Next Article in Journal
Nonlocal Symmetry, Painlevé Integrable and Interaction Solutions for CKdV Equations
Next Article in Special Issue
Local Structure and Dynamics of Functional Materials Studied by X-ray Absorption Fine Structure
Previous Article in Journal
Particle–Antiparticle Asymmetry in Relativistic Deformed Kinematics
Previous Article in Special Issue
Extracting Local Symmetry of Mono-Atomic Systems from Extended X-ray Absorption Fine Structure Using Deep Neural Networks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Base Catalysis of Sodium Salts of [Ta6−xNbxO19]8− Mixed-Oxide Clusters

1
Department of Chemistry, Graduate School of Science, Tokyo Metropolitan University, 1-1 Minami-Osawa, Hachioji, Tokyo 192-0397, Japan
2
Elements Strategy Initiative for Catalysts & Batteries (ESICB), Kyoto University, 1-30 Goryo-Ohara, Nishikyo-ku, Kyoto 615-8245, Japan
3
Laboratory for Materials and Structures, Tokyo Institute of Technology, 4259 Nagatsuta-cho, Midori-ku, Yokohama, Kanagawa 226-8503, Japan
4
Precursory Research for Embryonic Science and Technology (PRESTO), Japan Science and Technology Agency (JST), Tokyo 102-0076, Japan
*
Author to whom correspondence should be addressed.
Symmetry 2021, 13(7), 1267; https://doi.org/10.3390/sym13071267
Submission received: 31 May 2021 / Revised: 30 June 2021 / Accepted: 4 July 2021 / Published: 15 July 2021
(This article belongs to the Special Issue X-ray Absorption Fine Structure and Symmetry)

Abstract

:
The solid base catalysis of sodium salts of Lindqvist-type metal oxide clusters was investigated using a Knoevenagel condensation reaction. We successfully synthesized the sodium salts of Ta and Nb mixed-oxide clusters Na8−nHn[(Ta6−xNbx)O19]·15H2O (Na-Ta6−xNbx, n = 0, 1, x = 0–6) and found them to exhibit activity for proton abstraction from nitrile substrates with a pKa value of 23.8, which is comparable to that of the conventional solid base MgO. The Ta-rich Na-Ta6 and Na-Ta4Nb2 exhibited high activity among Ta and Nb mixed-oxide clusters. Synchrotron X-ray diffraction (SXRD) measurements, Fourier-transform infrared (FT-IR) spectroscopy, and X-ray absorption spectroscopy (XAS) revealed the structure of Na-Ta6−xNbx: (1) The crystal structure changed from Na7H[M6O19]·15H2O to Na8[M6O19]·15H2O (M = Ta or Nb) by the anisotropic expansion of the unit cell with an increase in Ta content; (2) Highly symmetrical Lindqvist [Ta6−xNbxO19]8− was generated in Na-Ta4Nb2 and Na-Ta6 because of the symmetrical association of Na+ ions with [Ta6−xNbxO19]8− in the structure. DFT calculation revealed that the Lindqvist structures with high symmetry have large NBO charges on surface oxygen species, which are strongly related to base catalytic activity, whereas the composition hardly affects the NBO charges. The above results showed that the Brønsted base catalysis was sensitive to the symmetry of the Lindqvist [Ta6−xNbxO19]8− structure. These findings contribute to the design of solid base catalysts composed of anionic metal oxide clusters with alkaline-metal cations.

1. Introduction

Metal oxide clusters (MOCs), referred to as polyoxometalates, are well known as super acid catalysts [1,2,3,4], photocatalysts [5,6], and redox catalysts [7,8]. Especially, biomass conversion using the acidity and redox property is one of the attractive applications of MOCs [9,10]. Recently, the base catalysis of MOCs attracted substantial attention because negatively charged MOCs proved to have base catalytic properties [11,12,13]. Mizuno and co-workers reported that monooxometalate [WO4]2− [11,12] and defective Keggin-type MOC [γ-H2GeW10O36]6− [13] worked as homogeneous base catalysts. The density functional theory (DFT) revealed that the basicity of MOCs is related to the negative charges of the oxygen atoms within them. Thus, they achieved the strong basicity of MOCs by increasing the negative charge of the clusters with lacunary Keggin-type structures. Our group and Ge et al. found that group V (Nb, Ta) MOCs such as [Nb6O19]8− [14,15], [Ta6O19]8− [14], [Nb10O28]6− [16,17,18,19], and [SiNb12O40]16− [20] showed stronger Brønsted and Lewis base catalysis than group VI (Mo, W) MOCs for Knoevenagel condensation and CO2 fixation reactions, respectively. Among them, [Ta6O19]8− exhibited the strongest Brønsted base properties being able to abstract protons from nitriles with a pKa value of 23.8 in a homogeneous catalytic system because [Ta6O19]8− has negatively charged oxygen atoms, as calculated by DFT [14]. These MOCs also work as base catalysts in solid form. Zhao et al. reported that solid Na8H[PW9O34] salt was active for the Knoevenagel condensation reactions of benzaldehyde with malononitrile (pKa = 11.1) and ethyl cyanoacetate (pKa = 13.1) [21]. Similar base strength was obtained for Na16[SiNb12O40xH2O [20] and K7H[Nb6O19]·13H2O salts [15]. However, to date, the base strength of solid MOCs has not been investigated. It could be considered that oxygen atoms with high electron density should be designed to form strong basic sites. Besides, heteroatom doping is one possible way of tuning the electron density of surface oxygen atoms of MOCs [22,23].
Here, we synthesized the sodium salts of Ta–Nb MOCs with a Lindqvist structure Na8nH[Ta6xNbxO19]·15H2O (Na-Ta6−xNbx, n = 0, 1, x = 0−6) and investigated their base strength using a Knoevenagel condensation reaction, which includes the proton abstraction from nitriles with various pKa-value C–H bonding [24,25]. We found that the alkali-metal MOC salts showed superior Brønsted base catalytic properties compared with the conventional metal oxide base catalysts. Interestingly, Brønsted basic activity of Na-Ta6−xNbx was not in the order of negative charge of the oxygen atoms of free [Ta6−xNbxO19]8− clusters, which is one of the significant factors for the base strength [13]. The structural analysis revealed that the symmetry of the MO6 unit in the Lindqvist [Ta6−xNbxO19]8− was affected by the packing of Na+ in the crystal and that Na-Ta6 and Na-Ta4Nb2 had high MO6 octahedral symmetry among Na-Ta6−xNbx. In addition, we found that the symmetry of the MO6 unit in the Lindqvist [Ta6−xNbxO19]8− was strongly related to the base catalytic activity. These findings contribute to the design of solid base catalysts composed of anionic metal oxide clusters with alkaline-metal cations.

