Next Article in Journal
Functional Properties of Corn Byproduct-Based Emulsifier Prepared by Hydrothermal–Alkaline
Previous Article in Journal
Chiral Covalent-Organic Framework MDI-β-CD-Modified COF@SiO2 Core–Shell Composite for HPLC Enantioseparation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A {Co9}-Added Polyoxometalate for Efficient Visible-Light-Driven Hydrogen Evolution

MOE Key Laboratory of Cluster Science, School of Chemistry and Chemical Engineering, Beijing Institute of Technology, Beijing 102488, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(2), 664; https://doi.org/10.3390/molecules28020664
Submission received: 14 November 2022 / Revised: 11 December 2022 / Accepted: 6 January 2023 / Published: 9 January 2023
(This article belongs to the Section Inorganic Chemistry)

Abstract

:
A polyanion cluster H6Na8Cs3[Co93-OH)3(H2O)6(HPO4)2(B-α-PW9O34)3]Cl·40H2O (1) was made with the guidance of the lacunary directing strategy under hydrothermal conditions. Compound 1 was characterized by single-crystal X-ray diffraction, powder X-ray diffraction, and thermogravimetric analysis, respectively. Single-crystal X-ray diffraction analyses showed that 1 consists of three anions [B-α-PW9O34]9− and a cyclic cationic [Co93-OH)3(H2O)6]15+ and two anions HPO42−. Variable-magnetic properties indicate antiferromagnetic interactions in 1. Visible-light-driven hydrogen evolution tests demonstrated that 1 was an efficient water reduction catalyst with an H2 evolution rate of 1217.6 μmol h−1 g−1.

1. Introduction

Polyoxometalates (POMs) have attracted increasing attention due to their intriguing structural diversity [1] and diverse applications in drugs [2,3], magnetism [4,5], and catalysis [4,5,6,7,8,9,10,11]. Lacunary POMs as precursors can induce transition metal (TM) cations’ aggregation to form TM oxo-clusters under hydrothermal conditions, producing a class of TM-added POMs (TMAPs) with interesting physicochemical properties [12,13,14,15]. Up to now, hydrothermal synthesis was successfully used in creating a large number of TMAPs. Since 2007, we have reported a large number of organic–inorganic hybrid Ni6-added POMs under the guidance of the lacunary directing strategy, using the lacunary sites of tri-lacunary POMs fragments {XW9O34}n− (X = P/Si/Ge, n = 9, 10) as structure-directing agents (SDAs) to induce the TM-oxo clusters aggregation to form novel TMAPs [16,17,18,19,20,21].
POMs have unique redox properties and modifiable Lewis acidity but their inherent poor stability and tendency to aggregate in solution have hindered further development [20]. TMAPs are widely used in photocatalysis [22,23] and electrocatalysis [24] fields due to their physicochemical properties and reversible multi-electron transfer ability. TMAPs have been widely studied due to their unique structural diversity, magnetic properties, and catalytic properties. Taking the above into account, TMAPs have demonstrated a massive potential for application for visible-light-driven Hydrogen evolution as heterogeneous water reduction catalysts (WRCs) [14,15,25].
Co-containing complexes were applied in photocatalytic water-splitting reactions with good reactivity [26] and Co2+ can be induced by tri-lacunary POMs fragments {XW9O34}n−, thus improving the performance of Co-added POMs (CoAPs) as the catalytic active center in visible-light-driven H2 evolution. Herein, a polyanion (HPO4)2@-{Co9(PW9)3} cluster was made with the guidance of the lacunary directing strategy under hydrothermal conditions. Interestingly, the HPO4 groups of 1 originate from the central heteroatom of the tri-lacunary precursor [A-α-PW9O34]9− and assemble with the cyclic (HPO4)2@{Co9} cluster to form CoAP. Moreover, visible-light-driven H2 evolution tests have demonstrated that CoAP is an efficient WRC with the H2 evlution rate of 1217.6 μmol h−1 g−1, indicating that 1 is a potential heterogeneous WRC.

2. Experimental Section

2.1. General Procedure

All chemicals were acquired commercially and used without further purification. The synthetic method of the Na10[A-α-PW9O34]·7H2O ({PW9}) was derived from previous literature methods [27]. Powder X-ray diffraction (PXRD) patterns were recorded on a Bruker D8 Advance XRD diffractometer (Karlsruhe, Germany) with Cu Kα radiation (λ = 1.54056 Å). FT-IR spectra were measured by using a Thermo Scientific Nicolet iS10 FT-IR spectrometer (Waltham, MA, USA) in the range of 400–4000 cm−1 with KBr pallets. Thermogravimetric analyses were conducted under air flowing on a Mettler−Toledo TGA/DSC 1000 (Zurich, Switzerland) with a heating rate of 10 °C min−1 from 25 to 1000 °C. UV–Vis absorption spectra were obtained using a Shimadzu UV3600 spectrometer (Kyoto, Japan).

