Encapsulation of Phosphotungstic Acid into Metal–Organic

Oct 3, 2018 - Fujian Provincial Key Laboratory of Agroecological Processing and Safety Monitoring, College of Life Sciences, Fujian Agriculture and Fo...
0 downloads 0 Views 5MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Encapsulation of Phosphotungstic Acid into Metal−Organic Frameworks with Tunable Window Sizes: Screening of PTA@MOF Catalysts for Efficient Oxidative Desulfurization Zu-Jin Lin,*,†,‡,§ He-Qi Zheng,‡,§ Jin Chen,§ Wan-E Zhuang,§ Yue-Xu Lin,§ Jin-Wei Su,§ Yuan-Biao Huang,‡ and Rong Cao*,‡

Inorg. Chem. Downloaded from pubs.acs.org by UNIV OF SUNDERLAND on 10/03/18. For personal use only.



Fujian Provincial Key Laboratory of Agroecological Processing and Safety Monitoring, College of Life Sciences, Fujian Agriculture and Forestry University, Fuzhou, Fujian 350002, People’s Republic of China ‡ State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou, Fujian 350002, People’s Republic of China § Department of Applied Chemistry, College of Life Sciences, Fujian Agriculture and Forestry University, Fuzhou, Fujian 350002, People’s Republic of China S Supporting Information *

ABSTRACT: Clean fuels with extremely low sulfur content are highly desirable due to environmental concerns. Herein, three water-stable and ecofriendly metal−organic frameworks with tunable window diameters, denoted as MOF-808X, have been employed as PTA solid supports. An array of PTA@ MOF-808X composites were facilely synthesized via the encapsulation strategy. With tunable window sizes and adjustable PTA loading amounts, the obtained PTA@MOF-808X composites were screened for catalytic oxidative desulfurization (ODS) with H2O2 serving as oxidant. The experiments found that 42% PTA@MOF-808A had the highest catalytic ODS activity and could completely remove dibenzothiophene (DBT) in a model fuel with an initial sulfur content of 1000 ppm within 30 min, which falls far below the acceptable limits for fuel standards (10 ppm). Further investigations revealed that this high catalytic activity could be attributed to the cooperative catalysis of metal clusters in the host framework and the guest PTA molecules. Moreover, 42%PTA@MOF-808A could be facilely recovered and reused for at least five runs without loss of catalytic activity. Having a combination of eco-sustainability, high stability, high catalytic activity, and good recyclability, 42%PTA@MOF-808A therefore represents a new benchmark material for catalytic ODS and provides a new perspective for ultradeep desulfurization.

1. INTRODUCTION The combustion of sulfur species in fossil fuels results in the massive emission of hazardous sulfur oxides (SOx), which leads to air pollution, acid rain, equipment corrosion, and catalyst deactivation. Due to environmental concerns, fuels with a maximum sulfur concentration of 10 ppm have been strictly implemented in some countries of Europe and the USA.1,2 Furthermore, an extremely low sulfur content ( 0.98) demonstrated that the ODS over 42%PTA@MOF-808A obeys a pseudo-first-order kinetic model. As expected, the kinetic constant (k) increased with the increase in reaction temperature, and the value reached 0.16 min−1 at 60 °C. As shown in Table 2, this value is the largest among the reported excellent heterogeneous catalysts for DBT oxidation. The large kinetic constant further confirmed that 42%PTA@MOF-808A is an effective catalyst for ODS. The half-life of the reaction (t1/2) calculated from the pseudo-first-order kinetic equation under the optimal conditions is only 4.3 min, which is the smallest among the reported representative catalysts (Table 2), indicating the high efficiency of the 42%PTA@MOF-808A composite in catalytic DBT oxidation. The turnover frequency (TOF) was calculated to quantitatively compare the catalytic oxidative activity with that of other catalysts, and the value reached 48.6 h−1 on the basis of the number of active PTA sites. The value is considerably larger than those of representative Ti-containing catalysts such as TiO2/porous glass (3.7 h−1), where H2O2 was used as oxidant,12 and is comparable to that of the benchmark of the Ti-containing mesoporous silica catalyst Ti-SBA-2 (48.8 h−1), where tertbutyl hydroperoxide was used as oxidant.56 On the basis of the

Figure 8. (a) Kinetic profiles for the DBT oxidation over 42%PTA@ MOF-808A at various temperatures. (b) Pseudo-first-order kinetic plots for the oxidation of DBT at various temperatures over 42% PTA@MOF-808A catalyst.

