A Stable Mesoporous Zr-Based Metal Organic Framework for Highly

1 day ago - ... mesoporous Zr-based metal–organic framework, [Zr6O4(OH)8(H2O)4(TADIBA)4]·24DMF·45H2O (DMF = N,N-dimethylformamide, H2TADIBA ...
23 downloads 0 Views 4MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

A Stable Mesoporous Zr-Based Metal Organic Framework for Highly Efficient CO2 Conversion Xiaodong Sun, Jiaming Gu, Yang Yuan, Chengyang Yu, Jiantang Li, Hongyan Shan, Guanghua Li, and Yunling Liu* State Key Laboratory of Inorganic Synthesis and Preparative Chemistry, College of Chemistry, Jilin University, Changchun 130012, P. R. China Inorg. Chem. Downloaded from pubs.acs.org by UNIV OF SOUTHERN INDIANA on 05/10/19. For personal use only.

S Supporting Information *

ABSTRACT: With the help of modulator synthesis method, a mesoporous Zr-based metal−organic framework, [Zr6O4(OH)8(H2O)4(TADIBA)4]·24DMF·45H2O (DMF = N,N-dimethylformamide, H2TADIBA = 4,4′-(2H-1,2,4-triazole-3,5-diyl) dibenzoic acid), namely, JLU-MOF58, was successfully constructed. JLU-MOF58 having reo topology is constructed by the bent ligands with Lewis basic sites and 8connected Zr6 clusters with Lewis and Brønsted acid sites. It not only possesses two types of mesoporous cages: octahedral and cuboctahedral (2.76 and 4.10 nm), with a pore volume of 1.76 cm3 g−1, but also displays excellent chemical stability in acid and base aqueous phase. Benefiting from the Brønsted and Lewis acid sites, Lewis basic sites, excellent chemical stability, and the mesoporous cages incorporated in the framework, JLU-MOF58 can be regarded as a remarkably efficient heterogeneous catalyst that exhibits excellent catalytic efficiency for CO2 conversion.



INTRODUCTION Carbon dioxide (CO2), a kind of anthropogenic greenhouse gas, is considered to be the culprit of the global warming, which has been listed as the greatest environmental issue influencing our daily life.1−3 However, CO2 is also an abundant C1 resource, which can be used to produce many valuable chemical feedstocks.4−8 Therefore, developing effective CO2 capture and conversion (CCC) technology is proved to be an alternative sustainable method for eliminating the adverse effect of greenhouse gas.9,10 Cycloaddition of CO2 with different kinds of epoxides to synthesis cyclic carbonates, which are precursors for synthesizing polycarbonates and other chemicals, is considered as an effective method for CO2 conversion.11−13 Herein, many kinds of heterogeneous catalysts have been prepared for cycloaddition of CO2 with epoxides. However, due to high chemical stability of the CO bond of CO2, the reaction can only be triggered under high temperature (>100 °C) and pressure (>3 MPa).14−18 In recent years, metal−organic frameworks (MOFs), because of their high surface areas, adjustable pore sizes, and surface chemistry, were widely studied as a new type of heterogeneous catalyst for CO2 conversion through cycloaddition of CO2 with epoxides.19−36 However, as a kind of heterogeneous catalyst, MOFs materials always suffer from some problems: (1) chemical stability remains a significant obstacle hindering MOFs for practical applications, especially in catalysis;37−40 (2) confinement effect of the channel size always restricts the diffusion of some larger substrates to the © XXXX American Chemical Society

active sites of the microporous MOF, thus resulting in low catalytic efficiency;41−44 (3) high operating pressures (>3 MPa) and temperatures (>100 °C) are still often required for cyclic carbonate production, which will increase the cost considering the energy and capital input.45,46 Herein, design and synthesis of stable mesoporous MOFs, which are able to do cycloaddition of CO2 with epoxides under relatively mild conditions, are urgently demanded and still have great challenges.47−50 In this case, Zr-MOFs, which possess ultrahigh surface area and outstanding thermal and chemical stabilities, make great contributions to overcome this problem.40 As is known, most of the Zr-MOFs that are constructed by ditopic carboxylate ligands are 12-connected.51−57 However, Zr-MOFs with lower connectivity have their unique advantages in CO2 conversion, because the reduced connectivity in Zr-MOFs not only generates additional space for the formation of mesopores, which is convenient for the diffusion of larger substrates, but also provide high densities of Lewis and Brønsted acid sites for CO2 conversion.58−63 On the basis of the above considerations, we employed a bent ligand with Lewis basic sites and highly stable Zr6 clusters for the assembly of Zr-MOF. As expected, an 8-connected mesoporous Zr-MOF [Zr6O4(OH)8(H2O)4(TADIBA)4]· 24DMF·45H 2 O (DMF = N,N-dimethylformamide, H2TADIBA = 4,4′-(2H-1,2,4-triazole-3,5-diyl) dibenzoic Received: March 11, 2019

