Selectively Photocatalytic Activity of an Open-Framework

1 day ago - The photocatalysis process with high selectivity is a very important research forefront for the semiconductor photocatalytic decomposition...
1 downloads 0 Views 3MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Selectively Photocatalytic Activity of an Open-Framework Chalcogenide Built from Corner-Sharing T4 Supertetrahedral Clusters Huan Pei,† Li Wang,*,† and Ming-Hua Zeng‡ †

College of Chemistry and Chemical Engineering, Xinjiang Normal University, Urumqi 830054, China Key Laboratory for the Synthesis and Application of Organic Functional Molecules, College of Chemistry and Chemical Engineering, Hubei University, Wuhan 430062, P. R. China

Downloaded via NOTTINGHAM TRENT UNIV on September 5, 2019 at 03:17:02 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: The photocatalysis process with high selectivity is a very important research forefront for the semiconductor photocatalytic decomposition of organic pollutants. However, the rational design of efficient photocatalysts with high selectivity is still a challenge. Here, we present an openframework chalcogenide (Heta)8[In14Sn2Zn4Se33] (Heta = ethanolamine-H+) (compound 1) constructed from T4 supertetrahedral clusters [In14Sn2Zn4Se35]12− with visiblelight-driven selectively photocatalytic degradation activity. Single-crystal XRD analysis shows that compound 1 crystallizes in I41/acd (no. 142) space group, with a = b = 24.3462(2) Å, c = 45.0062(9) Å, V = 26676.9(7) Å3, and Z = 8. Under visible-light irradiation, the selectively photocatalytic activities of 1 were evaluated by photodegradation of two kinds of cationic dye molecules, i.e., methylene blue (MB) and rhodamine B (RhB), against two anionic dyes, methyl orange (MO) and Kermes red (KR), with different sizes. We show that the adsorption capability and charge-matching between organic dyes and the supertetrahedral cluster together with a suitable band structure make it an excellent and selective photocatalyst. This is the first example of an open-framework chalcogenide based on supertetrahedral T4 for the selectively semiconductor photocatalytic decomposition of organic dyes.



separation and photocatalytic degradation of organic dyes.22,23 To our knowledge, crystalline open-framework chalcogenide superlattices based on the T4 cluster with selective photocatalytic activities for degrading organic pollutants have not been reported so far. In this work, an open-framework chalcogenide (Heta)8[In14Sn2Zn4Se33] (compound 1) constructed from T4 supertetrahedral clusters [In14Sn2Zn4Se35]12− has been prepared through solvothermal synthesis. Under visible-light irradiation, the selective photocatalytic activities of 1 were assessed by photodegradation of two kinds of cationic dye molecules (MB and RhB) against two anionic dyes (MO and KR) with different sizes. Their selectivity and photocatalytic performance are related to the adsorption capability and charge-matching between the organic dyes and the supertetrahedral cluster, together with suitable band structures.

INTRODUCTION Visible-light triggered degradation of organic pollutants with the aid of a semiconductor photocatalyst is of great interest because of the environmental friendly and generally mild conditions required for the process.1,2 Highly selective photocatalytic decomposition of mixed organic pollutants is, in particular, an important way for environmental remediation. The most applied method for enhancing photocatalytic selectivity is the modification of traditional nonporous photocatalysts of oxide, sulfide, etc.3−7 These methods focus on the control of morphology,8 phase,9 doping level,10 cocatalysts,11 structural composites,12 and surface treatment.13 However, rationally designing and synthesizing an efficient and highly selective photocatalyst remains a great challenge. Open-framework chalcogenides, as excellent candidates for visible-light photocatalysis as they possess both semiconductivity and porosity properties, have attracted increasing attention.14−17 A few open-framework chalcogenides based on tetrahedral clusters have been reportedly applied for photocatalytic decomposition of organic pollutants and hydrogen evolution.18,19 Despite successful efforts made in preparing robust open-framework chalcogenides,20,21 only a limited number of examples are available in the selective © XXXX American Chemical Society



EXPERIMENTAL SECTION

Materials and Synthesis. A mixture of Se (0.1737 g, 2.20 mmol), SnCl2·2H2O (0.0345 g, 0.15 mmol), In(NO3)3·5H2O (0.2332 g, 0.71