2.2. Synthesis of 1

Co(NO3)·6H2O (0.536 g, 1.84 mmol), {PW9} (0.597 g, 0.233 mmol), and CsCl (0.308 g, 1.829 mmol) were mixed in 8 mL H2O and 5 drops acetic acid. After stirring for 1 h, the mixed solution was transferred to a 25 mL Teflon-lined autoclave and kept at 80 °C for 24 h. When the reaction was complete and cooled to room temperature, the product was filtered and the hexagonal pink crystals were washed with distilled water in a yield of 0.052 g (0.058 mmol), a yield of 7.51% (based on {PW9}). IR data (KBr pellet, cm−1): 3430 (s), 1630 (m), 1030 (s), 933 (s), 883 (s), 806 (s), 721 (s).

2.3. X-ray Crystallography

A suitable crystal was selected and put on a Bruker APEX-II CCD diffractometer. This crystal was kept at 296.15 K during data collection. Using Olex2 [28], the structure was solved with the ShelXT [29] structure solution program using Intrinsic Phasing and refined with the ShelXL [30] refinement package using Least Squares minimization. The contribution of these disordered solvent molecules to the overall intensity data of the structure was treated using the SQUEEZE method [31] in PLATON (25 lattice water molecules in 1 are estimated by TGA). The crystallographic data structure refinement is listed in Table 1. Detailed crystallographic data of the title crystal were deposited on the Cambridge Crystallographic Data Centre with CCDC reference number 2213769 for 1. These data can be obtained free of charge via www.ccdc.cam.ac.uk/conts/retrieving.html accessed on 29 June 2022, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB21EZ, UK; Fax: +44-1223 336 033.

3. Result and Discussion

3.1. Structure of 1

The structure of 1 crystallized in the hexagonal space group P63/m. The structure contains one [Co93-OH)3(H2O)6(HPO4)2(B-α-PW9O34)3]16− (abb. as (HPO4)2@{Co9(PW9)3}, Figure 1a) cluster, three Cs+, eight Na+ and 15 water molecules. Bond valance sum (BVS) calculations show that the valence of W, Co, O, Cl and P atoms are +6, +2, −2, −1 and +5, respectively. The calculations reveal low values for O1 (1.299) and O3 (1.084) demonstrating that O1 and O3 further bond to H+ as OH groups (Table S1, Supplementary Materials). The polyanion (HPO4)2@{Co9(PW9)3} cluster can be viewed as consisting of three [Co3O23-OH)2(H2O)2(B-α-PW9O34)]9− (abb. as {Co3PW9}) units (Figure 1b) and two HPO4 tetrahedra. Co2+ are six-coordinate octahedral configurations with Co–O bond lengths ranging from 1.998 to 2.182 Å. In addition, two Co2+ (Co2, Co2A) have coordinated water molecules (O1W, O1WA), and the other Co2+ (Co3) carries a μ3-OH (O3, O3B) in the Co3O13 ({Co3}) cluster (Figure 1b). Three Co2+ share edges to form a triangular {Co3} cluster, and the {Co3} cluster cap on the [B-α-PW9O34]9− fragment to form the {Co3PW9} unit (Figure 1b,c). Three {Co3PW9} units are connected to each other through three μ3-OH forming a cyclic {Co9(PW9)3} unit. Two openings existed in the cyclic {Co9(PW9)3} unit, in which each consists of three O atoms forming a triangle with a side length of 2.501 Å and matching the size of the HPO42− tetrahedron (Figure 1d). The two HPO4 units are capped on both sides of the ring to form a cluster (HPO4)2@{Co9(PW9)3} (Figure 1a). Interestingly, [Cl@{Cs3(H2O)6}] (Figure S1, Supplementary Materials) fills the middle of the anion clusters as can be seen from the stacking diagram (Figure S2, Supplementary Materials).
From a topological point of view, 1 can be simplified as a three-connected 2D network (Figure 2a), in which (HPO4)2@{Co9(PW9)3} (red nodes) and [Cl@{Cs3(H2O)6}] (green nodes) can be simplified as three-connected (3-c) nodes (Figure 2b), respectively. Interestingly, the 2D layers are arranged in an –ABAB– manner along the c-axis (Figure 2c).