kinetic constants at various temperatures, the apparent activation energy (Ea) for the DBT oxidation over 42% PTA@MOF was 29 kJ mol−1, which was estimated using the Arrhenius equation (Figure S11). This value is one of the lowest among the reported apparent activation energies (Table 2), which further confirmed the high effectiveness of the catalyst. To illustrate the relationship between the structure of the substrate and the catalytic activity over 42%PTA@MOF-808A, the catalytic oxidations of other refractory organosulfur compounds including BT and 4,6-dimethyldibenzothiophene (DMDBT) were also conducted under the optimal reaction conditions. Figure 9 and Figure S12 show that both the substrate conversions and the rate constants followed the same order: DBT > BT > DMDBT. Two key factors, the electron density of sulfur atoms and the steric hindrance around the sulfur atoms in the substrates, would largely affect the ODS activity. The sulfur atom electron densities of DMDBT, DBT, and BT are 5.760, 5.758, and 5.739, respectively.5 In principle, high electron density on sthe ulfur atom benefits its oxidation. Therefore, the low electron density on the sulfur atom of BT (5.739) contributes to its low oxidation activity. That is why the ODS activity of BT is lower than that of DBT. As the electron densities on sulfur atoms of DBT (5.758) and DMDBT (5.760) are almost identical with each other, the lower ODS activity of DMDBT in comparison to DBT could be ascribed to the relatively large steric hindrance around the sulfur atom in DMDBT. What is more, the size of the substrates may also exert an important effect on the activity in this ODS system. DMDBT has the largest size among these three substrates (Table S2), and its size is closest to the window size of 42%PTA@MOF-808A. Therefore, the mass transfer of DMDBT and its oxidative products within the G

DOI: 10.1021/acs.inorgchem.8b02272 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Table 2. Comparison of the Pseudo-First-Order Kinetic Constants (k) and the Apparent Activation Energies (Ea) for ODS of DBT over Various Reported Excellent Catalysts with H2O2 Serving as Oxidant catalyst

T (°C)

k (min−1)

t1/2 (min)a

Ea (kJ mol−1)

ref

42%PTA@MOF-808A 2PTA@MIL-53(Al)-NH2 PTA@MIL-101(Fe) MDC-C MDC-P mesoporous WO3/TiO2 HPW-IL/SBA-15

60 60 70 80 80 50 60

0.16 ̵ 0.065 0.12 0.041 0.11 0.14

4.3 ̵ 10.7 5.8 16.9 6.3 4.9

29.0 34.1 ̵ 19 25 54 56.6

this work 47 5 57 57 58 59

a

t1/2 = ln 2/k.

DBT oxidation. In view of the fact that almost no activity was found in H3BTC ligands, the oxidation activity of MOF-808A could be attributed to its Zr6O4(μ3-OH)4(A)6 clusters. Very recently, some work has reported that the Zr−OH(OH2) can catalyze the decomposition of H2O2 into active •OH, thus promoting the oxidation reactions.60 The catalytic oxidation activity of MOF-808A may also occur by this mechanism. Meanwhile, the low catalytic activity of pure PTA may result in its low special surface area. Remarkably, the oxidative activity of a physical mixture of PTA and MOF-808A was almost equal to that of pure MOF-808A and was obviously lower than that of the 42%PTA@MOF-808A composite. The PTA molecules in the solution may have strong interactions with the surface of MOF-808A particles, thus partially blocking the channels of MOF-808A. This is why the catalytic activity of the PTA and MOF-808A mixture was lower than that of MOF-808A. Remarkably, the oxidative activity of 42%PTA@MOF-808A is not only higher than that of MOF-808A or PTA or their physical mixture but also higher than the total activity of pure MOF-808A and PTA. Such a high catalytic oxidation activity can be mainly attributed to the uniform and good dispersion of PTA molecules in the framework, which can serve as single active sites and thus speed the ODS process. In addition, the enhancement of the catalytic activity of the 42%PTA@MOF808A composite may also result from the synergistic effect of the host MOF-808A framework and the encapsulated PTA molecules.61 On the basis of the above experimental results and some literature reports,11,62,63 a possible mechanism was illuminated for the ODS over the 42%PTA@MOF-808A catalyst. First, the DBT in the oil phase quickly diffused into the polar acetonitrile phase. Then, the porous 42%PTA@MOF-808A catalyst in the polar phase adsorbed both the DBT and H2O2 into its pores, thus enriching their concentration in the framework. Some of the adsorbed H2O2 could readily oxidize the loaded PTA to W(VI)−peroxo, which can oxidize the DBT to sulfone.42 In view of the moderate surface area and high PTA dispersion, the encapsulated PTA molecules in the composite may act as single active sites to improve the oxidation activity. In addition, H2O2 also can be decomposed to active •OH by Zr− OH(OH2) on the Zr6 clusters, and the obtained active •OH can directly oxidize refractory sulfur-containing compounds such as DBT to the corresponding sulfones.64,65 Therefore, the cooperative catalysis of PTA and metal nodes of the framework contributed to the high catalytic oxidation activity. 3.6. Recycling of the Catalyst. The recyclability of a heterogeneous catalyst is extremely vital for its practical industrial applications. In order to investigate the reusability of PTA@MOF-808X, 42%PTA@MOF-808A was selected as a model catalyst and the ODS of DBT was chosen as a model

