Next Article in Journal
Slater–Pauling Behavior in Half-Metallic Heusler Compounds
Previous Article in Journal
Direct Optical Patterning of Quantum Dots: One Strategy, Different Chemical Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Hexadecanuclear Cobalt-Added Tungstogermanate Containing Counter Cobalt Hydrates: Synthesis, Structure and Photocatalytic Properties

MOE Key Laboratory of Cluster Science, School of Chemistry and Chemical Engineering, Beijing Institute of Technology, Beijing 102488, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(13), 2009; https://doi.org/10.3390/nano13132009
Submission received: 11 June 2023 / Revised: 3 July 2023 / Accepted: 4 July 2023 / Published: 5 July 2023

Abstract

:
The synthesis and exploration of the properties of structurally-new polyoxometalates (POMs) has been attracting considerable research interest. In this work, a hexadecanuclear cobalt-added tungstogermanate, H31(NH4)5Na16{Co(H2O)6}4{[Co43-OH)3(PO4)]4(A-α-GeW9O34)4}2·23-H2O (1), was synthesized under hydrothermal conditions and characterized by various techniques. Compound 1 can effectively drive the heterogeneous photocatalytic hydrogen evolution reaction in the presence of [Ir(ppy)2(dtbbpy)][PF6] as the photosensitizer, with triethanolamine (TEOA) and N-Hydroxy succinimide (NHS) used as the dual sacrificial reagents. Control experiments revealed the important role of NHS in enhancing the hydrogen-evolution activities. Under optimal catalytic conditions, a hydrogen yield of 54.21 μmol was achieved after 10-h photocatalysis, corresponding to a hydrogen evolution rate of 1807.07 μmol·g−1·h−1. Stability studies demonstrated that catalyst 1 can be isolated and reused for three successive photocatalytic cycles with negligible decline of the H2 yield, indicating the stability and recycling robustness of catalyst 1.

1. Introduction

Polyoxometalates (POMs), a class of nano-sized metal-oxo clusters, are formed through the condensation of metalates in their high oxidation states (e.g., MoVI, WVI, VV, NbV, TaV) [1,2], which usually exhibit some advantages: tunable molecular structures, reversible multi-electron-storing abilities, and excellent redox properties [3,4]. The explicit structures of POMs can be easily determined by modern crystallographic characterization techniques, making it convenient to investigate their physicochemical properties at the atomic level [5,6]. To date, POMs have been widely used in the fields of catalysis [7,8,9], electrochemistry [10,11], magnetism [12,13], and pharmaceutical chemistry [14]. Typically, lacunary Keggin-type POMs (e.g., {XW9O34}, X = Si, Ge, P, etc.) can be generated by removing one or several {WO6}s from the plenary {XW12O40} (X = Si, Ge, P, etc.) structures. The lacunary tungsten-oxo clusters can act as multidentate inorganic ligands, incorporating transition metal ions to form transition metal-added POMs (TMAPs) [13,15]. By following the lacunary-directing, synthetic strategy, numerous TMAPs have been reported [15,16,17,18,19,20,21,22,23,24]. TMAPs have been comprehensively studied due to their integrated advantages, including added transition metal ions and lacunary POM ligands, which result in good structural stability and interesting catalytic properties [20,21].
Among various TMAPs, the synthesis and application of high-nuclearity, cobalt-added polyoxometalates (CoAPs) remain one of the most attractive research directions for global researchers. By controlling reaction conditions, lacunary tungsten-oxo clusters can be made to undergo diverse configuration changes. Different lacunary tungsten-oxo clusters possess contrasting coordination abilities, which can therefore induce cobalt ions to aggregate in various forms. Some representative high-nuclearity CoAPs include [{(B-α-PW9O34)Co3(OH)(H2O)2(O3PC(O)(C3H6NH3)PO3)}2Co]14− [25], [Co9(OH)3(H2O)6(PO4)2(B-α-PW9O34)3]21− [26], [Co6(H2O)30{Co9Cl2(OH)3(H2O)9(β-SiW8O31)3}]5− [27], [{Co4(OH)3PO4}4− (PW9O34)4]28− [12,20], and [Co21(H2O)4(OH)12(SiW10O37)6]30− [28]. In these reported CoAPs, the high-nuclearity aggregation of cobalt-oxo clusters can be achieved with the aid of some organic/inorganic linkers. With respect to the synthetic approaches, most of these reported CoAPs have been synthesized using the conventional aqueous solution-based method, which usually requires good solubility of starting materials in water. In contrast, the emerging hydro(solvo)thermal synthetic approach could be more applicable to synthetic systems even though the solubility of starting reagents is relatively poor. In addition, the high-temperature and -pressure environment associated with hydro(solvo)thermal reactions is beneficial for the formation of some structurally-new POM intermediates, which contrasts with the thermodynamically dominant products synthesized by the conventional solution-based method [1,2,3].
Cobalt-based complexes have exhibited interesting catalytic applications, especially, given the increasing concerns about the energy crisis and environmental problems, in solar-driven water splitting [29,30]. Due to the complexity of total water splitting reactions, they have often been studied in separate half-parts: oxygen evolution reactions (OERs) and hydrogen evolution reactions (HERs) [31]. To date, the CoAPs have been widely investigated as water oxidation catalysts in both photo-/electrochemical systems [9,20,25,32,33,34,35,36,37,38]. The well-defined structures of CoAPs are beneficial for comprehensive studies on the relationship between the catalytic mechanism and structure properties. In contrast, very few works have been published that introduce CoAPs as catalysts in visible-light-driven water-reduction reactions [39,40,41], although many TMAPs were reported as the multi-electron-transfer catalysts for HERs [7,42,43,44,45]. In this regard, the continuous design and photocatalytic HER exploration of novel, high-nuclearity CoAPs remains a critical and interesting research direction.
Herein, we have successfully synthesized and systematically characterized water-insoluble and stable CoAPs, H31(NH4)5Na16{Co(H2O)6}4{[Co43-OH)3(PO4)]4(A-α-GeW9O34)4}2·23H2O (1), under hydrothermal conditions by introducing {GeW9O34} as the multidentate ligand and phosphate as a bridging linker to the central cobalt-oxo clusters. Compound 1 was investigated as a stable and recyclable heterogeneous catalyst to drive the photocatalytic HER. In the photocatalytic system, dual sacrificial reagents were first applied for improved performance.