3.2. IR Spectrum and Optical Band Gap

FT-IR spectra were measured in the range of 400–4000 cm−1 with KBr pallets (Figure S3, Supplementary Materials). The strong peaks at around 3430 and 1630 cm−1 are dominated by the stretching and bending modes of the water molecules. The characteristic bands derived from the Keggin POM fragments in the 721–1030 cm−1 region. In detail, the peak at 1030 and 933 cm−1 for 1 is attributable to ν(P-Oa) and ν(W-Ot). The peak at 806, 721, and 692 cm−1 is attributable to ν(M-O-M) (M = W or Co), respectively. UV–Vis absorption and optical diffuse reflectance spectra of 1 were obtained in the wavelength range of 200–800 nm (Figure S4, Supplementary Materials). Absorption peaks of 260 nm and 535 nm correspond to the charge transfer between O→W and the d-d charge transfer between Co2+, respectively (illustration). The band-gap energy (Eg) is obtained by extrapolating the linear part of the rising curve to zero [32]. The Eg of 1 was estimated as 2.58 eV.

3.3. PXRD and Thermogravimetric Analysis

The pure phase was confirmed by powder X-ray diffraction (PXRD) patterns (Figure S5, Supplementary Materials). Thermogravimetric analysis was conducted under air flowing with the heating rate of 10 °C min−1 from 25 to 1000 °C (Figure S6, Supplementary Materials). The TG curve indicates that the weight of 1 shows one step loss of 10.25% (calcd 10.3%) from 25 to 600 °C, corresponding to the release of 40 lattice water molecules. six coordination water molecules, three hydroxy groups (1.5 H2O) and six protons (3 H2O) for 1, respectively.

3.4. Photocatalytic H2 Evolution Performance

Herein, the three-component system was examined: [Ir(ppy)2(dtbbpy)][PF6] (ppy = 2-phenylpyridine; dtbbpy = 5,5′-di-tert-butyl-2,2′-bipyridine) as a photosensitizer, triethanolamine (TEOA) as sacrificial electron donor, and 1 as water reduction catalyst (WRC). The device used for the photocatalytic reaction is a multichannel photocatalytic reactor (Beijing Perfectlight Technology, Beijing, China), which uses white LED lamp beads (400–800 nm, electric power: 10 W), and is continuously illuminated for 10 h at a temperature of 25 °C. The reaction system was deaerated for 20 min under an Ar/CH4 (4:1) atmosphere. H2 was analyzed using a GC979011 gas chromatograph (Fuli Instruments, Taizhou, China) with a TCD and a 5 Å molecular sieve column (3 m × 3 mm) with Ar as the carrier gas.
As shown in Figure 3a, the H2 production increases with increasing radiation time. With the increasing concentrations of TEOA (5 mM, 10 mM, 15 mM, 20 mM) in the reaction system, the production of H2 showed a trend of first increasing (11.88 μmol, 31.36 μmol, 36.53 μmol, respectively) and then decreasing (24.48 μmol) after 10 h of illumination. One possible explanation is that the addition of TEOA increases the pH value of the system, reducing the concentration of H+ in the system, and thereby reducing the catalytic effect. The H2 production (9.9 μmol, 36.5 μmol and 40.7 μmol, respectively) increases with the increasing quantity of catalyst (1, 3, 6 mg), and the H2 evolution rate was 745.6 μmol h−1 g−1, 1217.6 μmol h−1 g−1 and 745.6 μmol h−1 g−1, respectively (Figure 3b). In terms of catalytic efficiency, the maximum H2 production rate of 3 mg can reach 1217.6 μmol h−1 g−1. Considering the catalytic system is heterogeneous catalysis, with the increase in the amount of catalyst, part of the light is scattered and the catalytic efficiency decreases. Hence, efficient catalysis may be achieved using the appropriate amount of catalyst. When the concentration of the photosensitizer (0.05 mM, 0.1 mM, 0.2 mM) was changed under the optimal conditions, the amount of the H2 production was 20.6 μmol, 23.58 μmol and 36.53 μmol, respectively, and the amount of hydrogen production increased with the increase in the concentration of the photosensitizer (Figure 3c). The PXRD and FT-IR (Figure S7, Supplementary Materials) tests before and after the photocatalytic reaction proved that the structure of 1 was basically unchanged and had good catalytic stability.
For heterogeneous catalytic systems, the stability and recyclability of catalysts are always important factors. In the cycling experiments (Figure S8, Supplementary Materials), the yield of H2 was detected by GC analysis, and the catalyst was isolated by centrifugation and washed with CH3CN after two hours of reaction at a catalyst dosage of 9 mg. The experiments were repeated for three successive cycles, the amount of H2 production was similar in the first two cycles but decreased the third time. Such a decrease in activity might be ascribed to the loss of the catalyst sample during the isolating operation. The PXRD pattern and FT-IR spectra (Figure S9, Supplementary Materials) of the catalyst before and after cycling experiments proved that 1 remains unchanged, which further demonstrated the recyclability and stability of complex 1.
In conclusion, the best reaction conditions explored were: 1 (3 mg), TEOA (15 mM) and photosensitizer (0.2 mM), and the best H2 production efficiency can reach 1217.6 μmol h−1 g−1. This value is comparable to that of previously reported results for H2 evolution (Table S3, Supplementary Materials). Considering the factors of catalyst efficiency, 3 mg 1 was used for in-depth studies. As shown in Figure 3d, H2 was produced only when the photosensitizer [Ir(ppy)2(dtbbpy)]-[PF6] and TEOA electron sacrificial reagent were added with light irradiation, and a small amount of H2 was produced when the equimolar amount of {PW9} and CoCl2·6H2O was used as the control catalyst revealing the superiority of CoAPs in the visible-light-driven H2 evolution. According to the reported literature [22,33,34], a possible photocatalytic mechanism for this visible-light-driven H2 evolution process was proposed (Figure S10, Supplementary Materials). In visible-light-driven catalytic systems, the photosensitizer’s excited state can function as either an oxidant or reductant and thus can be quenched by an electron donor or an acceptor. A POMs-based catalyst and TEOA can oxidatively and reductively quench the excited state of [Ir(ppy)2(dtbbpy)]+ * in the visible-light-driven H2 evolution process catalyzed by 1.