2. Materials and Methods

The Na8nH[Ta6xNbxO19]·15H2O (Na-Ta6−xNbx, n = 0, 1, x = 0, 2, 3, 4, and 6) were fabricated by the combination of the modified solid-state reaction and the dissolution-precipitation method [26]. The mixture of M2O5 (M = Ta or Nb), Na2C2O4, and (NH2)2CO at a molar ratio of (Ta + Nb):Na:(NH2)2CO of 1:1:4 was calcined at 773 K for 4 h and then 1373 K for 4 h. The obtained powder with a Na3MO4 structure was treated with NaOH solution. The resulting white precipitate was centrifugated and washed (Na-Ta6−xNbx). NaTaO3 and NaNbO3 were synthesized by a modified solid-state reaction [26] in which the mixture of Ta2O5 (or Nb2O5), Na2C2O4, and (NH2)2CO at a molar ratio of M (M = Ta or Nb):Na:(NH2)2CO of 1:1:4 was calcined at 773 K for 4 h [27]. MgO was supplied from the Catalytic Society of Japan (JRC-MGO-4). The high-resolution synchrotron XRD (SXRD) measurements were performed for Na-Ta6−xNbx using the large Debye–Scherrer camera equipped at beamline BL02B2 in SPring-8. Incident beams were monochromatized to 0.77531 Å. Powder samples were loaded into Pyrex capillaries with an i.d. of 0.2 mm. The sealed capillary was rotated during the measurements to reduce the effect of the preferred orientation of crystallites. The obtained SXRD data were analyzed by the RIETAN-FP program [28]. The structure model was displayed using the VESTA program [29]. The structures of MOCs in solid form were investigated using a Fourier-transform infrared spectrometer (JASCO, FT/IR-4600) equipped with an attenuated total reflection accessory (JASCO, ATR PRO ONE). The local structures of Na-Ta6−xNbx were investigated by Ta L3- and L1-, and Nb K-edges X-ray absorption spectroscopy (XAS) performed at BL01B1 in SPring-8. The spectra were recorded in transmittance mode using ionization chambers at room temperature. Si(111) and Si(311) double-crystal monochromators were used to obtain the incident X-ray beam for Ta L3- and L1-edges X-ray absorption fine structure (XAFS) and Nb-K edge XAFS spectra, respectively. XAFS spectra were analyzed using xTunes software [30]. The k3-weighted χ spectra were extracted as the extended X-ray absorption fine structure (EXAFS) after normalization. The EXAFS spectra in the k range 3.0–17.0 Å−1 were Fourier-transformed into r space to obtain FT-EXAFS spectra.
The base catalytic properties of the solid catalysts were investigated using Knoevenagel condensation reaction. Benzaldehyde (1 mmol), nitrile [1 mmol: a, ethyl cyanoacetate (pKa = 13.1); b, 4-cyanophenylacetonitrile (pKa = 16.0); c, (3-chlorophenyl)acetonitrile (pKa = 19.5); d, phenylacetonitrile (pKa = 21.9); e, (4-methoxyphenyl)acetonitrile (pKa = 23.8)], and biphenyl (internal standard, 0.1 mmol) were dissolved into dimethyl sulfoxide (2 mL). The Knoevenagel condensation reactions were started by the addition of alkali-metal MOC solid catalysts (5 μmol as MOC) into the reaction solution at 303 K or 343 K for 24 h. Metal oxide catalysts (5 mg) were activated by calcination at 473 K or 673 K prior to the reaction.
The DFT calculations and natural bonding orbital (NBO) analysis of the MOCs were performed by using the Gaussian 09 program, in accordance with the literature [14]. The optimized structures of the [Ta6−xNbxO19]8− were calculated by B3LYP incorporating the solvation effect of DMSO using PCM (dielectric constant = 46.826). We employed the LANL2DZ basis sets for Ta and Nb atoms and 6-31++G(d) basis sets for O atom to investigate the effect of the composition of the clusters on the NBO charge of O atoms. To determine the effect of structural distortion on the NBO charge, the NBO analysis was conducted using the crystal structures of Na-Nb6 and Na-Ta6 without structural optimization by B3LYP using basis sets of LANL2DZ for Ta and Nb atoms and 6-31++G(d) for O atom.