2. Experiments Section

2.1. General Procedure

The chemicals utilized in this work are commercially obtained and used without further purification. The tri-vacant POMs K8Na2[A-α-GeW9O34]·25H2O is prepared according to the previous report [46,47]. The basic characterization instruments and methods are as follows: (1) A Bruker D8 Advance X-ray diffractometer (Bruker, Karlsruhe, Germany) equipped with Cu Kα radiation (λ = 1.54056 Å) scanning from 5 to 50° was used to determine the powder X-ray diffraction (PXRD) patterns of 1. (2) The Nicolet iS10 FT-IR spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) was used for IR spectrum (400–4000 cm−1) measurements with KBr pellets. (3) The optical gap of 1 was measured and estimated from UV-Vis diffuse reflectance spectra (200–800 nm) using a Shimadzu UV3600 spectrometer (Shimadzu, Kyoto, Japan). (4) Thermal stability was confirmed by thermogravimetric analyses using a Mettler Toledo TGA/DSC 1100 analyzer (Mettler Toledo, Zurich, Switzerland). The testing temperature ranges from 25 to 1000 °C in air at a heating rate of 10 °C·h–1. (5) Inductively coupled plasma mass spectrometry was measured on an ICP-OES (iCAP 7400, Thermo, Waltham, MA, USA) to complete the elemental analysis and post-catalytic characterization.

2.2. Synthesis

A mixture of K8Na2[A-α-GeW9O34]·25H2O (0.30 g, 0.097 mmol), Co(NO3)2·6H2O (0.18 g, 0.62 mmol), NH4Cl (0.20 g, 3.73 mmol), and Na2HPO4·12H2O (0.10 g, 0.28 mmol) was introduced into 5 mL of deionized water. After stirring for 30 min, the pH of the solution was regulated to 9.50 by adding 4M of NaOH solution. The resulting purple-red turbid mixture was transferred into a 25 mL, Teflon-lined, stainless-steel autoclave and kept at 100 °C for 3 days. Pink spindle crystals were obtained after cooling the solution to room temperature. The crystals were further purified by washing and sonicating with deionized water. (Yield: 21%, based on K8Na2[A-α-GeW9O34]·25H2O). IR (KBr, cm−1): 3413 (vs), 1633 (vs), 1400 (vs), 1079 (w), 933 (vs), 869 (s), 806 (vs), 688 (vs). Elemental analysis (calcd) for compound 1 (wt %): Co 9.64 (9.30), Ge 3.01 (2.51), Na 1.90 (1.62), P 1.21 (1.09), W 58.21 (57.82).

2.3. X-ray Crystallography

A translucent pink crystal of 1 was sealed into a glass tube with Vaseline to collect the diffraction data by a Gemini A Ultra CCD diffractometer equipped with graphite monochromated Mo Kα (λ = 0.71073 Å) and radiation of 296(2) K. The structure was solved using the direct method and refined using full-matrix least-squares, which were fit using the F2 method in Olex2 [48] and the SHELX-2008 program package [49,50]. The contribution of the disordered water molecules to the diffraction data of 1 was examined using the SQUEEZE method in PLATON [51]. The crystalline water molecules cannot be fully identified by refinement (Table 1). The number of water molecules was quantified based on the TGA data shown in Figure S1, where 23 lattice water molecules exist in the structure. To obey the charge balance motif, 31 protons were added. The elemental compositions of compound 1 were further characterized by ICP-OES tests, which correlate with the single-crystal measurement results. The CCDC number of 1 is deposited as 2262758. Further data can be obtained from http://www.ccdc.cam.ac.uk/, accessed on 10 June 2023.

2.4. Photocatalytic Water Reduction Tests

The whole photocatalytic reaction system was comprised of acetonitrile/N,N-dimethylformamide (v/v = 1:3) as the mixed solvent, [Ir(ppy)2(dtbbpy)][PF6] as the photosensitizer, TEOA and NHS as the dual electron sacrificial reagents, 1 as the catalyst, and distilled water as the proton donor. The above components were transferred and sealed into the photoreactor. The reactor was deaerated by CH4/Ar (v/v = 4:1, CH4 as the internal standard substance) in a mixed atmosphere for ten minutes, and then transferred to a PCX-50C multi-channel photocatalytic reactor equipped with a white LED light plate (400–800 nm, electric power: 10 W) for photocatalytic reaction. During the photocatalysis, the reaction solution was continuously stirred at 400 rpm. The reaction temperature was kept at 25 °C by a water-circulating system. At an interval of two hours, 125 μL mix gas was extracted and qualitatively and quantitatively analyzed by a GC-979011 gas chromatograph analyzer.

3. Results and Discussion

3.1. Structure of Compound 1

1 crystallizes in the triclinic crystal system, P-1 space group. The structure consists of two hexadecanuclear cobalt-added tetrameric tungsten-oxo clusters {[Co43-OH)3(PO4)]4− (A-α-GeW9O34)4} [20,52], four free {Co(H2O)6} (Figure 1a,b and Figure S2), 16 Na+, 5 NH4+, 31 H+ as counter cations, and 23 crystalline water molecules. In the polyoxoanion of 1, four {A-α-GeW9 O34} POM ligands are located at the four vertices of a tetrahedron and further anchored to the {Co163-OH)12O36(PO4)4} cluster through μ2-O, μ3-O, and μ4-O atoms (Figure 1c,d). In another image, six vacant sites of each {A-α-GeW9O34} induce a defective-cubane-type tetranuclear cobalt-oxo cluster {Co4O3}, which gives rise to the {Co4GeW9}. Four identical {Co4GeW9} moieties are connected by four {PO4} linkers and four μ3-O atoms (Figure 1e).
All cobalt ions in the polyoxoanion of 1 exist in a hexacoordinated octahedron mode and are chemically present in bivalent forms as determined by the Bond Valence Sum (BVS) calculations [53]. The bond lengths of Co-O vary from 1.98 to 2.27 Å. For the four dissociated {Co(H2O)6}, the Co-O lengths are in the range of 1.85–2.03 Å. The BVS values of the four cobalt ions are 2.92, 2.89, 2.79, and 2.90, illustrating the existence of Co in the free cobalt hydrates. According to the Bond Valence Sum (BVS) calculation results [53], the chemical valance of W, Ge, and P is +6, +4, and +5, respectively. The oxygen in polyoxoanion 1 has three existing forms including O2−, OH, and H2O, as revealed by the structural analyses and BVS calculations.
The hexadecanuclear cobalt-oxo cluster {Co163-OH)12O36(PO4)4} can be structurally illustrated as three layers (Figure 2a). The first layer is {Co33-OH)6O9(PO4)} (Figure 2b,c), where a {PO4} linker is located in the center, with three hexacoordinated {Co(μ3-OH)2O4} links to {PO4} via three μ2-O atoms. The residual μ4-O of {PO4} is connected to the second layer that is a tri-phosphate-modified clover-like nonanuclear cobalt-oxo cluster {Co93-OH)9O19(PO4)3} (Figure 2d,e). Three hexacoordinated Co2+ are linked together by edge-sharing modes, creating one of the three “leaves” of the “clover”. The three “leaves” are held together by one μ4-O atom. Three {PO4} linkers are located in the same plane and fixed to the gap between the three “leaves” of the {Co93-OH)9O19(PO4)3}, connected by P-μ2-O-Co, P-μ3-O-Co, and P-μ4-O-Co linkages. Each of the three {PO4}s in the second layer can share a μ2-O to connect with the pedestal position Co of the third layer. The third layer is {Co43-OH)3O14} (Figure 2f,g), which is the traditional defective cubane-like configuration tetranuclear cobalt-oxo cluster. The vertex position Co of the third layer was linked to the center of the second layer through an edge-sharing mode.
In the polyoxoanion 1, {[Co43-OH)3(PO4)]4(A-α-GeW9O34)4} can also be simplified by omitting sectional oxygen ions from {A-α-GeW9O34} (Figure 3a). Four Co in each {Co43-OH)3O14} can form a {Co4} tetrahedron. The four Cos at the vertex position of four identical {Co4} tetrahedrons and four μ4-O atoms of {PO4} linkers can construct a {Co4O4} cubane in the center of {[Co43-OH)3(PO4)]4(A-α-GeW9O34)4} (Figure 3b,c).
From another view, {[Co43-OH)3(PO4)]4(A-α-GeW9O34)4} can also be presented as a tri-nested structure by simplifying {A-α-GeW9O34} as orange balls, cobalt ions as pink balls, and {PO4}s as yellow balls (Figure 4a). The four {A-α-GeW9O34} are located in the vertices of a tetrahedron, and the distance between each {A-α-GeW9O34} varies from 11.020 to 11.218 Å (Figure 4b). The arrangement of the four {PO4} linkers also shows a tetrahedron configuration. The distances between each {PO4} are in the range of 5.341–5.392 Å (Figure 4c). In the simplified {Co163-OH)12O36(PO4)4}, vertices of the four external {Co4} tetrahedrons can form an internal {Co4} tetrahedron. The distances between each Co are in the range of 3.017–3.811 Å (Figure 4d). The packing mode of 1 in the bc plane is shown in Figure S3.