Figure 9. Comparison of kinetic profiles for the DBT, BT, and DMDBT oxidations with initial sulfur contents of 1000 ppm over 42% PTA@MOF-808A at 60 °C.

catalyst was slowed down by the windows, leading to the lowest activity. 3.5. Plausible Mechanism. To understand the oxidation mechanism over 42%PTA@MOF-808A, control experiments were carried out. As shown in Figure 10, 42%PTA@MOF-

Figure 10. Control experiments conducted under the optimal conditions.

808A had an excellent catalytic activity for DBT oxidation, while no significant DBT oxidations were observed in the absence of either H2O2 or catalyst, demonstrating the indispensability of both H2O2 and catalyst to promote the DBT oxidation. In addition, the removal of 42%PTA@MOF808A catalyst after 5 min completely shut down the oxidative reaction, giving an identical DBT conversion after stirring for another 25 min. The result demonstrated the heterogeneous nature of 42%PTA@MOF-808A catalyst and no leaching of PTA molecules into the polar solution. Further experiments showed that MOF-808A and PTA could separately catalyze the H

DOI: 10.1021/acs.inorgchem.8b02272 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry reaction. 42%PTA@MOF-808A was first recovered by simple centrifugation and then subjected to the next ODS runs under optimal reaction conditions. As shown in Figure 11, no

constant in the DBT oxidation, further illustrating its high efficiency in catalytic oxidative desulfurization. Further investigation revealed that both the highly dispersed PTA and the uniformly decorated Zr−OH(H2O) in zirconium clusters contributed to the high catalytic ODS activity. Moreover, 42%PTA@MOF-808A could be easily recovered and reused for at least 5 runs without significant loss of catalytic activity. No PTA leaching was observed in the recycling experiments, confirming the heterogeneous nature of the catalyst. Considering the high stability and ecosustainability, 42%PTA@MOF-808A is the best heterogeneous catalyst among the POM@MOF catalysts and is comparable to other reported excellent materials in catalytic ODS. With the addition of facile synthesis, excellent recyclability, and low cost, 42%PTA@MOF-808A therefore represents a new benchmark material for catalytic ODS and provides a new perspective for ultradeep desulfurization.



Figure 11. Recycled experiments of 42%PTA@MOF-808A.

ASSOCIATED CONTENT

S Supporting Information *

significant decrease in catalytic activity was observed and the DBT conversion still could reach 100% after five successive catalytic runs under the optimal reaction conditions, which could be attributed to the high stability of the supports and the heterogeneous nature of the catalyst. In addition, the PTA content in the recycled sample was almost identical with that in the pristine composite (Table S1). This is because the window size of MOF-808A is smaller than that of a PTA molecule, thus effectively inhibiting the PTA leaching. The similar IR spectra (Figure S13) and PXRD patterns (Figure S14) of the recovered and pristine samples further confirmed the good retention of the catalyst structure after five catalytic runs. N2 sorption isotherms showed that only a slight reduction in the BET surface area was found for the recycled sample (Figure S15). This small decrease in surface area can be ascribed to the adsorption of the substrate or oxidized products on the catalyst. Moreover, SEM images revealed that the morphology of the catalyst was well retained after five successive catalytic runs (Figure S16). These results unambiguously confirmed the stability of the catalyst and its facile recycling and reuse for at least five runs. The combination of high catalytic efficiency and the easy reusability makes 42%PTA@MOF-808A a promising heterogeneous catalyst for application in ODS for fossil fuels.

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b02272.



Materials and methods, ICP data, PXRD patterns, N2 isotherms, TGA curves, solid-state 31P MAS NMR spectra, SEM-EDS elemental mapping, and other complementary data (PDF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail for Z.-J.L.: [email protected]. *E-mail for R.C.: [email protected]. ORCID

Zu-Jin Lin: 0000-0003-2515-3356 He-Qi Zheng: 0000-0002-8848-1165 Yuan-Biao Huang: 0000-0003-4680-2976 Rong Cao: 0000-0002-2398-399X Notes

The authors declare no competing financial interest.