3. Results and Discussion

Figure 1 shows the catalytic activities of Na-Ta6 and Na-Nb6 MOCs for the Knoevenagel condensation reaction (Scheme 1), which enables C–C bond formation via proton abstraction from nitrile by the base catalysts. The stronger base catalysts show activities for this reaction, with nitriles having higher pKa values. Both Na-Ta6 and Na-Nb6 showed catalytic activities for the Knoevenagel condensation reactions with a (pKa = 13.1) and b (pKa = 16.0) at 303 K. Na-Ta6 and Na-Nb6 also exhibited base catalytic activities for c (pKa = 19.5), d (pKa = 21.9), and e (pKa = 23.8) at 343 K, although the activities gradually decreased with increasing pKa values of nitriles. The activity of Na-Ta6 for all substrates was higher than that of Na-Nb6. Figure 2 summarizes the base catalytic activities of Na-Ta6−xNbx mixed-oxide MOCs, and common solid base metal oxides in the Knoevenagel condensation reaction with d (pKa = 21.9) at 343 K. The activity over Na-Ta6−xNbx MOCs was in the order of Na-Ta6 ≈ Na-Ta4Nb2 > Na-Ta3Nb3 ≈ Na-Ta2Nb4 > Na-Nb6. The base catalytic activities of Ta2O5, Nb2O5, and perovskite NaTaO3 and NaNbO3 were negligible. Therefore, the cluster structures are important for strong basicity. Among the metal oxide catalysts, activated MgO, treated at 673 K under vacuum conditions for 1 h, showed the base catalytic activity, whereas low activity was obtained over MgO treated at 473 K. This is because MgO has strong base sites of pKa ≥ 23 [31], which are passivated by the adsorption of H2O and CO2. Therefore, Na-Ta6−xNbx salts acted as solid base catalysts without pretreatment, and their strong basicity is comparable to that of activated MgO. Recently, we have reported that the surrounding environment of MOCs is significant for base catalysis because the Brønsted base catalytic activity was suppressed by blocking the base sites on MOCs with counter cations [18]. Next, the structural analysis of Na-Ta6−xNbx was performed.
Figure 3A shows the SXRD patterns of Na-Ta6−xNbx. All patterns can be indexed on an orthorhombic unit cell with a space group of Pmnn, which is consistent with the previous report of Na8[Ta6O19]·15H2O and Na7H[Nb6O19]·14H2O [32]. Here, the Na content in the unit cell is different between Na-Ta6 and Na-Nb6 MOCs, but the atomic positions are almost similar despite the additional Na-site in Na-Ta6 (Figure 4). No impurity phase was detected in all patterns. The lattice parameters obtained by Le Bail analysis are plotted against x in Figure 3B. The compounds show an anisotropic expansion of the lattice, where the a and c axes increase with x while the b axis decrease. The monotonic changes in the lattice suggest the successful formation of a solid solution. However, the lattice parameters deviate from Vegard’s law. In addition, the volume increases in the range of 0 ≤ x ≤ 4, whereas the volume decrease in the high x region. These behaviors possibly relate to the change in the Na content during the change in the Ta/Nb content. Since the structure contains two different Ta/Nb sites (Wyckoff positions of 4g and 8g sites), it is interesting to consider the site selectivity of the Ta/Nb positions. Thus, we attempted to carry out Rietveld refinements for the solid solution using Na7H[Nb6O19]·15H2O structure [32]. However, meaningful selectivity cannot be detected in all compositions.
The local structure of the Lindqvist M6O19 unit was investigated using FT-IR and XAS. Figure 5 shows the FT-IR spectra of Na-Ta6−xNbx in a region of metal-oxygen stretching patterns. The FT-IR spectrum of Na-Nb6 showed the characteristic absorption bands of the terminal Nb–Ot and bridging Nb–Ob–Nb vibration of the Lindqvist hexaniobate structure [33,34]. As for Na-Ta6, since Ta–Ot in Na8[Ta6O19]·15H2O (1.81 Å for Ta1–O4 and Ta2–O2) is longer than Nb–Ot in Na7H[Nb6O19]·15H2O (1.77 Å for Nb1–O4 and 1.79 Å for Nb2–O2), the absorption band of the Ta–Ot vibration in Na-Ta6 appeared at a lower frequency than that of Nb–Ot in Na-Nb6. The absorption band of M–Ot (M = Ta and Nb) in Na-Ta4Nb2 was also observed at a lower frequency than that of Nb–Ot in Na-Nb6, whereas those of Na-Ta2Nb4 and Na-Ta3Nb3 hardly shifted, being consistent with the deviation from Vegard’s law in lattice parameters, as shown in Figure 3B. Thus, the surroundings of the Lindqvist unit of Na-Ta4Nb2 are similar to those of Na-Ta6.