3.2. FT-IR Measurement

As illustrated in Figure S4, the FT-IR spectrum of 1 involves four major characteristic vibrational groups. First, the wide peaks in 3423–3203 cm−1 and 1633 cm−1 can be attributed to the existence of lattice and coordination water in 1. The distinct vibrational peak observed at ~1400 cm−1 could be attributed to the existence of NH4+ counter cations. In the fingerprint region, the absorption peak centered at 1079 cm–1 is ascribed to the ν(P–O) stretching vibration mode. The typical vibration peaks of the tri-vacant {A-α-GeW9O34} precursor (e.g., ν(Ge-O) ν(W-Ot), ν(W-Ob), and ν(W-Oc)) are observed at 933 (vs), 869 (vs), 806 (vs), and 688 cm−1, respectively.

3.3. Powder X-ray Diffraction Patterns

To confirm the phase purity of 1, it is noted that there is agreement between the experimental PXRD pattern and the simulated pattern from the single-crystal X-ray diffraction data (Figure S5). The modest shift in peak intensity may be attributed to a change in the favored orientation of the powder sample during the PXRD data-collecting process.

3.4. UV-Vis Diffuse Reflectance Test

Compound 1 exhibited strong absorption in the range of 200–800 nm. The wide absorption peak at 200–400 nm corresponds to the O-to-W charge transfer, while the absorption of 500–600 nm is attributed to the d-d transitions of Co centers. The obtained absorbance-wavelength data is further processed according to the Kubelka-Munk formula α/S = (1 − R)2/(2R) (R = reflection coefficient, α = absorption factor, S = scattering factor) to make the α/S-Energy plot, and the tangent plot is made in the near-linear range of the obtained curve; the intersection point with α/S = 0 is the optical band gap value. As shown in Figure S6, the optical gap of compound 1 is estimated as 2.65 eV.