ACKNOWLEDGMENTS This work was financially supported by the National Natural Science Foundation of China (NSFC) (21520102001, 21521061, and 21331006), Fujian Agriculture and Forestry University (118360020), Outstanding Youth Research Training Program of Fujian Agriculture and Forestry University (XJQ201616), FJ 2011 Program (2015-75), and the State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences (20170028).

CONCLUSIONS In summary, we have employed three highly stable and ecosustainable frameworks to serve as PTA supports. Through an encapsulation strategy, various amounts of PTA were successfully loaded into MOF-808X to form series of PTA@ MOF-808X composites. PTA@MOF-808X composites have been screened for catalytic oxidative desulfurization with green H2O2 being used as oxidant. The results revealed that 42% PTA@MOF-808A possesses the highest catalytic activity. The optimal catalytic conditions of 42%PTA@MOF-808A were investigated, and the result is 0.8 mol % catalyst dosage, 60 °C, and an O/S molar ratio of 5. Under these conditions, 42% PTA@MOF-808A can completely remove DBT in a model fuel with an initial sulfur content of 1000 ppm within 30 min. In addition to high chemical stability, 42%PTA@MOF-808A is either more efficient or more environmentally friendly than other reported POM@MOF catalysts for catalytic ODS. Kinetic investigations revealed that BT, DBT, and DMDBT oxidation over 42%PTA@MOF-808A followed a pseudo-firstorder model. 42%PTA@MOF-808A had a large kinetic



REFERENCES

(1) Stanislaus, A.; Marafi, A.; Rana, M. S. Recent advances in the science and technology of ultra low sulfur diesel (ULSD) production. Catal. Today 2010, 153, 1−68. (2) Gasoline Sulfur; Gasoline Standards, U.S. Environmental Protection Agency: https://www.epa.gov/gasoline-standards/ gasoline-sulfur, accessed May 1, 2018. (3) Song, C. An overview of new approaches to deep desulfurization for ultra-clean gasoline, diesel fuel and jet fuel. Catal. Today 2003, 86, 211−263.