Received: April 10, 2019

A

DOI: 10.1021/acs.inorgchem.9b01040 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry mmol), Zn(NO3)2·5H2O (0.1164 g, 0.42 mmol), ethanolamine (eta) (4 mL), and H2O (2 mL) was placed in a 23 mL Teflon-lined stainless container, and the mixture was stirred for 1 h and heated at 160 °C for 4 days. Following slow cooling of the autoclave to 25 °C, pale-yellow octahedral crystals (Figure S1) of 1 were obtained (yield: 90% based on In). Elemental analysis for 1 (wt %) calcd: C, 3.68; N, 2.15; H, 1.38. Found: C, 3.71; N, 2.30; H, 1.26. Instrumentation and Characterization. Single crystal XRD data were collected at a temperature of 190 (2) K on a Bruker APEXII CCD system with Cu Kα radiation (λ = 1.54178 Å). The structure was solved by direct methods and refined by full-matrix least-squares method on F2 using SHELX algorithms in Olex2.24,25 All nonhydrogen atoms were refined with anisotropic displacement parameters. Tables S1−S4 show the crystal data, atomic coordinates, selected bond distances, and selected bond angles. Powder X-ray diffraction, infrared spectra, and thermogravimetric analysis (TG) are shown in Figures S2−S4. The powder XRD pattern was acquired on a Bruker D8 ADVANCE instrument using Cu Kα radiation (2θ = 5− 60°). FT-IR spectra were measured on a Shimadzu IR Affinity-1 instrument (400−4000 cm−1) at room temperature. The TG was measured on a NETZSCHSTA 449F3 instrument. The sample and reference (Al2O3) were heated from room temperature to 700 °C at a heating rate of 10 °C/min under a N2 atmosphere. The UV−vis−NIR diffuse reflectance spectrum was recorded on a Shimadzu SolidSpec3700DUV spectrophotometer over a 350−800 nm wavelength range. The UV−vis absorption spectrum was recorded on a U-3310 UV− visible absorption spectrophotometer. The sorption isotherm for N2 was performed using an Autosorb iQ1 (Quantachrome) adsorption analyzer. Prior to the measurement, the as-synthesized crystal sample was soaked in 0.5 M CsCl for a period of 2 days. After being filtered and dried in a vacuum oven overnight, the crystalline sample was further dried by using the “degas” function of the surface area analyzer for 6 h at 120 °C. Liquid N2 was used as the temperature control media at 77 K. Quantum Chemical Calculations. Calculation of DOS was carried out using DFT methods (ADF 2016.106 program).26 GGA with the PBE exchange-correlation functional,27 TZ2P Slater basis sets were used.28 Frozen core approximations were used for the inner shells [1s2−4p6] for In and Zn atoms and [1s2−3p6] for Se atoms. SR effects were considered by ZORA.29 The atoms charges were also calculated using Mulliken charge population,30 Hirshfeld,31 Voronoi,32 and MDC approaches.33 Photocatalytic Experiments. The selective photocatalytic activities of 1 were evaluated by photocatalyzed degradation of two cationic dyes (MB and RhB) and two anionic dyes (MO and KR) in aqueous solutions. A 300 W Xe lamp equipped with a cutoff filter (λ ≥ 400 nm) was employed as the light source. A 250 mL Pyrex beaker was used as a photochemical reactor. A 50 mg sample of 1 (without thermal pretreatment) was suspended in 100 mL of 1 × 10−5 M organic dye aqueous solutions.

Figure 1. (a) A typical [In14Sn2Zn4Se35]12− T4 cluster. (b) Single-set topological open framework. (c) Simplified 2-fold interpenetrating framework. (d) Actual 2-fold interpenetration that occurred in the three-dimensional frameworks.