3.5. Photocatalytic Hydrogen Generation

Given the poor solubility of 1 in water, N,N-dimethylformamide, or acetonitrile, 1 was employed as a heterogeneous catalyst in the visible-light-driven hydrogen generation reaction. Various control experiments confirmed that the existence of the catalyst, photosensitizer, and sacrificial agent was indispensable for efficient photocatalysis (Figure 5a, Table S1). It is noted that the very small amount of hydrogen (0.42 μmol) was produced from the control experiment containing Ir photosensitizer and TEOA sacrificial donor, which is commonly observed in the literature [42,43,44,54]. Commonly, one type of electron donor is used in photocatalytic HER systems. Herein, when TEOA was used as the only sacrificial agent, only 6.81 μmol of H2 evolved. In 2019, Mialane’s group reported a tetradecanuclear nickel-added tungsten-oxo cluster [(A-α-SiW9O34)Ni14(AleH)5(Ale)2(H2O)11-(OH)7]12− (Ale = alendronate), which can act as a heterogeneous catalyst to drive the photocatalytic hydrogen evolution reaction in the presence of TEOA, 1-benzyl-1,4-dihydronicotinamide (BNAH), and [Ir(ppy)2(dtbbpy)][PF6]. Therein, BNAH served as the sacrificial electron donor in the system, and TEOA serves as a proton donor in the anhydrous system. The catalytic system obtained 18 mmol H2 after 7-h illumination [55]. Clearly, an additive sacrificial electron donor may improve the catalytic performance. Therefore, NHS was selected as an additive electron donor for the photocatalytic reaction in the present study. With the addition of NHS, other components showed similar control effects as that of the NHS-free system. To verify the potential proton donor in this dual sacrificial reagent system, an anhydrous system under otherwise identical conditions was tested and showed no H2 evolution (Figure 5a), indicating the important role of H2O in hydrogen generation. The optimal reaction condition was considered as 3 mg 1, 0.060 M TEOA, 0.153 mM NHS, 2 M H2O, and 0.20 mM [Ir(ppy)2(dtbbpy)][PF6]. After 10-h photocatalysis, the H2 yield of 54.21 μmol was obtained, corresponding to the highest hydrogen evolution rate of 1807.07 μmol·g−1·h−1. A nine-fold increase in hydrogen yield was achieved by introducing NHS into the catalytic system compared to that of NHS-free catalytic system (Figure 5a). To better illustrate the strong catalytic performance of compound 1, we have summarized the photocatalytic results of reported systems using cobalt-added tungsten-oxo clusters. Compared to the reported studies (Table S2), compound 1 exhibited one of the best photocatalytic performances in the absence of a noble metal co-catalyst (e.g., Pt). Also, to our knowledge, this is the first time that cobalt-added tungstogermanate has been investigated for visible-light-driven hydrogen production in the dual sacrificial reagent system. Considering that 1 contains Co2+, {GeW9O34} and {PO4}, additional control experiments using molar equivalents of Co(NO3)2·6H2O, K8Na2[A-α-GeW9O34], and Na2HPO4 as 3 mg of 1 have been further conducted to better understand the photocatalytic performance. As shown in Figure 5b, the systems catalyzed by Co(NO3)2·6H2O, K8Na2[A-α-GeW9O34], and Na2HPO4 resulted in the hydrogen yield of 6.32, 2.95, and 0.37 μmol, respectively. The mechanical mixture of all three components in equal molar proportions can catalyze the hydrogen production of 10.79 μmol (Figure 5b). All of these control experiment results demonstrated the importance of the intact molecular structure of 1 for effective photocatalysis.
In addition, it is worth mentioning that the color of the reaction system rapidly changed from bright yellow to dark blue upon light irradiation, simultaneously coupling the linear hydrogen evolution with the irradiation time. In the meanwhile, catalyst 1 was also transformed into the heteropoly blue intermediates, indicating the reduction of 1 during photocatalysis, which is commonly observed in the literature reports [42,54]. When the reaction system was moved to a dark environment, hydrogen production stopped immediately. The same phenomenon appeared in the light ON-OFF experiments (Figure 5c). After exposure to air, the heteropoly blue intermediates will return to their original pink color. The rapid formation and disappearance of heteropoly blue intermediates show the strong and reversible multielectron storage abilities of POMs. As discussed in previous reports, heteropoly blue intermediates act as an active species by receiving multiple electrons in the vacant W 5d orbitals, which promote the hydrogen evolution process [42,54]. The catalytic activity is also related to the amount of catalyst used. When the catalyst dosage was 1, 3, and 6 mg, the hydrogen yield reached 24.07, 54.21, and 51.44 μmol in 10 h, corresponding to the hydrogen production rate of 2407.62, 1807.07, and 857.46 μmol·g−1·h−1, respectively (Figure 5d and Figure S7). In addition, the concentration of photosensitizers is also a significant factor to the enhancement of hydrogen production. When the concentration of [Ir(ppy)2(dtbbpy)][PF6] was changed from 0.10 to 0.25 mM, the hydrogen evolution exhibited linear growth with reaction time, yielding 11.04 to 71.84 μmol H2 after 10-h. The hydrogen evolution rates ranged from 924.47 to 2394.91 μmol·g−1·h−1 (Figure 5e). Moreover, because the experiments varied the concentration of TEOA from 0.03 to 0.06 M, the hydrogen production ranged from 35.72 to 54.21 μmol. The calculated hydrogen evolution rate ranged from 1190.69 to 1807.07 μmol·g−1·h−1 (Figure 5f). It is noted that varying the concentration of NHS from 0.075 mM to 0.300 mM barely changed the H2 yield, which maintained a constant concentration of TEOA (0.060 M) (Figure S8).
In order to further assess the stability of 1 in the photocatalytic process, a long-term reaction lasting 24 h was carried out, and an amount of 85.40 μmol of hydrogen was cumulatively produced (Figure 6a). The hydrogen production rate gradually decreased over time due to the partial consumption of the sacrificial reagent and photosensitizer in the catalytic system. The catalysts after photocatalysis were isolated by centrifugation and washed with CH3CN. The isolated catalyst was further used for three successive photocatalytic cycles in the presence of fresh sacrificial reagents and the photosensitizer; the photocatalytic hydrogen production performance of 1 was maintained with negligible decline of the H2 yield (Figure 6b), proving the good stability and recycling robustness of catalyst 1. The structural stability of catalyst 1 was further confirmed by the perfectly matched FT-IR spectra of the isolated catalyst with that of the fresh compound 1 (Figure 6c). In order to prove the heterogeneity of the reaction system, an experiment was carried out to analyze the leaching of Co in the post-catalysis solution using the ICP-OES test. The results showed there is little Co (0.007 μmol) dissolution after the photocatalytic reaction (Table S3), which cannot significantly contribute to the hydrogen production activity, as demonstrated by the control experiment (Figure 5b, red line). Overall, all above experimental results and analyses demonstrated that compound 1 can work as a promising heterogeneous catalyst for visible-light-driven hydrogen production with decent activity, stability, and recyclability.

4. Conclusions

Under hydrothermal conditions, a water-insoluble and stable CoAPs, H31(NH4)5-Na16{Co(H2O)6}4{[Co43-OH)3(PO4)]4(A-α-GeW9O34)4}2·23H2O (1), was synthesized and characterized by various techniques. Distinguished from the reported hexadecanuclear cobalt-added POM, four additional free {Co(H2O)6} crystallized in 1. The introduction of phosphate linkers contributes to the aggregation of cobalt-oxo clusters. The hydrothermal synthetic method is beneficial for the synthesis of stable and structurally-new compound 1. While coupling with the photosensitizer and sacrificial reagent, 1 worked as a heterogeneous catalyst for visible-light-driven hydrogen evolution. Under optimal catalytic conditions, a hydrogen yield of 54.21 μmol was achieved after 10-h photocatalysis, corresponding to a hydrogen evolution rate of 1807.07 μmol·g−1·h−1. Compared to the classic three-component HER system, NHS was firstly applied to construct a dual sacrificial reagent system, which can significantly enhance the hydrogen evolution activity. The photocatalytic results confirm the enhanced effects of such dual-electron-sacrificial-reagents systems. This study provides new guidance for designing highly stable cobalt-added POMs and expanding their catalytic applications in solar conversion.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13132009/s1, Figure S1. TGA curve of 1. Figure S2. Ball and stick model diagram of 1. Color codes: W atom, black; Ge atom, orange; Co atom, turquoise; P atom, yellow; O atom, red. Figure S3. The packing mode of 1 in the bc plane. Color codes: WO6, red; CoO6, turquoise; PO4, yellow; GeO4, orange. Figure S4. FT-IR spectrum of 1. Figure S5. Experimental and simulated PXRD patterns of 1. Figure S6. UV-Vis diffuse reflectance spectrum of 1 and the corresponding α/S-Energy curve (inset). Figure S7. Relationship between hydrogen yield and hydrogen evolution rate in different qualities of 1. Figure S8. Hydrogen production comparison of different concentrations of NHS. Table S1. Control experiments for photocatalytic water reduction reaction in NHS-free conditions. Table S2. Overview of different photocatalytic HER performance using cobalt-added tungsten-oxo clusters [39,40,41,56,57]. Table S3. ICP-OES analyses of the cobalt contents in compound 1 and the post-catalytic solution after 10-h photocatalysis.