I

DOI: 10.1021/acs.inorgchem.8b02272 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Dependent Catalytic Performance. Inorg. Chem. 2017, 56, 13634− 13637. (22) Pan, M.; Zhu, Y.-X.; Wu, K.; Chen, L.; Hou, Y.-J.; Yin, S.-Y.; Wang, H.-P.; Fan, Y.-N.; Su, C.-Y. Epitaxial Growth of Hetero-LnMOF Hierarchical Single Crystals for Domain- and OrientationControlled Multicolor Luminescence 3D Coding Capability. Angew. Chem., Int. Ed. 2017, 56, 14582−14586. (23) Liao, W.-M.; Zhang, J.-H.; Yin, S.-Y.; Lin, H.; Zhang, X.; Wang, J.; Wang, H.-P.; Wu, K.; Wang, Z.; Fan, Y.-N.; Pan, M.; Su, C.-Y. Tailoring exciton and excimer emission in an exfoliated ultrathin 2D metal-organic framework. Nat. Commun. 2018, 9, 2401. (24) Chen, L.; Ye, J.-W.; Wang, H.-P.; Pan, M.; Yin, S.-Y.; Wei, Z.W.; Zhang, L.-Y.; Wu, K.; Fan, Y.-N.; Su, C.-Y. Ultrafast water sensing and thermal imaging by a metal-organic framework with switchable luminescence. Nat. Commun. 2017, 8, 15985. (25) Chen, Y.-Z.; Jiang, H.-L. Porphyrinic Metal−Organic Framework Catalyzed Heck-Reaction: Fluorescence “Turn-On” Sensing of Cu(II) Ion. Chem. Mater. 2016, 28, 6698−6704. (26) Yang, Q.; Xu, Q.; Jiang, H.-L. Metal-organic frameworks meet metal nanoparticles: synergistic effect for enhanced catalysis. Chem. Soc. Rev. 2017, 46, 4774−4808. (27) Yuan, S.; Qin, J.-S.; Xu, H.-Q.; Su, J.; Rossi, D.; Chen, Y.; Zhang, L.; Lollar, C.; Wang, Q.; Jiang, H.-L.; Son, D. H.; Xu, H.; Huang, Z.; Zou, X.; Zhou, H.-C. [Ti8Zr2O12(COO)16] Cluster: An Ideal Inorganic Building Unit for Photoactive Metal−Organic Frameworks. ACS Cent. Sci. 2018, 4, 105−111. (28) Zhang, L.; Yuan, S.; Feng, L.; Guo, B.; Qin, J.-S.; Xu, B.; Lollar, C.; Sun, D.; Zhou, H.-C. Pore-Environment Engineering with Multiple Metal Sites in Rare-Earth Porphyrinic Metal−Organic Frameworks. Angew. Chem., Int. Ed. 2018, 57, 5095−5099. (29) Huang, N.; Wang, K.; Drake, H.; Cai, P.; Pang, J.; Li, J.; Che, S.; Huang, L.; Wang, Q.; Zhou, H.-C. Tailor-Made Pyrazolide-Based Metal−Organic Frameworks for Selective Catalysis. J. Am. Chem. Soc. 2018, 140, 6383−6390. (30) Li, X.-M.; Dong, L.-Z.; Li, S.-L.; Xu, G.; Liu, J.; Zhang, F.-M.; Lu, L.-S.; Lan, Y.-Q. Synergistic Conductivity Effect in a Proton Sources-Coupled Metal−Organic Framework. ACS Energy Lett. 2017, 2, 2313−2318. (31) Lian, X.; Huang, Y.; Zhu, Y.; Fang, Y.; Zhao, R.; Joseph, E.; Li, J.; Pellois, J.-P.; Zhou, H.-C. Enzyme-MOF Nanoreactor Activates Nontoxic Paracetamol for Cancer Therapy. Angew. Chem., Int. Ed. 2018, 57, 5725−5730. (32) Cai, G.; Zhang, W.; Jiao, L.; Yu, S.-H.; Jiang, H.-L. TemplateDirected Growth of Well-Aligned MOF Arrays and Derived SelfSupporting Electrodes for Water Splitting. Chem. 2017, 2, 791−802. (33) Liao, W.-M.; Zhang, J.-H.; Wang, Z.; Yin, S.-Y.; Pan, M.; Wang, H.-P.; Su, C.-Y. Post-synthetic exchange (PSE) of UiO-67 frameworks with Ru/Rh half-sandwich units for visible-light-driven H2 evolution and CO2 reduction. J. Mater. Chem. A 2018, 6, 11337−11345. (34) Li, M.-T.; Yang, X.-Y.; Li, J.-S.; Sheng, N.; Liu, G.-D.; Sha, J.Q.; Lan, Y.-Q. Assembly of Multifold Helical Polyoxometalate-Based Metal−Organic Frameworks as Anode Materials in Lithium-Ion Batteries. Inorg. Chem. 2018, 57, 3865−3872. (35) Yang, X.-Y.; Wei, T.; Li, J.-S.; Sheng, N.; Zhu, P.-P.; Sha, J.-Q.; Wang, T.; Lan, Y.-Q. Polyoxometalate-Incorporated Metallapillararene/Metallacalixarene Metal-Organic Frameworks as Anode Materials for Lithium Ion Batteries. Inorg. Chem. 2017, 56, 8311−8318. (36) Zhang, A. M.; Zhang, M.; Lan, D.; Wang, H.-N.; Tang, Y.-J.; Wang, X.-L.; Dong, L.-Z.; Zhang, L.; Li, S.-L.; Lan, Y.-Q. Polyoxometalate-Based Metal−Organic Framework on Carbon Cloth with a Hot-Pressing Method for High-Performance LithiumIon Batteries. Inorg. Chem. 2018, 57, 11726. (37) Cheng, W.; Shen, F.-C.; Xue, Y.-s.; Luo, X.; Fang, M.; Lan, Y.Q.; Xu, Y. A Pair of Rare Three-Dimensional Chiral PolyoxometalateBased Metal−Organic Framework Enantiomers Featuring Superior Performance as the Anode of Lithium-Ion Battery. ACS Appl. Energy Mater. 2018, 1, 4931.