X-ray absorption near edge structure (XANES) spectra supported the structural change in the Lindqvist M6O19 unit of Na-Ta6−xNbx with a Ta-rich composition. The pre-edge-peak intensity of XANES spectra of group V, VI, and VII compounds is known as an indicator of the coordination number, d-electron number, and coordination symmetry [35,36,37]. Figure 6A shows Ta L1-edge XANES spectra of Na-Ta6−xNbx. Compared with Na-Ta6, Nb-rich Na-Ta6−xNbx MOCs exhibited high pre-edge peak intensity, which is attributed to the electronic transition from the 2p to 5d–6p hybrid orbital formed by the structural distortion from the octahedral (Oh) symmetry of the TaO6 unit. These results indicate that the Oh symmetry of TaO6 in the Lindqvist M6O19 of Na-Ta4Nb2 and Na-Ta6 is higher than that of Na-Ta3Nb3 and Na-Ta2Nb4. Nb K-edge XANES spectra exhibited a similar decrease in pre-edge-peak intensity (electronic transition from 2p to 4d–5p hybrid orbital) with a Ta-rich composition (Figure 6B). Thus, Na-Ta4Nb2 has a NbO6 unit with higher Oh symmetry in the Lindqvist structure than Na-Nb6. Ta L3- and Nb K-edges extended X-ray absorption fine structure (EXAFS) oscillations of Na-Ta6−xNbx robustly indicated the formation of solid solution. (Figure 6C,D). In Ta L3-edge Fourier-transformed EXAFS of Na-Ta6−xNbx (Figure 6E), the two peaks assigned to Ta–Nb and Ta–Ta bonds appeared in Na-Ta4Nb2, Na-Ta3Nb3, and Na-Ta2Nb4, and peak intensity of Ta–Ta increased with increasing Ta content. Similar results were obtained for Nb K-edge FT-EXAFS (Figure 6F) that the Nb–Ta peaks appeared at a long distance, and this intensity gradually increased with an increasing amount of Ta.
The structural analyses revealed that: (1) Na-Ta6−xNbx was composed of [Ta6−xNbxO19]8− mixed-oxide clusters; (2) Na-Ta6 and Na-Ta4Nb2 had Na8[Ta6−xNbxO19]·15H2O crystal structure whereas others were Na7H[Ta6−xNbxO19]·15H2O; (3) Oh symmetry of MO6 in Na-Ta6 and Na-Ta4Nb2 was higher than that in other Na-Ta6−xNbx. It was reported that the Brønsted base strength of MOCs depends on the electron density of the surface oxygen atoms [13,14]. The cluster compositions and structural distortion are expected to affect the electronic state of surface O atoms. First, we estimated the NBO charges of surface oxygen atoms of free [Ta6−xNbxO19]8− clusters by DFT calculations to investigate the effect of cluster composition on the NBO charge. The DFT calculations revealed that the order of the average NBO charge of surface oxygen atoms was [Ta6O19]8− (−1.01) > [Ta4Nb2O19]8− (−0.98) > [Ta3Nb3O19]8− (−0.97) > [Ta2Nb4O19]8− (−0.95) > [Nb6O19]8− (−0.93) as shown in Figure 7A. However, the NBO charge of Ta-coordinated O atoms hardly changed with Nb substitution both for terminal and bridging O atoms. Thus, Brønsted base activity of Na-Ta6−xNbx salts could not be explained by the NBO charge of the surface oxygen atoms in the free [Ta6−xNbxO19]8− clusters.
Next, we investigated the effect of structural distortion of MO6 in the clusters on the NBO charge by DFT calculation because XAFS analysis indicated that the Oh symmetry of MO6 in Na-Ta4Nb2 and Na-Ta6 was higher than that in other Na-Ta6−xNbx, which is induced by the symmetrical locations of Na ions in the crystal structure. The [M6O19]8− geometries in Na7H[M6O19]·15H2O as low Oh MO6 and Na8[M6O19]·15H2O as high Oh MO6 were used in this study (M = Ta or Nb). The average NBO charge of surface oxygen atoms of Lindqvist [M6O19]8− in Na8[M6O19]·15H2O was higher than those in Na7H[M6O19]·15H2O (Figure 7B). This result showed that the high Oh symmetry of MO6 in the Lindqvist structure with the symmetrical association of Na+ ions induces highly negatively charged O atoms. Besides, in the case of low Oh MO6, the cationic charge of M atoms decreases due to the interaction between 2p orbitals of the O atoms and d orbitals of M atoms. The higher base catalytic activities of Na-Ta4Nb2 and Na-Ta6 could be explained by the large NBO charges of surface O atoms induced by the high Oh symmetry of MO6. The above results showed that the Brønsted base catalysis was sensitive to the symmetry of the MO6 unit in the Lindqvist [Ta6−xNbxO19]8− structure.