Author Contributions

Conceptualization, Q.Z.; methodology, Q.Z.; formal analysis, Q.Z.; investigation, Q.Z.; data curation, Q.Z. and X.L.; writing—original draft preparation, Q.Z.; writing—review and editing, Q.Z., X.L., Y.W., H.L. and G.Y.; funding acquisition, G.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, 21871025, 21571016, 91122028, and 20725101.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zheng, S.-T.; Yang, G.-Y. Recent advances in paramagnetic-TM-substituted polyoxometalates (TM = Mn, Fe, Co, Ni, Cu). Chem. Soc. Rev. 2012, 41, 7623–7646. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, S.-S.; Yang, G.-Y. Recent Advances in Polyoxometalate-Catalyzed Reactions. Chem. Rev. 2015, 115, 4893–4962. [Google Scholar] [CrossRef] [PubMed]
  3. Zhao, J.-W.; Li, Y.-Z.; Chen, L.-J.; Yang, G.-Y. Research progress on polyoxometalate-based transition-metal-rare-earth hetero-metallic derived materials: Synthetic strategies, structural overview and functional applications. Chem. Commun. 2016, 52, 4418–4445. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, Z.-M.; Xin, X.; Zhang, M.; Li, Z.; Lv, H.-J.; Yang, G.-Y. Recent advances of mixed-transition-metal-substituted polyoxometalates. Sci. China Chem. 2022, 65, 1515–1525. [Google Scholar] [CrossRef]
  5. Kim, N.; Nam, J.-S.; Jo, J.; Seong, J.; Kim, H.; Kwon, Y.; Lah, M.-S.; Lee, J.-H.; Kwon, T.-H.; Ryu, J. Selective photocatalytic production of CH4 using Zn-based polyoxometalate as a nonconventional CO2 reduction catalyst. Nanoscale Horiz. 2021, 6, 379–385. [Google Scholar] [CrossRef]
  6. Zhao, W.; Zeng, X.; Huang, L.; Qiu, S.; Xie, J.; Yu, H.; Wei, Y. Oxidative dehydrogenation of hydrazines and diarylamines using a polyoxomolybdate-based iron catalyst. Chem. Commun. 2021, 57, 7677–7680. [Google Scholar] [CrossRef]
  7. Li, N.; Liu, J.; Dong, B.-X.; Lan, Y.-Q. Polyoxometalate-based Compounds for Photo- and Electrocatalytic Applications. Angew. Chem. Int. Ed. 2020, 59, 2–17. [Google Scholar] [CrossRef]
  8. Neumann, R.; Azaiza-Dabbah, D.; Vogt, C.; Wang, F.; Masip-Sanchez, A.; de Graaf, C.; Poblet, J.M.; Haviv, E. Molecular Transition Metal Oxide Electrocatalysts for the Reversible Carbon Dioxide-Carbon Monoxide Transformation. Angew. Chem. Int. Ed. 2022, 61, e202112915. [Google Scholar] [CrossRef]
  9. Blasco-Ahicart, M.; Soriano-López, J.; Carbó, J.-J.; Poblet, J.-M.; Galan-Mascaros, J.-R. Polyoxometalate Electrocatalysts Based on Earth-abundant Metals for Efficient Water Oxidation in Acidic Media. Nat. Chem. 2017, 10, 24–30. [Google Scholar] [CrossRef]
  10. Ni, L.-B.; Yang, G.; Liu, Y.; Wu, Z.; Ma, Z.; Shen, C.; Lv, Z.; Wang, Q.; Gong, X.; Xie, J.; et al. Self-assembled Polyoxometalate Nanodots as Bidirectional Cluster Catalysts for Polysulfide/Sulfide Redox Conversion in Lithium–sulfur Batteries. ACS Nano 2021, 15, 12222–12236. [Google Scholar] [CrossRef]
  11. Li, X.-X.; Ji, T.; Gao, J.-Y.; Chen, W.-C.; Yuan, Y.; Sha, H.-Y.; Faller, R.; Shan, G.-G.; Shao, K.-Z.; Wang, X.-L.; et al. An Unprecedented Fully Reduced {Mo60V} Polyoxometalate: From All-inorganic Molecular Light-absorber Model to Improved Photoelectronic Performance. Chem. Sci. 2022, 13, 4573–4580. [Google Scholar] [CrossRef] [PubMed]
  12. Ibrahim, M.; Lan, Y.; Bassil, B.-S.; Xiang, Y.; Suchopar, A.; Powell, A.-K.; Kortz, U. Hexadecacobalt(II)-containing Polyoxo-metalate-Based Single-molecule Magnet. Angew. Chem. Int. Ed. 2011, 50, 4708–4711. [Google Scholar] [CrossRef]
  13. 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]
  14. Rhule, J.-T.; Hill, C.-L.; Judd, D.-A.; Schinazi, R.-F. Polyoxometalates in medicine. Chem. Rev. 1998, 98, 327–357. [Google Scholar] [CrossRef]
  15. Zhao, J.-W.; Jia, H.-P.; Zhang, J.; Zheng, S.-T.; Yang, G.-Y. A Combination of Lacunary Polyoxometalates and High-nuclear Transition-metal Clusters under Hydrothermal Conditions. Part II: From Double Cluster, Dimer, and Tetramer to Three-dimensional Frameworks. Chem. Eur. J. 2007, 13, 10030–10045. [Google Scholar] [CrossRef]
  16. Zheng, S.-T.; Zhang, J.; Yang, G.-Y. Designed Synthesis of POM-organic Frameworks from {Ni6PW9} Building Blocks under Hydrothermal Conditions. Angew. Chem. Int. Ed. 2008, 47, 3909–3913. [Google Scholar] [CrossRef]
  17. Zheng, S.-T.; Zhang, J.; Clemente-Juan, J.-M.; Da-Qiang, Y.; Yang, G.-Y. Poly(polyoxotungstate)s with 20 Nickel Centers: From Nanoclusters to One-dimensional Chains. Angew. Chem. Int. Ed. 2009, 121, 7312–7315. [Google Scholar] [CrossRef]
  18. Zheng, S.-T.; Zhang, J.; Li, X.-X.; Fang, W.-H.; Yang, G.-Y. Cubic Polyoxometalate-organic Molecular Cage. J. Am. Chem. Soc. 2010, 132, 15102–15103. [Google Scholar] [CrossRef]
  19. 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]
  20. Han, X.-B.; Zhang, Z.-M.; Zhang, T.; Li, Y.-G.; Lin, W.-B.; You, W.-S.; Su, Z.-M.; Wang, E.-B. Polyoxometalate-based Cobalt-phosphate Molecular Catalysts for Visible Light-driven Water Oxidation. J. Am. Chem. Soc. 2014, 136, 5359–5366. [Google Scholar] [CrossRef]