shared corners of T4 clusters, compound 1 forms a threedimensional open-framework structure. If we do not consider atomic van der Waals radii, there are regular rectangular channels of 25.14 × 18.30 Å2 along the [100] direction in a single-set of the framework (Figure 1b). After the 2-fold interpenetration, there are rectangular channels of 25.14 × 9.10 Å2 along the [111] direction (Figure 1c and 1d), and the whole structure has 51% of the void space filled by protonated ethanolamines. Single-crystal and powder XRD, TG, IR, and elemental analyses establish the formula of (Heta)8[In14Sn2Zn4Se33]. This structure is a typical T4 cluster, which is similar to other reported T4 clusters, e.g., the undoped T4 [Zn4Ga16Se35] cluster34 and discrete chalcogenide [MxGa18−xSn2Q35]12−, etc.21 Infrared Spectroscopy. Hydroxyl group (−OH) peaks appeared at 3440 cm−1. Asymmetric stretching and symmetric bending vibrations of N−H group peaks appeared at 1603 and 1466 cm−1, respectively. The vibrational bands of the C−O group appeared at 1057 cm−1 (Figure S3). These IR data clearly indicate that ethanolaminium molecules are located in the framework, which is highly consistent with literature results.35 TG Analysis. Figure S4 of a TG curve shows that the loss of ∼10.0% (calcd as 9.5%) corresponds to the removal of ethanolaminium below ∼336 °C. Beyond 340 °C, the framework starts to decompose. The intensities of PXRD patterns obtained at 400 °C of 1 became weaker and some peaks disappeared, confirming the collapse of the framework integrity (Figure S2). UV−Vis Diffuse Reflectance Spectroscopy. Using the Kubelka−Munk function, optical diffuse reflectance spectra were converted to absorbance. By extrapolating the straight line, the Eg value of 1 was found to be 2.44 eV (Figure 2). Based on the equation EVB = X − Ee + 0.5Eg,36 the VB edge was estimated to be 1.53 eV, and the CB edge was estimated to be −0.91 eV. DFT Calculation. Figure 3 shows the total and partial DOS of 1, an indirect gap compound with a calculated band gap of 0.86 eV, which is smaller than the experimental one, probably due to the discontinuity in the derivative of exchange correlation energy within DFT.37 The top of VB is composed



RESULTS AND DISCUSSIONS Crystal Structure. Compound 1 crystallizes in the tetragonal space group I41/acd, and the asymmetric units consist of ten Se atoms, one Zn atom, and four In/Sn atoms. Crystallographic refinements, consideration of bond valence, and multiexperiments of ICP indicate that the four sites of In and Sn are distributed on one T4 cluster with the assigned site occupation factors converging to 0.875/0.125 for In/Sn (ICP measurements of Table S5). As shown in Figure 1a, there are 4 Se2− at four corners, 12 tri-coordinated Se2− on four faces, 18 bi-coordinated Se2− on six edges, and 1 tetra-coordinated Se2− at the center in a [In14Sn2Zn4Se35]12− T4 cluster. All In and mixed positions of In/Sn atoms are tetra-coordinated with 4 Se atoms, forming an (In/Sn)Se4 tetrahedron with (In/Sn)−Se distances ranging from 2.534 to 2.589 Å. The Zn atom is also tetra-coordinated by 4 Se atoms, forming a ZnSe4 tetrahedron with Zn−Se distances ranging from 2.442 to 2.455 Å. With B

DOI: 10.1021/acs.inorgchem.9b01040 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

concentration remained nearly at its initial level. Upon a prolonged irradiation time of 360 min, the residual concentration of RhB aqueous solution reached a level of 1.7%. The k (apparent rate constant) of MB is 0.66 min−1, which is 66 times that of RhB (0.01 min−1) (Figure S6). In comparison, under the irritation of the same period of 360 min, the residual concentrations of the anionic dyes, MO and KR, decreased slightly to 89% and 87%, respectively. Upon extending the irradiation time to 4320 min (3 days), no meaningful further decrease of C/C0 values for MO and KR was experimentally observed. Due to the electrostatic attraction, cationic dyes of MB+ and RhB+ are absorbed by anionic supertetrahedral clusters [In14Sn2Zn4Se35]12−, while MO− and KR− with an anionic charge are electrostatically repelled.38 On the other hand, the adsorption capability between different cationic dyes and the negatively charged supertetrahedral cluster also plays a role. The adsorption site of MB+ is the positively charged Ndimethylaminium moiety, while the RhB was absorbed on the supertetrahedral cluster through its positively charged Ndiethylaminium. Because the ethyl group is sterically larger than the methyl group, the adsorption capacity of RhB on the supertetrahedral clusters is weaker than that of MB (Scheme 1)

Figure 2. (a) Room-temperature UV−vis absorption spectrum of 1. (b) The corresponding plot of (F(R)hν)1/2 versus photon energy hν.