A

DOI: 10.1021/acs.inorgchem.9b00701 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. (a) From a topology point of view, H2TADIBA ligand can be simplified as linear rod, and 8-connected Zr6 clusters can be simplified as cube; (b) two types of mesoporous cages in the framework of JLU-MOF58: octahedral and cuboctahedral ones with diameters of 2.76 and 4.10 nm between the opposite Zr6 clusters; (c) ball-and-stick view of the JLU-MOF58 along [100] direction; (d) polyhedron view of the framework along [100] direction. Color scheme: carbon = gray; oxygen = red; zirconium = dark green. hot ethanol for 12 h (change fresh ethanol every 2 h), to remove all guest molecules from MOFs. Afterward, supercritical CO2 treatment was performed to activate the samples. Synthesis of JLU-MOF58. H2TADIBA (6 mg, 0.03 mmol), ZrCl4 (12 mg, 0.05 mmol), modulator (acetic acid, 0.3 mL, 0.16 mmol), and DMF (1 mL) were mixed together into a 20 mL vial. The vial was sealed and heated in the 120 °C oven for 2 d. Then the mixture was cooled to the room temperature. Colorless cubic block-shaped crystals were synthesized (70% yield, calculated based on H2TADIBA ligand). Elemental analysis (wt %) data for JLU-MOF58: found: C, 34.75; H, 6.32; N, 11.02, calculated: C, 35.45; H, 6.65; N, 10.95. The purity of the as-synthesized product was verified by the excellent similarity between the experimental and simulated PXRD patterns of JLUMOF58 (Figure S1). Single-Crystal X-ray Crystallography. Bruker D8 venture diffractometer, with graphite-monochromated Mo Kα (λ = 0.710 73 Å), was utilized to collect the crystallographic data of JLU-MOF58 at room temperature. The framework of JLU-MOF58 was solved through direct methods and refined by full-matrix least-squares on F2 using SHELXTL-2014.64 First, all the non-hydrogen atoms were determined in difference Fourier maps and refined anisotropically. Then the hydrogen atoms of the H2TADIBA ligand were generated theoretically and refined isotropically. The SQUEEZE routine in PLATON was used to remove the highly disordered guest molecules, and the results were appended in the CIF file. Finally, the formula of JLU-MOF58 was calculated based on the crystallographic data, elemental data, and thermogravimetric analysis data. Crystal parameters of JLU-MOF58 were presented in Tables S1 and S2. The CCDC number of JLU-MOF58 is 1870852, and the topology information was calculated by TOPOS 4.0 software.65

acid) (JLU-MOF58) with reo topology was successfully constructed. JLU-MOF58 not only contains two types of mesoporous cages: octahedral and cuboctahedral (2.76 and 4.10 nm), with a pore volume of 1.76 cm3 g−1, but also could maintain excellent chemical stability in aqueous phase with wide range of pH values (from 1 to 9). Given the presence of excellent stability, mesopores, Lewis and Brønsted acid sites and Lewis basic sites, JLU-MOF58 exhibits remarkable catalytic efficiency for CO2 conversion under a mild condition.



EXPERIMENTAL SECTION

Materials and Methods. All chemicals, which are reagent grade, were purchased from chemical companies. C, H, and N elemental analysis was tested by utilizing the vario MICRO (Elementar) instrument. Powder X-ray diffraction (PXRD) data were performed on Rigaku D/max-2550 diffractometer (λ = 1.5418 Å) over the 2θ ranging from 2.5 to 40°. PerkinElmer Optima 3300 Dv spectrometer was utilized to analyze the inductively coupled plasma (ICP) data. Thermogravimetric analyses (TGA) data were collected on a TGA Q500 thermogravimetric analyzer (30 to 800 °C, 10 °C min−1). Shimadzu GCMS-QP 2010 plus was used to collect the gas chromatography−mass spectrometry (GC-MS) data (injector temperature 250 °C, TR-wax-ms, 30 m × 0.25 mm × 0.25 μm). The N2 adsorption measurements were tested on the Micromeritics ASAP 2420. The alkane gas and CO2 adsorption data were collect from the Micromeritics 3-Flex. To activate the samples, as-synthesized JLUMOF58 (∼50 mg) was washed by fresh DMF to remove unreacted ligand, metal salt, and acetic acid. Then the samples were washed by B