4. Conclusions

Ta and Nb mixed-oxide clusters Na8nH[Ta6xNbxO19]·15H2O (Na-Ta6−xNbx, n = 0, 1, x = 0–6) in solid form exhibited activities for proton abstraction from nitrile substrates with a pKa value up to 23.8, which is comparable to that of the conventional solid base MgO. The Ta-rich Na-Ta6 and Na-Ta4Nb2 showed high activities among Ta and Nb mixed-oxide clusters. The structural analysis using SXRD, FT-IR, and XAS revealed that Na-Ta4Nb2 has a similar structure to Na8[Ta6O19]·15H2O, with a highly symmetrical MO6 in Lindqvist structure (M = Ta or Nb) compared with that of Nb6O19·in Na7H[Nb6O19]·15H2O. On the other hand, Na-Ta3Nb3 and Na-Ta2Nb4 have the same crystal structure to Na7H[Nb6O19]·15H2O. In a series of Na-Ta6−xNbx MOCs, the crystal structure had a large impact on the activity. NBO analysis using MO6 unit of Na8[Ta6O19]·15H2O and Na7H[Nb6O19]·15H2O revealed that high Oh symmetry of MO6 in Lindqvist structure induces the negative charges on O atoms. For the design of solid base catalytic sites, it is essential to control the symmetry of the MO6 unit in the MOC by using cation packing in the crystal structure.

Author Contributions

S.K. and S.Y. designed this study. M.T., K.S. (Kanako Shibata), Y.F. and J.H. synthesized and characterized catalysts and carried out the catalytic tests. K.S. (Kazuki Shibusawa) and N.N. carried out the DFT calculations and analyzed the NBO charges of the catalysts. T.Y. measured and analyzed the SXRD. S.K., M.T. and S.Y. measured and analyzed the XAFS spectra. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by the Elements Strategy Initiative for Catalysts & Batteries (ESICB), Japan Society for the Promotion of Science (JSPS) KAKENHI (Nos. 18K18982, 20K22467 and 21H01718), and Yazaki memorial foundation for science and technology.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

The synchrotron radiation experiment was performed at BL01B1 and BL02B2 in SPring-8 under the approval of the Japan Synchrotron Radiation Research Institute (JASRI) as 2019B1166, 2020A0715, 2020A1068, 2020A1219, and 2020A1851.

Conflicts of Interest

There is no conflict of interest.