3.5. Magnetic Properties

The variable-magnetic properties of 1 were measured in the temperature range of 2–300 K with a magnetic field of 5000 Oe. As shown in Figure 4a, the χmT value is 26.53 cm3 mol−1 K at 300 K. The χmT value decreases continuously with decreasing temperature. When the temperature drops below 50 K, the χmT value decreases rapidly. The χmT drops to 4.64 cm3 mol−1 K at 2 K. The above behaviors suggest antiferromagnetic interactions in 1. The magnetic susceptibility data of 1 are in accordance with the Curie–Weiss law χ = C/(Tθ) (Figure 4b) in the temperature range of 50–300 K with a constant of θ = −28.25 K, which is consistent with the overall antiferromagnetic interaction in 1.

4. Conclusions

In conclusion, a polyanion (HPO4)2@{Co9(PW9)3} cluster was made with the guidance of a lacunary directing strategy under hydrothermal conditions. The HPO4 of 1 originates from the central heteroatom of the tri-lacunary precursor [A-α-PW9O34]9− and assembles with a cyclic (HPO4)2@{Co9} cluster to form CoAP 1. Variable-magnetic properties indicate antiferromagnetic interactions in 1. Moreover, visible-light-driven H2 evolution tests demonstrated that 1 is an eco-friendly, efficient and stable WRC with an H2 evolution rate of 1217.6 μmol h−1 g−1. This work further proves that the lacunary directing strategy is an effective guideline for the synthesis of TMAPs under hydrothermal conditions. The photocatalytic performance showed that CoAP is an efficient WRC. In future work, the application of POMs in heterogeneous catalytic reactions is in progress.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28020664/s1, Table S1: Bond valence sum (BVS) calculations of all the W, P, Co, O, and Cl atoms in 1; Figure S1: View of the asymmetric unit of 1. Symmetry codes: A (x, y, 1.5-z), B (1-y, 1+x-y, z); Figure S2: (a) Polyhedral view of compound 1 along the c-axis; (b) Ball-and-stick model of the Cl@{Cs3(H2O)6} cluster; Figure S3: FT-IR spectra of 1; Figure S4: UV–Vis spectra of 1; Figure S5: Simulated and experimental powder X-ray diffraction patterns of 1; Figure S6: TG curves of 1; Figure S7: Stability test of 1 after catalysis: (a) PXRD patterns before and after photocatalysis; (b) FT-IR patterns before and after photocatalysis; Figure S8: Yield of H2 for 1 (9 mg) as a photocatalyst in three continuous runs; Figure S9: Stability test of 1 after three-time recycling: (a) The powder X-ray diffraction patterns after three-time recycling of catalyst 1; (b) The FT-IR spectra after three-time recycling of catalyst 1; Figure S10: Proposed mechanism for visible-light-driven H2 evolution tests by 1 with oxidative and reductive quenching mechanism; Table S2: Visible-light-driven H2 evolution catalyzed by different TMAP-based photocatalysts. References [14,15,35,36,37] are cited in the Supplementary Materials.