DOI: 10.1021/acs.inorgchem.9b00701 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 2. (a) PXRD patterns of the JLU-MOF58 samples after being soaked in acid and alkaline aqueous solutions (pH = 1 to 9) for 3 d; (b) N2 adsorption isotherms of JLU-MOF58 after soaking in HCl (pH = 1), NaOH (pH = 9), and H2O aqueous solutions for 24 h; (inset) pore size distribution plot of JLU-MOF58 calculated using nonlocalized density functional theory method. Catalytic Performance of JLU-MOF58. All the catalytic reactions were tested in a 10 mL sealed test tube containing catalysts (0.05, 0.1, and 0.2 mol % calculated based on the activated JLUMOF58 molecular weight), epoxide (10 mmol), and cocatalyst of tetra-n-tert-butylammonium bromide (TBAB, 0.1, 0.3, and 0.5 mol %). Then the reaction tube was purged of air with CO2 four times under 1 atm. After the CO2 was introduced, the reaction mixtures were stirred under a relatively mild condition (25 and 80 °C for propylene oxide and larger epoxides, respectively) for specific times. The catalysts were then recycled by centrifugation and prepared for the next cycle experiment. The final products were detected by GCMS with the peak area normalization method and determined by 1H NMR.

MOF58 viewed along [100], [101], and [111] directions illustrated that it had empty skeleton (Figure S3). The solvent-accessible volume of JLU-MOF58 was calculated to be 169 497 Å3 per unit cell, which counts for ∼75% of the cell volume. Because of the large cavities and high porosity of the framework, JLU-MOF58 must be carefully activated with supercritical CO2. As shown in Figure 2b, typical type IV N2 adsorption isotherm of JLU-MOF58 exhibits an increase at a relative pressure of P/P0 = 0.13, corresponding to a mesoporous cage of 2.5 nm in JLU-MOF58. The Brunauer− Emmett−Teller (BET) and Langmuir surface areas are calculated to be 3663 and 6808 m2 g−1, respectively. The BET surface area surpasses many reported Zr-MOFs (Table S3). The experimental total pore volume is 1.76 cm3 g−1 for JLU-MOF58, which is close to the theoretical value (2.17 cm3 g−1). Remarkably, JLU-MOF58 possesses not only large mesopores but also excellent thermal and chemical stabilities. The high collapse temperature (∼450 °C), which can be confirmed by TGA curves and the temperature-dependent PXRD patterns of JLU-MOF58, reveals its excellent thermal stability (Figure S2). As shown in Figure 2a, the material structure can also be well-maintained after being immersed in acid and base aqueous solutions (pH = 1 to 9). Moreover, N2 adsorption isotherms of the samples, which were treated by water and strong acid solutions, further indicated that JLUMOF58 possessed remarkable chemical stability. The excellent stability of this Zr-MOF is due to the high positive charge density on Zr (IV), which can generate strong interactions between the ligand carboxylates and metal clusters.40 After the permanent porosity was established, the CO2 sorption ability of JLU-MOF58 was investigated first. As illustrated in Figure S6, the CO2 uptake values of JLU-MOF58 are determined to be 49 and 28 cm3 g−1 at 273 and 298 K under 1 bar, respectively. Because of the empty skeleton of the JLU-MOF58 framework, there were no multiple interactions between CO2 molecules and the framework. Herein the CO2 uptake value of JLU-MOF58 was lower. At zero loading, the isosteric heat (Qst) of CO2 for the JLU-MOF58 is 19 kJ mol−1. Because of the high density of Lewis and Brønsted acid sites, Lewis basic sites, excellent chemical stability, and the mesoporous cages in the framework, we intend to evaluate the catalytic performance of JLU-MOF58 for cycloaddition of



RESULTS AND DISCUSSION Single X-ray crystallographic analysis reveals that JLU-MOF58 crystallizes in cubic space group Fm3̅m with the lattice parameter a = 60.912(7) Å. The total framework is constructed by 8-connected Zr6 cluster and bent aromatic ditopic carboxylates (H2TADIBA) with Lewis basic sites. As is known, combining Zr6 clusters and dicarboxylate ligands usually generates classical 12-connected UiO-series MOFs.51−57 However, in JLU-MOF58, eight edges of the Zr6 octahedral cluster are bridged by carboxylates, and the remaining sites are occupied by terminal −OH and H2O molecules. Consequently, the Zr6 clusters become 8-connected nodes with the symmetry reduced from Oh to D4h (Figure 1a). From a topology point of view, H2TADIBA ligand can be simplified as linear rod, and 8-connected Zr6 clusters can be simplified as cube. Herein, the whole framework is 8connected with reo topology (Figure 1c,d). Because of the reduced connectivity, which can potentially provide additional space, two kinds of mesoporous cages, octahedral and cuboctahedral, are formed in JLU-MOF58 (Figure S4). As shown in Figure 1b, the octahedral cage consists of six 8connected Zr6 clusters and 12 bent ligands with 2.76 nm between the opposite Zr6 clusters. The cuboctahedral cage is constructed by six square and eight triangular windows with a 4.10 nm inner pore size. Therefore, the octahedral cages connected with cuboctahedral cages by sharing the edges of the triangular windows to construct a mesoporous framework. Connolly surface area and space-filling models of the JLUC