  21. Han, X.-B.; Li, Y.-G.; Zhang, Z.-M.; Tan, H.-Q.; Lu, Y.; Wang, E.-B. Polyoxometalate-Based Nickel Clusters as Visible Light-driven Water Oxidation Catalysts. J. Am. Chem. Soc. 2015, 137, 5486–5493. [Google Scholar] [CrossRef] [PubMed]
  22. Goura, J.; Bassil, B.; Ma, X.; Rajan, A.; Moreno-Pineda, E.; Schnack, J.; Ibrahim, M.; Powell, A.; Ruben, M.; Wang, J.-J.; et al. Ni36II-containing 54-tungsto-6-silicate: Synthesis, Structure, Magnetic and Electrochemical Studies. Chem. Eur. J. 2021, 27, 15081–15085. [Google Scholar] [CrossRef] [PubMed]
  23. Chang, Q.; Meng, X.-Y.; Ruan, W.-J.; Feng, Y.-Q.; Li, R.; Zhu, J.-Y.; Ding, Y.; Lv, H.-J.; Wang, W.; Chen, G.-Y.; et al. Metal-organic Cages with {SiW9Ni4} Polyoxotungstate Nodes. Angew. Chem. Int. Ed. 2022, 134, e202117637. [Google Scholar] [CrossRef]
  24. 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 Atoms. CCS Chem. 2022, 4, 1305–1314. [Google Scholar] [CrossRef]
  25. Hu, Q.-Y.; Chen, S.-S.; Wågberg, T.; Zhou, H.-S.; Li, S.-J.; Li, Y.-D.; Tan, Y.-L.; Hu, W.-Q.; Ding, Y.; Han, X.-B. Developing Insoluble Polyoxometalate Clusters to Bridge Homogeneous and Heterogeneous Water Oxidation Photocatalysis. Angew. Chem. Int. Ed. 2023, 2023, e202303290. [Google Scholar] [CrossRef]
  26. Galán-Mascarós, J.-R.; Gómez-Garcia, C.-J.; Borrás-Almenar, J.-J.; Coronado, E. High Nuclearity Magnetic Clusters: Magnetic Properties of a Nine Cobalt Cluster Encapsulated in a Polyoxometalate, [Co9(OH)3(H2O)6(HPO4)2(PW9O34)3]16−. Adv. Mater. 1994, 6, 221–223. [Google Scholar] [CrossRef]
  27. Bassil, B.-S.; Nellutla, S.; Kortz, U.; Stowe, A.-C.; Van Tol, J.; Dalal, N.-S.; Keita, B.; Nadjo, L. The Satellite-shaped Co-15 Polyoxotungstate, [Co6(H2O)30{Co9Cl2(OH)3(H2O)9(β-SiW8O31)3}]5−. Inorg. Chem. 2005, 44, 2659–2665. [Google Scholar] [CrossRef]
  28. Shi, D.-Y.; Cui, C.-J.; Sun, C.-X.; Du, J.-P.; Liu, C.-S. A New [Co21(H2O)4(OH)12]30+ Unit-incorporating Polyoxotungstate for Sensitive Detection of dichlorvos. New J. Chem. 2020, 44, 11336–11341. [Google Scholar] [CrossRef]
  29. Dempsey, J.-L.; Brunschwig, B.-S.; Winkler, J.-R.; Gray, H.-B. Hydrogen Evolution Catalyzed by Cobaloximes. Acc. Chem. Res. 2009, 42, 1995–2004. [Google Scholar] [CrossRef] [Green Version]
  30. Artero, V.; Chavarot-Kerlidou, M.; Fontecave, M. Splitting Water with Cobalt. Angew. Chem. Int. Ed. 2011, 50, 7238–7266. [Google Scholar] [CrossRef]
  31. Jiao, Y.; Zheng, Y.; Jaroniecb, M.; Qiao, S.-Z. Design of electrocatalysts for oxygen- and hydrogen-involving energy conversion reactions. Chem. Soc. Rev. 2015, 44, 2060–2086. [Google Scholar] [CrossRef] [PubMed]
  32. Huang, Z.-Q.; Luo, Z.; Geletii, Y.-V.; Vickers, J.-W.; Yin, Q.; Wu, D.; Hou, Y.; Ding, Y.; Song, J.; Musaev, D.-G.; et al. Efficient light-driven carbon-free cobalt-based molecular catalyst for water oxidation. J. Am. Chem. Soc. 2011, 133, 2068–2071. [Google Scholar] [CrossRef]
  33. Lv, H.-J.; Geletii, Y.-V.; Zhao, C.-C.; Vickers, J.-W.; Zhu, G.-B.; Luo, Z.; Song, J.; Lian, T.-Q.; Musaev, D.-G.; Hill, C.-L. Polyoxo-metalate water oxidation catalysts and the production of green fuel. Chem. Soc. Rev. 2012, 41, 7572–7589. [Google Scholar] [CrossRef]
  34. Lv, H.-J.; Song, J.; Geletii, Y.-V.; Vickers, J.-W.; Sumliner, J.-M.; Musaev, D.-G.; Kögerler, P.; Zhuk, P.-F.; Bacsa, J.; Zhu, G.-B.; et al. An Exceptionally Fast Homogeneous Carbon-free Cobalt-based Water Oxidation Catalyst. J. Am. Chem. Soc. 2014, 136, 9268–9271. [Google Scholar] [CrossRef]
  35. Stracke, J.-J.; Finke, R.-G. Water Oxidation Catalysis Beginning with [Co4(H2O)2(PW9O34)2]10− When Driven by the Chemical Oxidant Ruthenium(III)tris(2,2′-bipyridine): Stoichiometry, Kinetic, and Mechanistic Studies en Route to Identifying the True Catalyst. ACS Catal. 2014, 4, 79–89. [Google Scholar] [CrossRef]
  36. Haider, A.; Bassil, B.-S.; Soriano-López, J.; Qasim, H.-M.; De Pipaón, C.S.; Ibrahim, M.; Dutta, D.; Koo, Y.-S.; Carbó, J.-J.; Poblet, J.-M.; et al. 9-Cobalt(II)-Containing 27-Tungsto-3-germanate(IV): Synthesis, Structure, Computational Modeling, and Heterogeneous Water Oxidation Catalysis. Inorg. Chem. 2019, 58, 11308–11316. [Google Scholar] [CrossRef]
  37. Dong, K.-L.; Ma, P.-T.; Wu, H.-C.; Wu, Y.-K.; Niu, J.-Y.; Wang, J.-P. Cobalt- and Nickel-Containing Germanotungstates Based on Open Wells–Dawson Structure: Synthesis and Characterization of Tetrameric Anion. Inorg. Chem. 2019, 58, 6000–6007. [Google Scholar] [CrossRef]
  38. Al-Sayed, E.; Nandan, S.-P.; Tanuhadi, E.; Giester, G.; Arrigoni, M.; Madsen, G.-K.-H.; Cherevan, A.; Eder, D.; Rompel, A. Phosphate-Templated Encapsulation of a {CoII4O4} Cubane in Germanotungstates as Carbon-Free Homogeneous Water Oxidation Photocatalysts. ChemSusChem 2021, 14, 2529–2536. [Google Scholar] [CrossRef]