Author Contributions

Methodology, formal analysis, data curation, writing—original draft and preparation, Z.-W.W.; Conceptualization, writing—reviewing and editing, supervision, and funding acquisition: G.-Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant numbers 21831001, 21571016, 91122028 and 20725101.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Liu, J.X.; Zhang, X.B.; Li, Y.L.; Huang, S.L.; Yang, G.Y. Polyoxometalate Functionalized Architectures. Coord. Chem. Rev. 2020, 414, 213260. [Google Scholar] [CrossRef]
  2. Wang, S.S.; Yang, G.Y. Recent Advances in Polyoxometalate-Catalyzed Reactions. Chem. Rev. 2015, 115, 4893–4962. [Google Scholar] [CrossRef] [PubMed]
  3. Kong, X.; Chen, Q.; Wan, G.; Yang, Y.; Yu, H.; Li, B.; Wu, L. Hyaluronic Acid-Enwrapped Polyoxometalate Complex for Synergistic Near Infrared-II Photothermal/Chemo-Therapy and Chemodynamic Therapy. Biomacromolecules 2022, 23, 3752–3765. [Google Scholar] [CrossRef] [PubMed]
  4. Lydon, C.; Sabi, M.M.; Symes, M.D.; Long, D.L.; Murrie, M.; Yoshii, S.; Nojiri, H.; Cronin, L. Directed Assembly of Nanoscale Co(II)-Substituted {Co9[P2W15]3} and {Co14[P2W15]4} Polyoxometalates. Chem. Commun. 2012, 48, 9819–9821. [Google Scholar] [CrossRef] [PubMed]
  5. Ibrahim, M.; Lan, Y.; Bassil, B.S.; Xiang, Y.; Suchopar, A.; Powell, A.K.; Kortz, U. Hexadecacobalt(II)-Containing Polyoxometalate -Based Single-Molecule Magnet. Angew. Chem. Int. Ed. 2011, 50, 4708–4711. [Google Scholar] [CrossRef]
  6. Chen, X.; Zhang, G.; Li, B.; Wu, L. An Integrated Giant Polyoxometalate Complex for Photothermally Enhanced Catalytic Oxidation. Sci. Adv. 2021, 7, eabf8413. [Google Scholar] [CrossRef] [PubMed]
  7. Chen, X.; Yang, A.; Wang, G.; Wei, M.; Liu, N.; Li, B.; Wu, L. Reinforced Catalytic Oxidation of Polyoxometalate@charge Transfer Complex by On-Site Heating from Photothermal Conversion. Chem. Eng. J. 2022, 446, 137134. [Google Scholar] [CrossRef]
  8. Miras, H.N.; Yan, J.; Long, D.-L.; Cronin, L. Engineering Polyoxometalates with Emergent Properties. Chem. Soc. Rev. 2012, 41, 7403–7430. [Google Scholar] [CrossRef]
  9. Hill, C.L. Introduction:  PolyoxometalatesMulticomponent Molecular Vehicles to Probe Fundamental Issues and Practical Problems. Chem. Rev. 1998, 98, 1–2. [Google Scholar] [CrossRef] [Green Version]
  10. Chen, W.; Li, H.; Song, J.; Zhao, Y.; Ma, P.; Niu, J.; Wang, J. Binuclear Ru(III)-Containing Polyoxometalate with Efficient Photocatalytic Activity for Oxidative Coupling of Amines to Imines. Inorg. Chem. 2022, 61, 2076–2085. [Google Scholar] [CrossRef]
  11. Yang, L.; Zhang, Z.; Zhang, C.N.; Li, S.; Liu, G.C.; Wang, X.L. An excellent multifunctional photocatalyst with a polyoxometalate-viologen framework for CEES oxidation, Cr(VI) reduction and dye decolorization under different light regimes. Inorg. Chem. Front. 2022, 9, 4824–4833. [Google Scholar] [CrossRef]
  12. Xu, B.; Xu, Q.; Wang, Q.; Liu, Z.; Zhao, R.; Li, D.; Ma, P.; Wang, J.; Niu, J. A Copper-Containing Polyoxometalate-Based Metal–Organic Framework as an Efficient Catalyst for Selective Catalytic Oxidation of Alkylbenzenes. Inorg. Chem. 2021, 60, 4792–4799. [Google Scholar] [CrossRef] [PubMed]