Scheme 1. Adsorption Sites of MB and RhB towards the Supertetrahedral Cluster

Figure 3. Calculated total and partial density of states for 1. The Fermi level is chosen as the energy reference at 0 eV.

of In-5p and Se-4p, together with a minor contribution from Sn-5p, while the bottom of CB is composed of In-5p and Se4p. Selective Photocatalysis. To study the selective photocatalytic decomposition of organic dyes, two cationic dyes (MB, 5.30 × 14.27 × 4.89 Å3; RhB, 11.79 × 14.87 × 6.76 Å3) and two anionic dyes (MO, 5.10 × 15.07 × 4.10 Å3; KR, 11.34 × 15.03 × 6.27 Å3) with different molecular sizes were selected (Figure S5). Figure 4a and 4b show the kinetic absorption spectra of photocatalytically degraded dye solutions in the presence of 1 under the visible-light irradiation. The residual concentration (C/C0) × 100% of a MB aqueous solution significantly decreased to 0.1% after 10 min of irradiation. On the other hand, after 10 min of light exposure, the RhB

due to the steric hindrance effect. The steric-hindranceinduced variation of adsorption capability could also be observed by the large difference between the residual

Figure 4. (a) Absorbance under the photocatalytic degradation conditions for MB and RhB. (b) C/C0 × 100% as a function of irradiation time for MB, RhB, MO, and KR. C

DOI: 10.1021/acs.inorgchem.9b01040 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

decreased to 49% after 10 min of visible-light irradiation and then to 1.3% after 30 min in the presence of CdS (Figure S8). We also compared the photocatalytic performance of 1 with commercial TiO2 (P25), and the apparent photodegradation rate constant k of 1 is 22 times of that of TiO2 (Figure S9). While comparing it with Fe3O4@MIL-100(Fe) MOF,43 the k of compound 1 is nearly 6 times higher under the same reaction conditions.

concentrations of MB (55.6%) and RhB (90.4%) with no light irradiation. Figure 5 indicates that the photocatalytic efficiency



CONCLUSION In summary, an open-framework chalcogenide (Heta)8[In14Sn2Zn4Se33] which can selectively decompose cationic MB and RhB, but against anionic dye MO and KR, was prepared. The selectivity and photocatalytic activity are closely related to the adsorption capability and charge-matching between organic dyes and supertetrahedral cluster and the favorable band structures in such a chalcogenide. This openframework chalcogenide demonstrated applicability in the selective and efficient photocatalytic area.

Figure 5. Performance stability of the MB/compound 1 system shown in three photocatalytic degradation cycles.

of compound 1 toward MB still remained high after 3 cycles. After the photocatalytic experiments, powder XRD and IR spectra show that compound 1 kept a good stability and that the protonated ethanolamine existed in channels and was not replaced the cationic dyes (Figures S2 and S3). We performed analysis for the sorption isotherm of nitrogen gas before and after Cs+-exchange of 1. Due to the pore blockage by bulky cationic agents, the as-synthesized 1 showed negligible gas adsorption. So far, few open-framework chalcogenides have exhibited microporosity by gas adsorption experiments because it is difficult to remove or exchange the guest organic molecules completely.39 In the semiconductor photocatalytic process, an electron transits from the valence band to the conduction band. The socalled photoelectron may be captured by an adsorbed O2 molecule to produce an •O2−. The left hole is captured by OH− to form an •OH radical.40,41 The radical •OH and/or •O2− then react with the organic dyes until the mineralization. On the basis of the above calculated band positions, the CB is −0.91 eV, which is more negative than E⊖ of O2/•O2− (−0.28 eV vs NHE), so the photogenerated electrons can reduce the surface chemically absorbed O2 to yield the strong oxidizing species •O2− (Figure 6).42 In order to confirm this



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.9b01040. (Table S1) Crystal data and structure refinement; (Table S2) final coordinates and equivalent isotropic displacement parameters; (Table S3) selected bond distances; (Table S4) selected bond angles; (Table S5) the In/Sn ratio determined by ICP-MS; (Figure S1) image for single crystals; (Figure S2) PXRD patterns; (Figure S3) IR spectra; (Figure S4) TG curve; (Figure S5) chemical structures and dimensions of different dye molecules; (Figure S6) linear plots for the photocatalytic degradation of MB and RhB; (Figure S7) control experiments; (Figure S8) photocatalytic activities; and (Figure S9) photocatalytic degradation curve and linear plots (PDF) Accession Codes

CCDC 1943161 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



Figure 6. Schematic representation of electron−hole pair formation and band-edge positions of 1.