(4) van Haandel, L.; Bremmer, G. M.; Hensen, E. J. M.; Weber, T. The effect of organic additives and phosphoric acid on sulfidation and activity of (Co)Mo/Al2O3 hydrodesulfurization catalysts. J. Catal. 2017, 351, 95−106. (5) Wang, X. S.; Li, L.; Liang, J.; Huang, Y. B.; Cao, R. Boosting Oxidative Desulfurization of Model and Real Gasoline over Phosphotungstic Acid Encapsulated in Metal−Organic Frameworks: The Window Size Matters. ChemCatChem 2017, 9, 971−979. (6) Li, S.-W.; Yang, Z.; Gao, R.-M.; Zhang, G.; Zhao, J.-s. Direct synthesis of mesoporous SRL-POM@MOF-199@MCM-41 and its highly catalytic performance for the oxidesulfurization of DBT. Appl. Catal., B 2018, 221, 574−583. (7) Ahmed, I.; Jhung, S. H. Adsorptive desulfurization and denitrogenation using metal-organic frameworks. J. Hazard. Mater. 2016, 301, 259−276. (8) Dinamarca, M. A.; Rojas, A.; Baeza, P.; Espinoza, G.; IbacacheQuiroga, C.; Ojeda, J. Optimizing the biodesulfurization of gas oil by adding surfactants to immobilized cell systems. Fuel 2014, 116, 237− 241. (9) Kumar, S.; Srivastava, V. C.; Nanoti, S. M. Extractive Desulfurization of Gas Oils: A Perspective Review for Use in Petroleum Refineries. Sep. Purif. Rev. 2017, 46, 319−347. (10) Ibrahim, M. H.; Hayyan, M.; Hashim, M. A.; Hayyan, A. The role of ionic liquids in desulfurization of fuels: A review. Renewable Sustainable Energy Rev. 2017, 76, 1534−1549. (11) Gomez-Paricio, A.; Santiago-Portillo, A.; Navalon, S.; Concepcion, P.; Alvaro, M.; Garcia, H. MIL-101 promotes the efficient aerobic oxidative desulfurization of dibenzothiophenes. Green Chem. 2016, 18, 508−515. (12) Shen, C.; Wang, Y. J.; Xu, J. H.; Luo, G. S. Oxidative desulfurization of DBT with H2O2 catalysed by TiO2/porous glass. Green Chem. 2016, 18, 771−781. (13) Wang, X.-S.; Huang, Y.-B.; Lin, Z.-J.; Cao, R. Phosphotungstic acid encapsulated in the mesocages of amine-functionalized metalorganic frameworks for catalytic oxidative desulfurization. Dalton Trans 2014, 43, 11950−11958. (14) Chen, L.; Luque, R.; Li, Y. Controllable design of tunable nanostructures inside metal-organic frameworks. Chem. Soc. Rev. 2017, 46, 4614−4630. (15) Julião, D.; Gomes, A. C.; Pillinger, M.; Valença, R.; Ribeiro, J. C.; Gonçalves, I. S.; Balula, S. S. A recyclable ionic liquidoxomolybdenum(vi) catalytic system for the oxidative desulfurization of model and real diesel fuel. Dalton Trans 2016, 45, 15242−15248. (16) Li, J.-K.; Dong, J.; Wei, C.-P.; Yang, S.; Chi, Y.-N.; Xu, Y.-Q.; Hu, C.-W. Controllable Synthesis of Lindqvist Alkoxopolyoxovanadate Clusters as Heterogeneous Catalysts for Sulfoxidation of Sulfides. Inorg. Chem. 2017, 56, 5748−5756. (17) Buru, C. T.; Li, P.; Mehdi, B. L.; Dohnalkova, A.; Platero-Prats, A. E.; Browning, N. D.; Chapman, K. W.; Hupp, J. T.; Farha, O. K. Adsorption of a Catalytically Accessible Polyoxometalate in a Mesoporous Channel-type Metal−Organic Framework. Chem. Mater. 2017, 29, 5174−5181. (18) Zhong, X.; Lu, Y.; Luo, F.; Liu, Y.; Li, X.; Liu, S. A Nanocrystalline POM@MOFs Catalyst for the Degradation of Phenol: Effective Cooperative Catalysis by Metal Nodes and POM Guests. Chem. - Eur. J. 2018, 24, 3045−3051. (19) Huang, Y.-B.; Liang, J.; Wang, X.-S.; Cao, R. Multifunctional metal-organic framework catalysts: synergistic catalysis and tandem reactions. Chem. Soc. Rev. 2017, 46, 126−157. (20) Chen, C.-X.; Wei, Z.-W.; Jiang, J.-J.; Zheng, S.-P.; Wang, H.-P.; Qiu, Q.-F.; Cao, C.-C.; Fenske, D.; Su, C.-Y. Dynamic Spacer Installation for Multirole Metal−Organic Frameworks: A New Direction toward Multifunctional MOFs Achieving Ultrahigh Methane Storage Working Capacity. J. Am. Chem. Soc. 2017, 139, 6034−6037. (21) Fan, W.; Wang, Y.; Xiao, Z.; Zhang, L.; Gong, Y.; Dai, F.; Wang, R.; Sun, D. A Stable Amino-Functionalized Interpenetrated Metal−Organic Framework Exhibiting Gas Selectivity and Pore-SizeJ