  39. Zhao, J.-L.; Ding, Y.; Wei, J.; Du, X.-Q.; Yu, Y.-Z.; Han, R.-X. A Molecular Keggin Polyoxometalate Catalyst with High Efficiency for Visible-light Driven Hydrogen Evolution. Int. J. Hydrog. Energy 2014, 39, 18908–18918. [Google Scholar] [CrossRef]
  40. Wu, W.-M.; Teng, T.; Wu, X.-Y.; Dui, X.-J.; Zhang, L.; Xiong, J.-H.; Wu, L.; Lu, C.-Z. A Cobalt-based Polyoxometalate Catalyst for Efficient Visible-light-driven H2 Evolution from Water Splitting. Catal. Commun. 2015, 64, 44–47. [Google Scholar] [CrossRef]
  41. Wang, Z.-W.; Yang, G.-Y. A {Co9}-Added Polyoxometalate for Efficient Visible-light-driven Hydrogen Evolution. Molecules 2023, 28, 664. [Google Scholar] [CrossRef]
  42. Lv, H.-J.; Guo, W.-W.; Wu, K.-F.; Chen, Z.-Y.; Bacsa, J.; Musaev, D.-G.; Geletii, Y.-V.; Lauinger, S.-M.; Lian, T.-Q.; 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]
  43. Lv, H.-J.; Gao, Y.-Z.; Guo, W.-W.; Lauinger, S.M.; Chi, Y.-N.; Bacsa, J.; Sullivan, K.P.; Wieliczko, M.; Musaev, D.G.; Hill, C.L. Cu-based Polyoxometalate Catalyst for Efficient Catalytic Hydrogen Evolution. Inorg. Chem. 2016, 55, 6750–6758. [Google Scholar] [CrossRef] [PubMed]
  44. 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] [PubMed]
  45. Shen, Q.-B.; Gómez-García, C.-J.; Sun, W.-L.; Lai, X.-Y.; Pang, H.-J.; Ma, H.-Y. Improving the Photocatalytic H2 Evolution Activity of Keggin Polyoxometalates Anchoring Copper-azole Complexes. Green Chem. 2021, 23, 3104–3114. [Google Scholar] [CrossRef]
  46. Haraguchi, N.; Okaue, Y.; Isobe, T.; Matsuda, Y. Stabilization of Tetravalent Cerium upon Coordination of Unsaturated Heteropolytungstate Anions. Inorg. Chem. 1994, 33, 1015–1020. [Google Scholar] [CrossRef]
  47. Bi, L.-H.; Kortz, U.; Nellutla, S.; Stowe, A.-C.; Van Tol, J.; Dalal, N.-S.; Keita, B.; Nadjo, L. Structure, Electrochemistry, and Magnetism of the Iron(III)-substituted Keggin Dimer, [Fe6(OH)3(A-α-GeW9O34(OH)3)2]11−. Inorg. Chem. 2005, 44, 896–903. [Google Scholar] [CrossRef]
  48. 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]
  49. Sheldrick, G. SHELXT-Integrated space-group and crystal-structure determination. Acta Cryst. 2015, A71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  50. Sheldrick, G. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  51. Spek, A. PLATON SQUEEZE: A tool for the calculation of the disordered solvent contribution to the calculated structure factors. Acta Cryst. 2015, C71, 9–18. [Google Scholar] [CrossRef] [Green Version]
  52. Ibrahim, M.; Haider, A.; Lan, Y.-H.; Bassil, B.-S.; Carey, A.-M.; Liu, R.-J.; Zhang, G.-J.; Keita, B.; Li, W.-H.; Kostakis, G.E.; et al. Multinuclear Cobalt(II)-containing Heteropolytungstates: Structure, Magnetism, and Electrochemistry. Inorg. Chem. 2014, 53, 5179–5188. [Google Scholar] [CrossRef] [PubMed]
  53. Brown, I.-D.; Altermatt, D. Bond-valence Parameters Obtained from a Systematic Analysis of the Inorganic Crystal Structure Database. Acta Cryst. 1985, B41, 244–247. [Google Scholar] [CrossRef] [Green Version]
  54. Li, H.-L.; Zhang, M.; Lian, C.; Lang, Z.-L.; Lv, H.; 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]
  55. Paille, G.; Boulmier, A.; Bensaid, A.; Ha-Thi, M.-H.; Tran, T.-T.; Pino, T.; Marrot, J.; Rivière, E.; Hendon, C.H.; Oms, O.; et al. An Unprecedented {Ni14SiW9} Hybrid Polyoxometalate with High Photocatalytic Hydrogen Evolution Activity. Chem. Commun. 2019, 55, 4166–4169. [Google Scholar] [CrossRef] [Green Version]
  56. Wang, Z.-L.; Lu, Y.; Li, Y.-G.; Wang, S.-M.; Wang, E.-B. 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]
  57. Shang, X.-K.; Liu, R.-J.; Zhang, G.-J.; Zhang, S.-J.; Cao, H.-B.; Gu, Z.-J. Artificial photosynthesis for solar hydrogen generation over transition-metal substituted Keggin-type titanium tungstate. New J. Chem. 2014, 38, 1315–1320. [Google Scholar] [CrossRef]
Figure 1. The polyhedron diagrams of (a) polyoxoanion and counter cations {Co(H2O)6} in 1; (b) {[Co43-OH)3(PO4)]4(A-α-GeW9O34)4}; (c) four {A-α-GeW9O34}; (d) {Co163-OH)12O36(PO4)4}; (e) {Co4GeW9O34}. Color codes: WO6, red; GeO4, orange; CoO6, turquoise; PO4, yellow.
Figure 1. The polyhedron diagrams of (a) polyoxoanion and counter cations {Co(H2O)6} in 1; (b) {[Co43-OH)3(PO4)]4(A-α-GeW9O34)4}; (c) four {A-α-GeW9O34}; (d) {Co163-OH)12O36(PO4)4}; (e) {Co4GeW9O34}. Color codes: WO6, red; GeO4, orange; CoO6, turquoise; PO4, yellow.
Nanomaterials 13 02009 g001
Figure 2. (a) Polyhedron diagram of {Co163-OH)12O36(PO4)4} from the front view. The top view of (b) polyhedron diagram of {Co33-OH)6O9(PO4)}; (c) ball and stick diagram of {Co33-OH)6O9(PO4)}; (d) polyhedron diagram of {Co93-OH)9O19(PO4)3}; (e) ball and stick diagram of {Co93-OH)9O19-(PO4)3}; (f) polyhedron diagram of {Co43-OH)3O14}; (g) ball and stick diagram of {Co43-OH)3O14}. Color codes: CoO6, turquoise; PO4, yellow; Co atom, turquoise; P atom, yellow; O atom, red.