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected].

•O2−

ORCID

scavenger (benzoquinone, BZQN) assumption, we did experiments. Figure S7 shows that the photodegradation efficiency of MB and RhB on 1 was sharply decreased respectively to 59.0% and 57.0% after adding •O2− scavenger BZQN. This result suggests that the superoxide radicals (•O2−) are the predominant active species contributing to the degradation of the organic dyes. We finally compared the photocatalytic decomposition performance of MB with visible-light-driven photocatalyst CdS (80.0 m2 g−1, 0.41 cm3 g−1) under the same condition. The residual concentration (C/C0)% of MB aqueous solution

Li Wang: 0000-0002-7425-6872 Ming-Hua Zeng: 0000-0002-7227-7688 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (51762039) and by the Tianshan XueSong Program (2017XS03). D

DOI: 10.1021/acs.inorgchem.9b01040 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry



(19) Zheng, N. F.; Bu, X. H.; Vu, H.; Feng, P. Y. Open-framework Chalcogenides as Visible-light Photocatalysts for Hydrogen Generation from Water. Angew. Chem., Int. Ed. 2005, 44, 5299−5303. (20) Li, H.; Kim, J.; Groy, T. L.; O’Keeffe, M.; Yaghi, O. M. 20 Å Cd4In16S3514‑ Supertetrahedral T4 Clusters as Building Units in Decorated Cristobalite Frameworks. J. Am. Chem. Soc. 2001, 123, 4867−4868. (21) Wu, T.; Wang, L.; Bu, X. H.; Chau, V.; Feng, P. Y. Largest Molecular Clusters in the Supertetrahedral Tn Series. J. Am. Chem. Soc. 2010, 132, 10823−10831. (22) Chen, X. T.; Bu, X. H.; Lin, Q. P.; Mao, C. Y.; Zhai, Q. G.; Wang, Y.; Feng, P. Y. Selective Ion Exchange and Photocatalysis by Zeolite-Like Semiconducting Chalcogenide. Chem. - Eur. J. 2017, 23, 11913−11919. (23) Zhao, X.; Bu, X. H.; Wu, T.; Zheng, S. T.; Wang, L.; Feng, P. Y. Selective Anion Exchange with Nanogated Isoreticular Positive Metalorganic Frameworks. Nat. Commun. 2013, 4, 2344. (24) Sheldrick, G. M. A Short History of SHELX. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112−122. (25) Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A.; Puschmann, H. OLEX2: A Complete Structure Solution, Refinement and Analysis Program. J. Appl. Crystallogr. 2009, 42, 339−341. (26) Te Velde, G.; Bickelhaupt, F. M.; Baerends, E. J.; Fonseca Guerra, C.; van Gisbergen, S. J. A.; Snijders, J. G.; Ziegler, T. Chemistry with ADF. J. Comput. Chem. 2001, 22, 931−967. (27) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868. (28) Van Lenthe, E.; Baerends, E. J. Optimized Slater-Type Basis Sets for the Elements 1−118. J. Comput. Chem. 2003, 24, 1142−1156. (29) Lenthe, E. v.; Baerends, E. J.; Snijders, J. G. Relativistic Regular Two-Component Hamiltonians. J. Chem. Phys. 1993, 99, 4597−4610. (30) Mulliken, R. S. Electronic Population Analysis on LCAO−MO Molecular Wave Functions. I. J. Chem. Phys. 1955, 23, 1833−1840. (31) Wiberg, K. B.; Rablen, P. R. Comparison of Atomic Charges Derived via Different Procedures. J. Comput. Chem. 1993, 14, 1504− 1518. (32) Bickelhaupt, F. M.; Van Eikema Hommes, N. J.; Fonseca Guerra, C.; Baerends, E. J. The Carbon−Lithium Electron Pair Bond in (CH3Li) n (n= 1, 2, 4). Organometallics 1996, 15, 2923−2931. (33) Swart, M.; Van Duijnen, P. T.; Snijders, J. G. A Charge Analysis Derived from An Atomic Multipole Expansion. J. Comput. Chem. 2001, 22, 79−88. (34) Wu, T.; Bu, X. H.; Zhao, X.; Khazhakyan, R.; Feng, P. Y. Phase Selection and Site-selective Distribution by Tin and Sulfur in Supertetrahedral Zinc Gallium Selenides. J. Am. Chem. Soc. 2011, 133, 9616−9625. (35) Pongstabodee, S.; Pornaroontham, P.; Pintuyothin, N.; Pootrakulchote, N.; Thouchprasitchai, N. CO2 Capture Performance of Bi-functional Activated Bleaching Earth Modified with Basicalcoholic Solution and Functionalization with Monoethanolamine: Isotherms, Kinetics and Thermodynamics. J. Environ. Sci. 2016, 48, 126−137. (36) Chen, Z. H.; Wang, W. L.; Zhang, Z. G.; Fang, X. M. Highefficiency Visible-light-driven Ag3PO4/AgI Photocatalysts: Z-scheme Photocatalytic Mechanism for Their Enhanced Photocatalytic Activity. J. Phys. Chem. C 2013, 117, 19346−19352. (37) Lin, Z. S.; Kang, L.; Zheng, T.; He, R.; Huang, H.; Chen, C. T. Strategy for the Optical Property Studies in Ultraviolet Nonlinear Optical Crystals from Density Functional Theory. Comput. Mater. Sci. 2012, 60, 99−104. (38) Yao, R. X.; Hailili, R.; Cui, X.; Wang, L.; Zhang, X. M. A Perfectly Aligned 63 Helical Tubular Cuprous Bromide Single Crystal for Selective Photo-catalysis, Luminescence and Sensing of Nitroexplosives. Dalton Trans. 2015, 44, 3410−3416. (39) Feng, P. Y.; Bu, X. H.; Zheng, N. F. The Interface Chemistry between Chalcogenide Clusters and Open Framework Chalcogenides. Acc. Chem. Res. 2005, 38, 293−303.