DOI: 10.1021/acs.inorgchem.9b00701 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry CO2 with different kinds of epoxides. Taking advantage of the mesoporous cages in JLU-MOF58, transforming styrene oxide, a larger substrate, with CO2 into styrene carbonate was selected as a typical reaction. To optimize the reaction conditions, a series of control tests were performed (Table S4). When the reactions were performed in the absence of JLUMOF58 or TBAB, only a small amount of substrate were converted (Table 1, entries 5 and 6). While combining both Table 1. Cycloaddition of CO2 with Styrene Epoxide under Different Conditions

entry

catalysts

time, h

conversion, %

selectivity, %

yield, %

1 2 3 4 5 6 7

JLU-MOF58 JLU-MOF58a JLU-MOF58b JLU-MOF58c no catalyst no TBAB ligand

12 24 24 24 24 24 24

68 99 85 94 28 31 28

95 96 97 94 75 72 58

65 95 82 88 21 22 16

Figure 3. A possible cooperative catalytic mechanism for cycloaddition of styrene oxide and CO2 catalyzed by JLU-MOF58/TBAB. The blue ball represents the Br− of TBAB, and light blue cube represents the Lewis basic sites of the framework. The parts selected by the dotted line (red and blue) represent the Lewis and Brønsted acid sites in the framework. Color code: Zr = green, O = red, H = white, and C = gray.

a

Reaction conditions: styrene oxide (10 mmol), JLU-MOF58 (0.1 mol %), and TBAB (nBu4NBr, 5 mol %) (1 atm CO2, 80 °C). Catalytic performance was tested by 1H NMR and GC-MS. bJLUMOF58 (0.05 mol %). cTBAB (1 mol %).

facilitated by binding the oxygen atom of epoxide with the Lewis acidic zirconium site and forming hydrogen bonds between the (−OH) group of Zr6 cluster and oxygen atom of epoxide. Then, the Br− of TBAB as nucleophiles attacks the polarized ring, forming an opened ring, and exists in the form of intermediate oxygen anion. Subsequently, it rapidly reacts with the contiguous CO2 molecules, which have been activated by Lewis basic sites of the H2TADIBA ligand in the framework, to generate alkycarbonate anions. Finally, the cycloaddition reaction is achieved through the ring-closing step, and TBAB could be recycled simultaneously. The mesoporous cages of JLU-MOF58 could also boost the diffusion of the substrates and products, thus promoting the cycloaddition reaction. In summary, the whole cycloaddition reaction is synergistically promoted by the Lewis, Brønsted acid and basic sites incorporated in the mesoporous framework. Encouraged by the excellent catalytic performance of JLUMOF58 for styrene oxide, another four epoxides with different sizes and ring tension substrates were selected to further study the effect of spatial confinement effect on catalytic efficiency under the optimized conditions. As expected, because of the presence of large mesoporous cages in the frameworks, the catalysis performance of JLU-MOF58 is not influenced by the size or shape of the substrates. As shown in Table 2, JLUMOF58 exhibited outstanding catalytic activity for different sizes and shape substrates, from the smallest substrate: PO (95% yield, 4.2 × 2.6 Å) to the largest substrate: glycidyl 2methylphenyl ether (yield: 81%, 9.8 × 5.2 Å). It is worth mentioning that, due to the pore size restriction in microporous MOFs, only a few works evaluated the catalytic performance for cycloaddition of benzyl phenylglycidyl ether. However, because of the mesopores, JLU-MOF58 is proved to be an efficient catalyst for this reaction with a high yield of 90%, which is comparable to many well-known MOF