  13. Patel, A.; Narkhede, N.; Singh, S.; Pathan, S. Keggin-Type Lacunary and Transition Metal Substituted Polyoxometalates as Heterogeneous Catalysts: A Recent Progress. Catal. Rev. 2016, 58, 337–370. [Google Scholar] [CrossRef]
  14. Wang, S.S.; Kong, X.Y.; Wu, W.; Wu, X.Y.; Cai, S.; Lu, C.Z. Synergic Coordination of Multicomponents for the Formation of a {Ni30} Cluster Substituted Polyoxometalate and Its In-Situ Assembly. Inorg. Chem. Front. 2022, 9, 4350–4358. [Google Scholar] [CrossRef]
  15. Wang, Z.W.; Zhao, Q.; Chen, C.A.; Sun, J.J.; Lv, H.; Yang, G.Y. Chiral {Ni6PW9} Cluster–Organic Framework: Synthesis, Structure, and Properties. Inorg. Chem. 2022, 61, 7477–7483. [Google Scholar] [CrossRef]
  16. Zheng, S.T.; Yuan, D.Q.; Jia, H.P.; Zhang, J.; Yang, G.Y. Combination between Lacunary Polyoxometalates and High-Nuclear Transition Metal Clusters under Hydrothermal Conditions: I. from Isolated Cluster to 1-D Chain. Chem. Commun. 2007, 18, 1858–1860. [Google Scholar] [CrossRef]
  17. Zheng, S.T.; Zhang, J.; Clemente-Juan, J.M.; Yuan, D.Q.; Yang, G.Y. Poly(Polyoxotungstate)s with 20 Nickel Centers: From Nanoclusters to One-Dimensional Chains. Angew. Chem. Int. Ed. 2009, 48, 7176–7179. [Google Scholar] [CrossRef]
  18. Huang, L.; Zhang, J.; Cheng, L.; Yang, G.Y. Poly(Polyoxometalate)s Assembled by {Ni6PW9} Units: From Ring-Shaped Ni24-Tetramers to Rod-Shaped Ni40-Octamers. Chem. Commun. 2012, 48, 9658–9660. [Google Scholar] [CrossRef]
  19. Li, H.L.; Zhang, M.; Lian, C.; Lang, Z.L.; Lv, H.J.; Yang, G.Y. Ring-Shaped Polyoxometalate Built by {Mn4PW9} and PO4 Units for Efficient Visible-Light-Driven Hydrogen Evolution. CCS Chem. 2021, 3, 2095–2103. [Google Scholar] [CrossRef]
  20. Chen, Y.; Guo, Z.W.; Li, X.X.; Zheng, S.T.; Yang, G.Y. Multicomponent Cooperative Assembly of Nanoscale Boron-Rich Polyoxotungstates with 22 and 30 Boron Atomsa. CCS Chem. 2021, 4, 1305–1314. [Google Scholar] [CrossRef]
  21. Huang, L.; Wang, S.S.; Zhao, J.W.; Cheng, L.; Yang, G.Y. Synergistic Combination of Multi-ZrIV Cations and Lacunary Keggin Germanotungstates Leading to a Gigantic Zr24-Cluster-Substituted Polyoxometalate. J. Am. Chem. Soc. 2014, 136, 7637–7642. [Google Scholar] [CrossRef] [PubMed]
  22. Lv, H.; Guo, W.; Wu, K.; Chen, Z.; Bacsa, J.; Musaev, D.G.; Geletii, Y.V.; Lauinger, S.M.; Lian, T.; Hill, C.L. A Noble-Metal-Free, Tetra-Nickel Polyoxotungstate Catalyst for Efficient Photocatalytic Hydrogen Evolution. J. Am. Chem. Soc. 2014, 136, 14015–14018. [Google Scholar] [CrossRef] [PubMed]
  23. Amthor, S.; Knoll, S.; Heiland, M.; Zedler, L.; Li, C.; Nauroozi, D.; Tobaschus, W.; Mengele, A.K.; Anjass, M.; Schubert, U.S.; et al. A Photosensitizer–Polyoxometalate Dyad That Enables the Decoupling of Light and Dark Reactions for Delayed on-Demand Solar Hydrogen Production. Nat. Chem. 2022, 14, 321–327. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, Y.R.; Huang, Q.; He, C.T.; Chen, Y.; Liu, J.; Shen, F.C.; Lan, Y.Q. Oriented Electron Transmission in Polyoxometalate-Metalloporphyrin Organic Framework for Highly Selective Electroreduction of CO2. Nat Commun. 2018, 9, 4466. [Google Scholar] [CrossRef] [Green Version]
  25. Jiao, L.; Dong, Y.; Xin, X.; Qin, L.; Lv, H. Facile Integration of Ni-Substituted Polyoxometalate Catalysts into Mesoporous Light-Responsive Metal-Organic Framework for Effective Photogeneration of Hydrogen. Appl. Catal. B 2021, 291, 120091. [Google Scholar] [CrossRef]