References

  1. Olah, G.A.; Sommer, J.; Namanworth, E. Stable carbonium ions. XLI: Protonated aliphatic alcohols and their cleavage to carbonium ions. J. Am. Chem. Soc. 1967, 89, 3576–3581. [Google Scholar] [CrossRef]
  2. Birchall, T.; Gillespie, R. Nuclear magnetic resonance studies of the protonation of weak bases in fluorosulfuric acid: V. Ketones, carboxylic acids, and some other oxygen bases. Can. J. Chem. 1965, 43, 1045–1051. [Google Scholar] [CrossRef]
  3. Mizuno, N.; Misono, M. Heterogeneous catalysis. Chem. Rev. 1998, 98, 199–218. [Google Scholar] [CrossRef]
  4. Kozhevnikov, I.V. Catalysis by heteropoly acids and multicomponent polyoxometalates in liquid-phase reactions. Chem. Rev. 1998, 98, 171–198. [Google Scholar] [CrossRef] [PubMed]
  5. Suzuki, K.; Mizuno, N.; Yamaguchi, K. Polyoxometalate photocatalysis for liquid-phase selective organic functional group transformations. ACS Catal. 2018, 8, 10809–10825. [Google Scholar] [CrossRef]
  6. Iwase, Y.; Tomita, O.; Naito, H.; Higashi, M.; Abe, R. Molybdenum-substituted polyoxometalate as stable shuttle redox mediator for visible light driven Z-scheme water splitting system. J. Photochem. Photobiol. A Chem. 2018, 356, 347–354. [Google Scholar] [CrossRef]
  7. Sadakane, M.; Steckhan, E. Electrochemical properties of polyoxometalates as electrocatalysts. Chem. Rev. 1998, 98, 219–238. [Google Scholar] [CrossRef] [PubMed]
  8. Wang, S.S.; Yang, G.Y. Recent advances in polyoxometalate–catalyzed reactions. Chem. Rev. 2015, 115, 4893–4962. [Google Scholar] [CrossRef] [PubMed]
  9. Albert, J.; Mehler, J.; Tucher, J.; Kastner, K.; Streb, C. One-step synthesizable Lindqvist−isopolyoxometalates as promising new catalysts for selective conversion of glucose as a model substrate for lignocellulosic biomass to formic acid. ChemistrySelect 2016, 1, 2889–2894. [Google Scholar] [CrossRef]
  10. Cheng, F.; Wang, H.; Rogers, R.D. Oxygen enhances polyoxometalate-based catalytic dissolution and delignification of woody biomass in ionic liquids. ACS Sustain. Chem. Eng. 2014, 2, 2859–2865. [Google Scholar] [CrossRef]
  11. Kimura, T.; Kamata, K.; Mizuno, N. A bifunctional tungstate catalyst for chemical fixation of CO2 at atmospheric pressure. Angew. Chem. Int. Ed. Engl. 2012, 51, 6700–6703. [Google Scholar] [CrossRef]
  12. Kimura, T.; Sunaba, H.; Kamata, K.; Mizuno, N. Efficient [WO4]2−-catalyzed chemical fixation of carbon dioxide with 2-aminobenzonitriles to quinazoline-2,4(1H,3H)-diones. Inorg. Chem. 2012, 51, 13001–13008. [Google Scholar] [CrossRef] [PubMed]
  13. Sugahara, K.; Kimura, T.; Kamata, K.; Yamaguchi, K.; Mizuno, N. A highly negatively charged gamma-Keggin germanodecatungstate efficient for Knoevenagel condensation. Chem. Commun. 2012, 48, 8422–8424. [Google Scholar] [CrossRef]
  14. Hayashi, S.; Sasaki, N.; Yamazoe, S.; Tsukuda, T. Superior base catalysis of group 5 hexametalates [M6O19]8– (M = Ta, Nb) over group 6 hexametalates [M6O19]2– (M = Mo, W). J. Phys. Chem. C 2018, 122, 29398–29404. [Google Scholar] [CrossRef]
  15. Xu, Q.; Niu, Y.; Wang, G.; Li, Y.; Zhao, Y.; Singh, V.; Niu, J.; Wang, J. Polyoxoniobates as a superior Lewis base efficiently catalyzed Knoevenagel condensation. Mol. Catal. 2018, 453, 93–99. [Google Scholar] [CrossRef]
  16. Hayashi, S.; Yamazoe, S.; Koyasu, K.; Tsukuda, T. Application of group V polyoxometalate as an efficient base catalyst: A case study of decaniobate clusters. RSC Adv. 2016, 6, 16239–16242. [Google Scholar] [CrossRef] [Green Version]
  17. Hayashi, S.; Yamazoe, S.; Koyasu, K.; Tsukuda, T. Lewis base catalytic properties of [Nb10O28]6− for CO2 fixation to epoxide: Kinetic and theoretical studies. Chem. Asian J. 2017, 12, 1635–1640. [Google Scholar] [CrossRef] [PubMed]
  18. Hayashi, S.; Yamazoe, S.; Tsukuda, T. Base catalytic activity of [Nb10O28]6–: Effect of countercations. J. Phys. Chem. C 2020, 124, 10975–10980. [Google Scholar] [CrossRef]
  19. Gutierrez, L.F.; Nope, E.; Rojas, H.A.; Cubillos, J.A.; Sathicq, A.G.; Romanelli, G.P.; Martínez, J.J. New application of decaniobate salt as basic solid in the synthesis of 4H-pyrans by microwave assisted multicomponent reactions. Res. Chem. Intermed. 2018, 44, 5559–5568. [Google Scholar] [CrossRef]
  20. Ge, W.; Wang, X.; Zhang, L.; Du, L.; Zhou, Y.; Wang, J. Fully-occupied Keggin type polyoxometalate as solid base for catalyzing CO2 cycloaddition and Knoevenagel condensation. Catal. Sci. Technol. 2016, 6, 460–467. [Google Scholar] [CrossRef]