REFERENCES

(1) Chen, Z.; Xu, J.; Ren, Z.; He, Y.; Xiao, G. Low Temperature Synthesis of ZnIn2S4 Microspheres as A Visible Light Photocatalyst for Selective Oxidation. Catal. Commun. 2013, 41, 83−86. (2) Inumaru, K.; Yasui, M.; Kasahara, T.; Yamaguchi, K.; Yasuda, A.; Yamanaka, S. Nanocomposites of Crystalline TiO2 Particles and Mesoporous Silica: Molecular Selective Photocatalysis Tuned by Controlling Pore Size and Structure. J. Mater. Chem. 2011, 21, 12117−12125. (3) Liu, S.; Yu, J.; Jaroniec, M. Tunable Photocatalytic Selectivity of Hollow TiO2 Microspheres Composed of Anatase Polyhedra with Exposed {001} Facets. J. Am. Chem. Soc. 2010, 132, 11914−11916. (4) Shiraishi, Y.; Sugano, Y.; Ichikawa, S.; Hirai, T. Visible Light Induced Partial Oxidation of Cyclohexane on WO3 Loaded with Pt Nanoparticles. Catal. Sci. Technol. 2012, 2, 400−405. (5) Qamar, M.; Elsayed, R. B.; Alhooshani, K. R.; Ahmed, M. I.; Bahnemann, D. W. Chemoselective and Highly Efficient Conversion of Aromatic Alcohols into Aldehydes Photo-Catalyzed by Ag3PO4 in Aqueous Suspension under Simulated Sunlight. Catal. Commun. 2015, 58, 34−39. (6) Liu, Y.; Huang, B.; Dai, Y.; Zhang, X.; Qin, X.; Jiang, M.; Whangbo, M. H. Selective Ethanol Formation from Photocatalytic Reduction of Carbon Dioxide in Water with BiVO4 Photocatalyst. Catal. Commun. 2009, 11, 210−213. (7) Ye, L.; Su, Y.; Jin, X.; Xie, H.; Zhang, C. Recent Advances in BiOX (X = Cl, Br and I) Photocatalysts: Synthesis, Modification, Facet Effects and Mechanisms. Environ. Sci.: Nano 2014, 1, 90−112. (8) Ikeda, S.; Ikoma, Y.; Kobayashi, H.; Harada, T.; Torimoto, T.; Ohtani, B.; Matsumura, M. Encapsulation of Titanium(IV) Oxide Particles in Hollow Silica for Size-Selective Photocatalytic Reactions. Chem. Commun. 2007, 36, 3753−3755. (9) Tian, J.; Zhao, Z.; Kumar, A.; Boughton, R. I.; Liu, H. Recent Progress in Design, Synthesis and Applications of One-Dimensional TiO2 Nanostructured Surface Heterostructures: A Review. Chem. Soc. Rev. 2014, 43, 6920−6937. (10) Devi, L. G.; Kavitha, R. A Review on Non Metal Ion Doped Titania for the Photocatalytic Degradation of Organic Pollutants under UV/Solar Light: Role of Photogenerated Charge Carrier Dynamics in Enhancing the Activity. Appl. Catal., B 2013, 140, 559− 587. (11) Abedi, S.; Morsali, A. Ordered Mesoporous Metal Organic Frameworks Incorporated with Amorphous TiO2 As Photocatalyst for Selective Aerobic Oxidation in Sunlight Irradiation. ACS Catal. 