  26. Sun, W.; Zhu, J.; Zhang, M.; Meng, X.; Chen, M.; Feng, Y.; Chen, X.; Ding, Y. Recent Advances and Perspectives in Cobalt-Based Heterogeneous Catalysts for Photocatalytic Water Splitting, CO2 Reduction, and N2 Fixation. Chin. J. Catal. 2022, 43, 2273–2300. [Google Scholar] [CrossRef]
  27. Ginsberg, A.P. Inorganic Syntheses; John Wiley & Sons: New York, NY, USA, 1990; p. 108. [Google Scholar]
  28. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.a.K.; Puschmann, H. OLEX2: A Complete Structure Solution, Refinement and Analysis Program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  29. Sheldrick, G.M. SHELXT–Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr. Sect. A Found. Crystallogr. 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  30. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. Sect. C: Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  31. Spek, A.L. PLATON SQUEEZE: A Tool for the Calculation of the Disordered Solvent Contribution to the Calculated Structure Factors. Acta Crystallogr. Sect. C: Struct. Chem. C. 2015, 71, 9–18. [Google Scholar] [CrossRef]
  32. Xue, H.; Zhao, J.W.; Pan, R.; Yang, B.F.; Yang, G.Y.; Liu, H.S. Diverse Ligand-Functionalized Mixed-Valent Hexa-Manganese Sandwiched Silicotungstates with Single-Molecule Magnet Behavior. Chem. Eur. J. 2016, 22, 12322–12331. [Google Scholar] [CrossRef] [PubMed]
  33. Qin, L.; Zhao, C.; Yao, L.Y.; Dou, H.; Zhang, M.; Xie, J.; Weng, T.C.; Lv, H.; Yang, G.Y. Efficient Photogeneration of Hydrogen Boosted by Long-Lived Dye-Modified Ir(III) Photosensitizers and Polyoxometalate Catalyst. CCS Chem. 2022, 4, 259–271. [Google Scholar] [CrossRef]
  34. Ding, Y.S.; Wang, H.Y.; Ding, Y. Visible-Light-Driven Hydrogen Evolution Using a Polyoxometalate-Based Copper Molecular Catalyst. Dalton Trans. 2020, 49, 3457–3462. [Google Scholar] [CrossRef]
  35. Wang, Z.; Lu, Y.; Li, Y.; Wang, S.; Wang, E. Visible-Light Photocatalytic H2 Evolution over a Series of Transition Metal Substituted Keggin-Structure Heteropoly Blues. Chin. Sci. Bull. 2012, 57, 2265–2268. [Google Scholar] [CrossRef] [Green Version]
  36. Shang, X.; Liu, R.; Zhang, G.; Zhang, S.; Cao, H.; Gu, Z. Artificial Photosynthesis for Solar Hydrogen Generation over Transition-Metal Substituted Keggin-Type Titanium Tungstate. New J. Chem. 2014, 38, 1315–1320. [Google Scholar] [CrossRef]
  37. Sun, J.J.; Wang, W.D.; Li, X.Y.; Yang, B.F.; Yang, G.Y. {Cu8} Cluster-Sandwiched Polyoxotungstates and Their Polymers: Syntheses, Structures, and Properties. Inorg. Chem. 2021, 60, 10459–10467. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) View of polyoxoanion of 1; (b) the coordination environment of [Co3O23-OH)2(H2O)2- (B-α-PW9O34)]9−. Symmetry codes: A (x, y, 1.5 − z), B (1 − y, 1 + x − y, z); (c) polyhedral view of [Co3O23-OH)2 -(H2O)2(B-α-PW9O34)]9−; (d) the {Co9} cluster capped by two HPO4 groups.
Figure 1. (a) View of polyoxoanion of 1; (b) the coordination environment of [Co3O23-OH)2(H2O)2- (B-α-PW9O34)]9−. Symmetry codes: A (x, y, 1.5 − z), B (1 − y, 1 + x − y, z); (c) polyhedral view of [Co3O23-OH)2 -(H2O)2(B-α-PW9O34)]9−; (d) the {Co9} cluster capped by two HPO4 groups.
Molecules 28 00664 g001