  21. Zhao, S.; Chen, Y.; Song, Y.-F. Tri-lacunary polyoxometalates of Na8H[PW9O34] as heterogeneous Lewis base catalysts for Knoevenagel condensation, cyanosilylation and the synthesis of benzoxazole derivatives. Appl. Catal. A 2014, 475, 140–146. [Google Scholar] [CrossRef]
  22. Simms, C.; Kondinski, A.; Parac-Vogt, T.N. Metal-addenda substitution in plenary polyoxometalates and in their modular transition metal analogues. Eur. J. Inorg. Chem. 2020, 2020, 2559–2572. [Google Scholar] [CrossRef]
  23. Anderson, T.M.; Rodriguez, M.A.; Stewart, T.A.; Bixler, J.N.; Xu, W.; Parise, J.B.; Nyman, M. Controlled sssembly of [Nb6–xWxO19](8–x)– (x = 0–4) Lindqvist ions with (amine)copper Complexes. Eur. J. Inorg. Chem. 2008, 2008, 3286–3294. [Google Scholar] [CrossRef]
  24. Isobe, K.; Hoshi, T.; Suzuki, T.; Hagiwara, H. Knoevenagel reaction in water catalyzed by amine supported on silica gel. Mol. Divers. 2005, 9, 317–320. [Google Scholar] [CrossRef] [PubMed]
  25. Tran, U.P.N.; Le, K.K.A.; Phan, N.T.S. Expanding applications of metal−organic frameworks: Zeolite imidazolate framework ZIF-8 as an efficient heterogeneous catalyst for the Knoevenagel reaction. ACS Catal. 2011, 1, 120–127. [Google Scholar] [CrossRef]
  26. Yamazoe, S.; Shibata, K.; Kato, K.; Wada, T. Needle-like NaNbO3 synthesis via Nb6O198− cluster using Na3NbO4 precursor by dissolution–precipitation method. Chem. Lett. 2013, 42, 380–382. [Google Scholar] [CrossRef] [Green Version]
  27. Fukada, M.; Shibata, K.; Imai, T.; Yamazoe, S.; Hosokawa, S.; Wada, T. Fabrication of lead-free piezoelectric NaNbO3 ceramics at low temperature using NaNbO3 nanoparticles synthesized by solvothermal method. J. Ceram. Soc. Jpn. 2013, 121, 116–119. [Google Scholar] [CrossRef] [Green Version]
  28. Izumi, F.; Momma, K. Three-dimensional visualization in powder diffraction. Solid State Phenom. 2007, 130, 15–20. [Google Scholar] [CrossRef]
  29. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  30. Asakura, H.; Yamazoe, S.; Misumi, T.; Fujita, A.; Tsukuda, T.; Tanaka, T. xTunes: A new XAS processing tool for detailed and on-the-fly analysis. Radiat. Phys. Chem. 2020, 175, 108270–108273. [Google Scholar] [CrossRef]
  31. Matsuhashi, H.; Nagashima, K.; Naijo, N.; Aritani, H. Surface base sites of MgO covered with Al2O3: XANES analysis of Al and Mg K-edges. Top. Catal. 2010, 53, 659–663. [Google Scholar] [CrossRef]
  32. Anderson, T.M.; Rodriguez, M.A.; Bonhomme, F.; Bixler, J.N.; Alam, T.M.; Nyman, M. An aqueous route to [Ta6O19]8− and solid-state studies of isostructural niobium and tantalum oxide complexes. Dalton Trans. 2007, 4517–4522. [Google Scholar] [CrossRef]
  33. Ma, P.T.; Chen, G.; Wang, G.; Wang, J.P. Cobalt–sandwiched Lindqvist hexaniobate dimer [Co(III)H5(Nb6O19)2]8−. Russ. J. Coord. Chem. 2011, 37, 772–775. [Google Scholar] [CrossRef]
  34. Besserguenev, A.V.; 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]
  35. Farges, F.; Brown, G.E.; Rehr, J. Ti K-edge XANES studies of Ti coordination and disorder in oxide compounds: Comparison between theory and experiment. Phys. Rev. B 1997, 56, 1809–1819. [Google Scholar] [CrossRef]
  36. Yamazoe, S.; Hitomi, Y.; Shishido, T.; Tanaka, T. XAFS study of tungsten L1-and L3-edges: Structural analysis of WO3 species loaded on TiO2 as a catalyst for photo-oxidation of NH3. J. Phys. Chem. C 2008, 112, 6869–6879. [Google Scholar] [CrossRef]
  37. Asakura, H.; Shishido, T.; Yamazoe, S.; Teramura, K.; Tanaka, T. Structural analysis of group V, VI, and VII metal compounds by XAFS. J. Phys. Chem. C 2011, 115, 23653–23663. [Google Scholar] [CrossRef]
Figure 1. The conversions of nitriles with various pKa in Knoevenagel condensation reaction for 24 h over Na-Ta6 (yellow) and Na-Nb6 (green). Benzaldehyde, 1 mmol; nitrile substrates ae, 1 mmol; catalyst, 5 μmol; solvent, DMSO 2 mL; temperature, 303 K for a and b or 343 K for ce; atmosphere, N2 1 atm.
Figure 1. The conversions of nitriles with various pKa in Knoevenagel condensation reaction for 24 h over Na-Ta6 (yellow) and Na-Nb6 (green). Benzaldehyde, 1 mmol; nitrile substrates ae, 1 mmol; catalyst, 5 μmol; solvent, DMSO 2 mL; temperature, 303 K for a and b or 343 K for ce; atmosphere, N2 1 atm.
Symmetry 13 01267 g001
Scheme 1. Knoevenagel condensation reaction of nitrile substrates with benzaldehyde.