2014, 4, 1398−1403. (12) Zhang, N.; Zhang, Y.; Pan, X.; Yang, M. Q.; Xu, Y. J. Constructing Ternary CdS-Graphene-TiO2 Hybrids on the Flatland of Graphene Oxide with Enhanced Visible-Light Photoactivity for Selective Transformation. J. Phys. Chem. C 2012, 116, 18023−18031. (13) Tang, Z. R.; Yin, X.; Zhang, Y.; Xu, Y. J. One-Pot, High-Yield Synthesis of One-Dimensional ZnO Nanorods with Well-Defined Morphology as A Highly Selective Photocatalyst. RSC Adv. 2013, 3, 5956−5965. (14) Davis, M. E. Ordered Porous Materials for Emerging Applications. Nature 2002, 417, 813. (15) Cheetham, A. K.; Férey, G.; Loiseau, T. Open-framework Inorganic Materials. Angew. Chem., Int. Ed. 1999, 38, 3268−3292. (16) Osterloh, F. E. Inorganic Materials as Catalysts for Photochemical Splitting of Water. Chem. Mater. 2008, 20, 35−54. (17) Mao, C. Y.; Wang, Y. X.; Jiao, W.; Chen, X. T.; Lin, Q. P.; Deng, M. L.; Ling, Y.; Zhou, Y. M.; Bu, X. H.; Feng, P. Y. Integrating Zeolite-Type Chalcogenide with Titanium Dioxide Nanowires for Enhanced Photoelectrochemical Activity. Langmuir 2017, 33, 13634− 13639. (18) Yue, C. Y.; Lei, X. W.; Feng, L. J.; Wang, C.; Gong, Y. P.; Liu, X. Y. [Mn2Ga4Sn4S20]8− T3 Supertetrahedral Nanocluster Directed by a Series of Transition Metal Complexes. Dalton Trans. 2015, 44, 2416−2424. E

DOI: 10.1021/acs.inorgchem.9b01040 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (40) Adhikari, S.; Charanpahari, A. V.; Madras, G. Solar-light-driven Improved Photocatalytic Performance of Hierarchical ZnIn2S4 Architectures. ACS Omega 2017, 2, 6926−6938. (41) Lee, Y. Y.; Moon, J. H.; Choi, Y. S.; Park, G. O.; Jin, M.; Jin, L. Y.; Li, D. H.; Lee, J. Y.; Son, S. U.; Kim, J. M. Visible-Light Driven Photocatalytic Degradation of Organic Dyes over Ordered Mesoporous CdxZn1−xS Materials. J. Phys. Chem. C 2017, 121, 5137−5144. (42) Kundu, J.; Khilari, S.; Pradhan, D. Shape-dependent Photocatalytic Activity of Hydrothermally Synthesized Cadmium Sulfide Nanostructures. ACS Appl. Mater. Interfaces 2017, 9, 9669−9680. (43) Zhang, C. F.; Qiu, L. G.; Ke, F.; Zhu, Y. J.; Yuan, Y. P.; Xu, G. S.; Jiang, X. A Novel Magnetic Recyclable Photocatalyst Based on a Core−shell Metal−organic Framework Fe3O4@MIL-100(Fe) for the Decolorization of Methylene Blue Dye. J. Mater. Chem. A 2013, 1, 14329−14334.

F

DOI: 10.1021/acs.inorgchem.9b01040 Inorg. Chem. XXXX, XXX, XXX−XXX