Figure 2. (a) Polyhedral view of 1 along the c-axis. Color code: WO6: red, CoO6: pink, PO4: yellow, Cl@{Cs3(H2O)6}: green; (b) the topological network of 1 along the c-axis. (HPO4)2@{Co9(PW9)3} 3-c node: red; Cl@{Cs3(H2O)6} 3-c node: green; (c) the 2D layers are arranged in –ABAB– manner along the c-axis.
Figure 2. (a) Polyhedral view of 1 along the c-axis. Color code: WO6: red, CoO6: pink, PO4: yellow, Cl@{Cs3(H2O)6}: green; (b) the topological network of 1 along the c-axis. (HPO4)2@{Co9(PW9)3} 3-c node: red; Cl@{Cs3(H2O)6} 3-c node: green; (c) the 2D layers are arranged in –ABAB– manner along the c-axis.
Molecules 28 00664 g002
Figure 3. Photocatalytic H2 Evolution performance of 1: (a) Time-dependent H2 yield of different TEOA concentrations (5–20 mM) of 1. Reaction conditions: 1 (3 mg), white light (400–800 nm, 10 W), [Ir(ppy)2(dtbbpy)][PF6] (0.2 mM), 6 mL CH3CN/DMF (1:3) and H2O (2 M) degassed with Ar/CH4 (4:1); (b) Time-dependent H2 yield of different amounts of 1 (1~6 mg); (c) Time-dependent H2 yield of different concentrations of [Ir(ppy)2(dtbbpy)][PF6]; (d) Time-dependent H2 yield of light source and blank control under the condition of equimolar amount of 1, {PW9} and CoCl2.
Figure 3. Photocatalytic H2 Evolution performance of 1: (a) Time-dependent H2 yield of different TEOA concentrations (5–20 mM) of 1. Reaction conditions: 1 (3 mg), white light (400–800 nm, 10 W), [Ir(ppy)2(dtbbpy)][PF6] (0.2 mM), 6 mL CH3CN/DMF (1:3) and H2O (2 M) degassed with Ar/CH4 (4:1); (b) Time-dependent H2 yield of different amounts of 1 (1~6 mg); (c) Time-dependent H2 yield of different concentrations of [Ir(ppy)2(dtbbpy)][PF6]; (d) Time-dependent H2 yield of light source and blank control under the condition of equimolar amount of 1, {PW9} and CoCl2.
Molecules 28 00664 g003
Figure 4. (a) Temperature dependence of χm and χmT values for 1 at 5000 Oe; (b) temperature dependence of χm−1 values for 1 at 5000 Oe. The red line is fit to the Curie–Weiss law.
Figure 4. (a) Temperature dependence of χm and χmT values for 1 at 5000 Oe; (b) temperature dependence of χm−1 values for 1 at 5000 Oe. The red line is fit to the Curie–Weiss law.
Molecules 28 00664 g004
Table 1. Crystal data and structure refinements for 1.
Table 1. Crystal data and structure refinements for 1.
1
FormulaH103Na8ClCs3Co9O158P5W27
Mr8898.75
Crystal systemHexagonal
Space groupP63/m
a [Å]20.7840(8)
b [Å]20.7840(8)
c [Å]20.6385(9)
α [º]90
β [º]90
γ [º]120
V3]7721.1
Z, Dc [g cm−3]2
F(000)7316.0
GOF on F21.036
Final R indices [I > 2σ(I)]R1 = 0.0762 wR2 = 0.2026
R indices [all data] aR1 = 0.01328 wR2 = 0.2464
CCDC number2213769
a R1 = Σ||F0| − |Fc||/Σ|F0|. wR2 = [Σw(F20 − F2c)2/Σw(F20)2]1/2.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, Z.-W.; Yang, G.-Y. A {Co9}-Added Polyoxometalate for Efficient Visible-Light-Driven Hydrogen Evolution. Molecules 2023, 28, 664. https://doi.org/10.3390/molecules28020664

AMA Style

Wang Z-W, Yang G-Y. A {Co9}-Added Polyoxometalate for Efficient Visible-Light-Driven Hydrogen Evolution. Molecules. 2023; 28(2):664. https://doi.org/10.3390/molecules28020664

Chicago/Turabian Style

Wang, Zhen-Wen, and Guo-Yu Yang. 2023. "A {Co9}-Added Polyoxometalate for Efficient Visible-Light-Driven Hydrogen Evolution" Molecules 28, no. 2: 664. https://doi.org/10.3390/molecules28020664

Article Metrics

Back to TopTop