Figure 2. (a) Polyhedron diagram of {Co163-OH)12O36(PO4)4} from the front view. The top view of (b) polyhedron diagram of {Co33-OH)6O9(PO4)}; (c) ball and stick diagram of {Co33-OH)6O9(PO4)}; (d) polyhedron diagram of {Co93-OH)9O19(PO4)3}; (e) ball and stick diagram of {Co93-OH)9O19-(PO4)3}; (f) polyhedron diagram of {Co43-OH)3O14}; (g) ball and stick diagram of {Co43-OH)3O14}. Color codes: CoO6, turquoise; PO4, yellow; Co atom, turquoise; P atom, yellow; O atom, red.
Nanomaterials 13 02009 g002
Figure 3. The simplified arrangement diagrams of (a) the polyoxoanion in 1; (b) {Co163-OH)12-O36(PO4)4}; (c) {Co163-OH)12O52}. Color codes: GeW9, orange ball; Co4, turquoise tetrahedron; Co atom, turquoise; P atom, yellow; O atom, red.
Figure 3. The simplified arrangement diagrams of (a) the polyoxoanion in 1; (b) {Co163-OH)12-O36(PO4)4}; (c) {Co163-OH)12O52}. Color codes: GeW9, orange ball; Co4, turquoise tetrahedron; Co atom, turquoise; P atom, yellow; O atom, red.
Nanomaterials 13 02009 g003
Figure 4. The simplified arrangement diagrams of (a) the polyoxoanion in 1; (b) four {A-α-GeW9O34}; (c) four {PO4}; (d) {Co163-OH)12O52}. Color codes: GeW9 orange; Co atom, turquoise; P atom, yellow.
Figure 4. The simplified arrangement diagrams of (a) the polyoxoanion in 1; (b) four {A-α-GeW9O34}; (c) four {PO4}; (d) {Co163-OH)12O52}. Color codes: GeW9 orange; Co atom, turquoise; P atom, yellow.
Nanomaterials 13 02009 g004
Figure 5. (a) A series of blank experiments and the optimal reaction condition test of HER; (b) photocatalytic H2 evolution with different catalysts; (c) effect of the visible light towards the HER. Photocatalytic H2 evolution with different amounts of (d) 1; (e) photosensitizer; (f) TEOA. (Reaction conditions: 10 W white LED light, 1 (1, 3, 5 mg), TEOA (0.03, 0.050, 0.060 M), NHS (0.153 mM), H2O (2 M), [Ir(ppy)2(dtbbpy)][PF6] (0.1, 0.15, 0.2, 0.25 mM), CH3CN/DMF (1/3) as mixed solvent, the total reaction volume of 6 mL, stirring speed 400 rpm, at room temperature).
Figure 5. (a) A series of blank experiments and the optimal reaction condition test of HER; (b) photocatalytic H2 evolution with different catalysts; (c) effect of the visible light towards the HER. Photocatalytic H2 evolution with different amounts of (d) 1; (e) photosensitizer; (f) TEOA. (Reaction conditions: 10 W white LED light, 1 (1, 3, 5 mg), TEOA (0.03, 0.050, 0.060 M), NHS (0.153 mM), H2O (2 M), [Ir(ppy)2(dtbbpy)][PF6] (0.1, 0.15, 0.2, 0.25 mM), CH3CN/DMF (1/3) as mixed solvent, the total reaction volume of 6 mL, stirring speed 400 rpm, at room temperature).
Nanomaterials 13 02009 g005
Figure 6. (a) Long term photocatalysis of water reduction reaction test; (b) Recyclability tests of 1 towards hydrogen generation in three cycles. (Reaction conditions: 10 W white LED light, 1 (1 mg), TEOA (0.060 M), NHS (0.153 mM), [Ir(ppy)2(dtbbpy)][PF6] (0.2 mM), H2O (2 M), CH3CN/DMF (1/3) as mixed solvent, the total reaction volume of 6 mL, stirring speed 400 rpm, at room temperature); (c) FT−IR spectra of fresh and the isolated 1 after the catalytic cycles.
Figure 6. (a) Long term photocatalysis of water reduction reaction test; (b) Recyclability tests of 1 towards hydrogen generation in three cycles. (Reaction conditions: 10 W white LED light, 1 (1 mg), TEOA (0.060 M), NHS (0.153 mM), [Ir(ppy)2(dtbbpy)][PF6] (0.2 mM), H2O (2 M), CH3CN/DMF (1/3) as mixed solvent, the total reaction volume of 6 mL, stirring speed 400 rpm, at room temperature); (c) FT−IR spectra of fresh and the isolated 1 after the catalytic cycles.
Nanomaterials 13 02009 g006
Table 1. Crystallographic data and structure refinement for 1.
Table 1. Crystallographic data and structure refinement for 1.
1
Empirical formulaH169N5O375Na16P8Co36Ge8W72
Formula weight22,794.97
Temperature/K293(2)
Crystal systemTriclinic
Space groupP-1
a/Å24.4950 (8)
b/Å30.9701 (17)
c/Å36.8329 (17)
α/°72.049 (4)
β/°70.58
γ/°66.70
Volume/Å323,694.9 (19)
Z2
ρcalcg/cm33.113
μ/mm−119.239
F(000)19,358.0
Goodness-of-fit on F20.900
Final R indexes [I ≥ 2σ (I)] a,bR1 = 0.0643, wR2 = 0.1158
Final R indexes [all data] a,bR1 = 0.1428, wR2 = 0.1420
a R1 = ∑||Fo| − |Fc||/∑|Fo|; b wR2 = [Σw(Fo2Fc2)2]/Σw(Fo2)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

Zhao, Q.; Li, X.; Wang, Y.; Lv, H.; Yang, G. A Hexadecanuclear Cobalt-Added Tungstogermanate Containing Counter Cobalt Hydrates: Synthesis, Structure and Photocatalytic Properties. Nanomaterials 2023, 13, 2009. https://doi.org/10.3390/nano13132009

AMA Style

Zhao Q, Li X, Wang Y, Lv H, Yang G. A Hexadecanuclear Cobalt-Added Tungstogermanate Containing Counter Cobalt Hydrates: Synthesis, Structure and Photocatalytic Properties. Nanomaterials. 2023; 13(13):2009. https://doi.org/10.3390/nano13132009

Chicago/Turabian Style

Zhao, Qing, Xuyan Li, Yu Wang, Hongjin Lv, and Guoyu Yang. 2023. "A Hexadecanuclear Cobalt-Added Tungstogermanate Containing Counter Cobalt Hydrates: Synthesis, Structure and Photocatalytic Properties" Nanomaterials 13, no. 13: 2009. https://doi.org/10.3390/nano13132009

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