Scheme 1. Knoevenagel condensation reaction of nitrile substrates with benzaldehyde.
Symmetry 13 01267 sch001
Figure 2. Conversions of d (pKa = 21.9) in Knoevenagel condensation reaction over Na-Ta6−xNbx and reference catalysts. Reaction conditions: benzaldehyde, 1 mmol; d, 1 mmol; catalyst, 5 μmol for Na-Ta6−xNbx or 5 mg for other metal oxides; solvent, DMSO 2 mL; temperature, 303 K; atmosphere, N2 1 atm. MgO was activated at 673 K under vacuum, whereas other metal-oxide catalysts were pretreated at 473 K.
Figure 2. Conversions of d (pKa = 21.9) in Knoevenagel condensation reaction over Na-Ta6−xNbx and reference catalysts. Reaction conditions: benzaldehyde, 1 mmol; d, 1 mmol; catalyst, 5 μmol for Na-Ta6−xNbx or 5 mg for other metal oxides; solvent, DMSO 2 mL; temperature, 303 K; atmosphere, N2 1 atm. MgO was activated at 673 K under vacuum, whereas other metal-oxide catalysts were pretreated at 473 K.
Symmetry 13 01267 g002
Figure 3. (A) SXRD pattens of Na-Ta6−xNbx with the corresponding reference patterns. (a) x = 0, (b) 2, (c) 3, (d) 4, and (e) 6. (B) Lattice parameters for Na-Ta6−xNbx (x = 0, 2, 3, 4, and 6). Right axis represents the unit cell volume.
Figure 3. (A) SXRD pattens of Na-Ta6−xNbx with the corresponding reference patterns. (a) x = 0, (b) 2, (c) 3, (d) 4, and (e) 6. (B) Lattice parameters for Na-Ta6−xNbx (x = 0, 2, 3, 4, and 6). Right axis represents the unit cell volume.
Symmetry 13 01267 g003
Figure 4. The Lindqvist-unit structure of (A) Ta6O19 in Na8[Ta6O19]·15H2O and (B) Nb6O19 in Na7H[Nb6O19]·15H2O with surrounding Na atoms within 2.6 Å. Color codes: yellow, Ta; green, Nb; red, O; purple, Na; pink, 25%-occupied Na.
Figure 4. The Lindqvist-unit structure of (A) Ta6O19 in Na8[Ta6O19]·15H2O and (B) Nb6O19 in Na7H[Nb6O19]·15H2O with surrounding Na atoms within 2.6 Å. Color codes: yellow, Ta; green, Nb; red, O; purple, Na; pink, 25%-occupied Na.
Symmetry 13 01267 g004
Figure 5. (A) FT-IR spectra of Na-Ta6−xNbx. (a) x = 0, (b) 2, (c) 3, (d) 4, and (e) 6, and (B) corresponding extended view.
Figure 5. (A) FT-IR spectra of Na-Ta6−xNbx. (a) x = 0, (b) 2, (c) 3, (d) 4, and (e) 6, and (B) corresponding extended view.
Symmetry 13 01267 g005
Figure 6. (A) Ta L1- and (B) Nb K-edge XANES spectra. (C) Ta L3- and (D) Nb K-edge EXAFS oscillations and (E) Ta L3- and (F) Nb K-edge Fourier transforms. (a) Na-Ta6, (b) Na-Ta4Nb2, (c) Na-Ta3Nb3, (d) Na-Ta2Nb4, (e) Na-Nb6.
Figure 6. (A) Ta L1- and (B) Nb K-edge XANES spectra. (C) Ta L3- and (D) Nb K-edge EXAFS oscillations and (E) Ta L3- and (F) Nb K-edge Fourier transforms. (a) Na-Ta6, (b) Na-Ta4Nb2, (c) Na-Ta3Nb3, (d) Na-Ta2Nb4, (e) Na-Nb6.
Symmetry 13 01267 g006
Figure 7. (A) Average NBO charges on O atoms in free [Ta6O19]8−, [Ta4Nb2O19]8−, [Ta3Nb3O19]8−, [Ta2Nb4O19]8−, and [Nb6O19]8− clusters. (B) Average NBO charges on O atoms in [M6O19]8− (M = Ta or Nb) unit of Na8[Ta6O19]·15H2O and Na7H[Nb6O19]·15H2O. Color codes: yellow, Ta-coordinated O atoms; green, Nb-coordinated O atoms; red, total O atoms. Terminal and bridging O atoms represent solid and dashed lines, respectively.
Figure 7. (A) Average NBO charges on O atoms in free [Ta6O19]8−, [Ta4Nb2O19]8−, [Ta3Nb3O19]8−, [Ta2Nb4O19]8−, and [Nb6O19]8− clusters. (B) Average NBO charges on O atoms in [M6O19]8− (M = Ta or Nb) unit of Na8[Ta6O19]·15H2O and Na7H[Nb6O19]·15H2O. Color codes: yellow, Ta-coordinated O atoms; green, Nb-coordinated O atoms; red, total O atoms. Terminal and bridging O atoms represent solid and dashed lines, respectively.
Symmetry 13 01267 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kikkawa, S.; Tsukada, M.; Shibata, K.; Fujiki, Y.; Shibusawa, K.; Hirayama, J.; Nakatani, N.; Yamamoto, T.; Yamazoe, S. Base Catalysis of Sodium Salts of [Ta6−xNbxO19]8− Mixed-Oxide Clusters. Symmetry 2021, 13, 1267. https://doi.org/10.3390/sym13071267

AMA Style

Kikkawa S, Tsukada M, Shibata K, Fujiki Y, Shibusawa K, Hirayama J, Nakatani N, Yamamoto T, Yamazoe S. Base Catalysis of Sodium Salts of [Ta6−xNbxO19]8− Mixed-Oxide Clusters. Symmetry. 2021; 13(7):1267. https://doi.org/10.3390/sym13071267

Chicago/Turabian Style

Kikkawa, Soichi, Mio Tsukada, Kanako Shibata, Yu Fujiki, Kazuki Shibusawa, Jun Hirayama, Naoki Nakatani, Takafumi Yamamoto, and Seiji Yamazoe. 2021. "Base Catalysis of Sodium Salts of [Ta6−xNbxO19]8− Mixed-Oxide Clusters" Symmetry 13, no. 7: 1267. https://doi.org/10.3390/sym13071267

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