DOI: 10.1021/acs.inorgchem.8b02272 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (38) Ye, J.-J.; Wu, C.-D. Immobilization of polyoxometalates in crystalline solids for highly efficient heterogeneous catalysis. Dalton Trans 2016, 45, 10101−10112. (39) Lu, B.-B.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. A Polyoxovanadate− Resorcin[4]arene-Based Porous Metal−Organic Framework as an Efficient Multifunctional Catalyst for the Cycloaddition of CO2 with Epoxides and the Selective Oxidation of Sulfides. Inorg. Chem. 2017, 56, 11710−11720. (40) Liu, Y.; Liu, S.; Liu, S.; Liang, D.; Li, S.; Tang, Q.; Wang, X.; Miao, J.; Shi, Z.; Zheng, Z. Facile Synthesis of a Nanocrystalline Metal−Organic Framework Impregnated with a Phosphovanadomolybdate and Its Remarkable Catalytic Performance in Ultradeep Oxidative Desulfurization. ChemCatChem 2013, 5, 3086−3091. (41) Rafiee, E.; Nobakht, N. Keggin type heteropoly acid, encapsulated in metal-organic framework: A heterogeneous and recyclable nanocatalyst for selective oxidation of sulfides and deep desulfurization of model fuels. J. Mol. Catal. A: Chem. 2015, 398, 17− 25. (42) Ribeiro, S.; Granadeiro, C. M.; Silva, P.; Almeida Paz, F. A.; de Biani, F. F.; Cunha-Silva, L.; Balula, S. S. An efficient oxidative desulfurization process using terbium-polyoxometalate@MIL101(Cr). Catal. Sci. Technol. 2013, 3, 2404−2414. (43) Hu, X.; Lu, Y.; Dai, F.; Liu, C.; Liu, Y. Host−guest synthesis and encapsulation of phosphotungstic acid in MIL-101 via “bottle around ship”: An effective catalyst for oxidative desulfurization. Microporous Mesoporous Mater. 2013, 170, 36−44. (44) Lin, Z.-J.; Zheng, H.-Q.; Zheng, H.-Y.; Lin, L.-P.; Xin, Q.; Cao, R. Efficient Capture and Effective Sensing of Cr2O72− from Water Using a Zirconium Metal−Organic Framework. Inorg. Chem. 2017, 56, 14178−14188. (45) Masoomi, M. Y.; Bagheri, M.; Morsali, A. Application of Two Cobalt-Based Metal−Organic Frameworks as Oxidative Desulfurization Catalysts. Inorg. Chem. 2015, 54, 11269−11275. (46) Hao, X.-L.; Ma, Y.-Y.; Zang, H.-Y.; Wang, Y.-H.; Li, Y.-G.; Wang, E.-B. A Polyoxometalate-Encapsulating Cationic Metal− Organic Framework as a Heterogeneous Catalyst for Desulfurization. Chem. - Eur. J. 2015, 21, 3778−3784. (47) Zhang, L.; Wu, S.; Liu, Y.; Wang, F.; Han, X.; Shang, H. Immobilization of phosphotungstic acid in an amino-containing metal−organic framework for oxidative desulfurization. Appl. Organomet. Chem. 2016, 30, 684−690. (48) Furukawa, H.; Gándara, F.; Zhang, Y.-B.; Jiang, J.; Queen, W. L.; Hudson, M. R.; Yaghi, O. M. Water Adsorption in Porous Metal− Organic Frameworks and Related Materials. J. Am. Chem. Soc. 2014, 136, 4369−4381. (49) Liang, W.; Chevreau, H.; Ragon, F.; Southon, P. D.; Peterson, V. K.; D’Alessandro, D. M. Tuning pore size in a zirconiumtricarboxylate metal-organic framework. CrystEngComm 2014, 16, 6530−6533. (50) Ji, P.; Solomon, J. B.; Lin, Z.; Johnson, A.; Jordan, R. F.; Lin, W. Transformation of Metal−Organic Framework Secondary Building Units into Hexanuclear Zr-Alkyl Catalysts for Ethylene Polymerization. J. Am. Chem. Soc. 2017, 139, 11325−11328. (51) Ullah, L.; Zhao, G.; Xu, Z.; He, H.; Usman, M.; Zhang, S. 12Tungstophosphoric acid niched in Zr-based metal-organic framework: a stable and efficient catalyst for Friedel-Crafts acylation. Sci. China: Chem. 2018, 61, 402−411. (52) Mastikhin, V. M.; Terskikh, V. V.; Timofeeva, M. N.; Krivoruchko, O. P. 1H, 31P NMR MAS, infrared and catalytic studies of heteropolyacid H3PW12O40 supported on MgF2. J. Mol. Catal. A: Chem. 1995, 95, 135−140. (53) Ganapathy, S.; Fournier, M.; Paul, J. F.; Delevoye, L.; Guelton, M.; Amoureux, J. P. Location of Protons in Anhydrous Keggin Heteropolyacids H3PMo12O40 and H3PW12O40 by 1H{31P}/31P{1H} REDOR NMR and DFT Quantum Chemical Calculations. J. Am. Chem. Soc. 2002, 124, 7821−7828. (54) Maksimchuk, N. V.; Kovalenko, K. A.; Arzumanov, S. S.; Chesalov, Y. A.; Melgunov, M. S.; Stepanov, A. G.; Fedin, V. P.; Kholdeeva, O. A. Hybrid Polyoxotungstate/MIL-101 Materials:

Synthesis, Characterization, and Catalysis of H2O2-Based Alkene Epoxidation. Inorg. Chem. 2010, 49, 2920−2930. (55) Peng, Y.-L.; Liu, J.; Zhang, H.-F.; Luo, D.; Li, D. A sizematched POM@MOF composite catalyst for highly efficient and recyclable ultra-deep oxidative fuel desulfurization. Inorg. Chem. Front. 2018, 5, 1563−1569. (56) Shi, C.; Wang, W.; Liu, N.; Xu, X.; Wang, D.; Zhang, M.; Sun, P.; Chen, T. Low temperature oxidative desulfurization with hierarchically mesoporous titaniumsilicate Ti-SBA-2 single crystals. Chem. Commun. 2015, 51, 11500−11503. (57) Bhadra, B. N.; Song, J. Y.; Khan, N. A.; Jhung, S. H. TiO2containing carbon derived from a metal−organic framework composite: A highly active catalyst for oxidative desulfurization. ACS Appl. Mater. Interfaces 2017, 9, 31192−31202. (58) Qin, Y.; Xun, S.; Zhan, L.; Lu, Q.; He, M.; Jiang, W.; Li, H.; Zhang, M.; Zhu, W.; Li, H. Synthesis of mesoporous WO3/TiO2 catalyst and its excellent catalytic performance for the oxidation of dibenzothiophene. New J. Chem. 2017, 41, 569−578. (59) Xiong, J.; Zhu, W.; Ding, W.; Yang, L.; Chao, Y.; Li, H.; Zhu, F.; Li, H. Phosphotungstic Acid Immobilized on Ionic LiquidModified SBA-15: Efficient Hydrophobic Heterogeneous Catalyst for Oxidative Desulfurization in Fuel. Ind. Eng. Chem. Res. 2014, 53, 19895−19904. (60) Zheng, H.-Q.; Liu, C.-Y.; Zeng, X.-Y.; Chen, J.; Lü, J.; Lin, R.G.; Cao, R.; Lin, Z.-J.; Su, J.-W. MOF-808: A Metal−Organic Framework with Intrinsic Peroxidase-Like Catalytic Activity at Neutral pH for Colorimetric Biosensing. Inorg. Chem. 2018, 57, 9096−9104. (61) Cai, G.; Jiang, H.-L. A Modulator-Induced Defect-Formation Strategy to Hierarchically Porous Metal−Organic Frameworks with High Stability. Angew. Chem., Int. Ed. 2017, 56, 563−567. (62) Bertleff, B.; Claußnitzer, J.; Korth, W.; Wasserscheid, P.; Jess, A.; Albert, J. Extraction Coupled Oxidative Desulfurization of Fuels to Sulfate and Water-Soluble Sulfur Compounds Using Polyoxometalate Catalysts and Molecular Oxygen. ACS Sustainable Chem. Eng. 2017, 5, 4110−4118. (63) McNamara, N. D.; Neumann, G. T.; Masko, E. T.; Urban, J. A.; Hicks, J. C. Catalytic performance and stability of (V) MIL-47 and (Ti) MIL-125 in the oxidative desulfurization of heterocyclic aromatic sulfur compounds. J. Catal. 2013, 305, 217−226. (64) Granadeiro, C. M.; Ribeiro, S. O.; Karmaoui, M.; Valenca, R.; Ribeiro, J. C.; de Castro, B.; Cunha-Silva, L.; Balula, S. S. Production of ultra-deep sulfur-free diesels using a sustainable catalytic system based on UiO-66(Zr). Chem. Commun. 2015, 51, 13818−13821. (65) Ye, G.; Qi, H.; Li, X.; Leng, K.; Sun, Y.; Xu, W. Enhancement of Oxidative Desulfurization Performance over UiO-66(Zr) by Titanium Ion Exchange. ChemPhysChem 2017, 18, 1903−1908.

K

DOI: 10.1021/acs.inorgchem.8b02272 Inorg. Chem. XXXX, XXX, XXX−XXX