JLU-MOF58 and TBAB as a binary catalyst (Table 1, entry 2), styrene oxide was efficiently converted into styrene carbonate (Table 1, entries 2, 3, and 4), which indicates the critical role of JLU-MOF58 and TBAB (cocatalyst) during the cycloaddition reaction. After the factor of time was further optimized, along with the amount of catalyst and cocatalyst (Table 1, entries 1, 2, 3, and 4), JLU-MOF58 efficiently converted CO2 with a high yield of 95% and a remarkable turnover frequency (TOF) value of 40 under the reaction conditions of catalyst (0.1 mol %) and cocatalyst (5 mol %) at a relatively mild condition (1 atm and 80 °C) for 24 h (Table S4). Because of the pore size restriction, only a few MOF materials exhibit high performance for transforming styrene oxide. However, because of the presence of mesopores in JLUMOF58, its catalytic performance is comparable to many reported Zr-MOFs and other metal-based MOFs under the similar conditions, such as Zr-based MOF-892, MOF-893,58 PCN-700,66 Zr(H4L),67 and other metal-based Gea-MOF-1,68 Hf-NU-1000,69 and Cu-TPTC-NH2 20 materials (Table S5).70−72 Compared with Gea-MOF-1, JLU-MOF58 could promote the progress of the reaction at a relatively mild condition. According to previous reports, when MOF materials act as heterogeneous catalyst for cycloaddition of CO2 reaction, the Lewis/Brønsted acid sites in the framework could fix and polarize the epoxide. If there exist Lewis basic sites in the framework, it would polarize the carbon dioxide and promote the reaction.73−77 Herein, a possible cooperative catalytic mechanism for the cycloaddition of CO2 with epoxide into cyclic carbonate is proposed. As illustrated in Figure 3, the overall reaction consists of four steps. First, the reaction is initiated by polarizing the epoxide ring, which is synergistically D

DOI: 10.1021/acs.inorgchem.9b00701 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Table 2. Cycloaddition of CO2 and Different-Size Epoxides Catalyzed by JLU-MOF58 under the Optimum Conditionsa

a

Reaction conditions: different-size epoxide (10 mmol), catalyst (JLU-MOF58, 0.1 mol %), and cocatalyst (TBAB, 5 mol %) under 1 atm PCO2 for 24 h (25 and 80 °C for propylene oxide and larger epoxides, respectively). Catalytic performances were tested by 1H NMR.



materials, such as (I−)Meim-UiO-66,71 FJI-C10,10 and CuTPTC-NH220 (Table S6). The regenerative performance is very important for a catalyst; when the cost of material synthesis is considered, the recycling experiments for the CO2 cycloaddition of styrene oxide were conducted. As shown in Figure S15, after six rounds of circulation, JLU-MOF58 could be successfully recycled without any significant loss in its catalytic activity, which indicates the excellent practicability of the JLU-MOF58. Meanwhile, the ICP-optical emission spectrometry (OES) analysis and the PXRD patterns of the six times recycled samples further confirmed that JLU-MOF58 was an efficient and reusable heterogeneous MOF catalyst (Figure S16).



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.9b00701. Gas sorption data, TGA, PXRD, 1H NMR data, additional structural figures, and crystallographic data (PDF) Accession Codes

CCDC 1870852 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.

CONCLUSIONS

In conclusion, by using the modulator synthesis approach, we design and construct a Zr-based MOF containing two types of mesoporous cages (2.76 and 4.10 nm) with reo topology. Although MOFs with high surface areas tend to decompose, JLU-MOF58 exhibits excellent chemical stability in acid and alkaline aqueous. Given the presence of chemical stability, Lewis and Brønsted acid sites, and basic sites incorporated in the mesoporous framework, JLU-MOF58 exhibits efficient catalytic performance and outstanding recyclability for CO2 cycloaddition with epoxides under a relatively mild condition. Herein, design and synthesis of such Zr-MOFs owing mesoporous cages are highly desirable for CO2 fixation in the field of global climate treatment.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Jiantang Li: 0000-0002-8963-5402 Guanghua Li: 0000-0003-3029-8920 Yunling Liu: 0000-0001-5040-6816 Notes

The authors declare no competing financial interest. E

DOI: 10.1021/acs.inorgchem.9b00701 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry



(18) Yano, T.; Matsui, H.; Koike, T.; Ishiguro, H.; Fujihara, H.; Yoshihara, M.; Maeshima, T. Magnesium oxide-catalysed reaction of carbon dioxide with an epoxide with retention of stereochemistry. Chem. Commun. 1997, 1129−1130. (19) Dhakshinamoorthy, A.; Li, Z.; Garcia, H. Catalysis and photocatalysis by metal organic frameworks. Chem. Soc. Rev. 2018, 47, 8134−8172. (20) Sharma, V.; De, D.; Saha, R.; Das, R.; Chattaraj, P. K.; Bharadwaj, P. K. A Cu(II)-MOF capable of fixing CO2 from air and showing high capacity H2 and CO2 adsorption. Chem. Commun. 2017, 53, 13371−13374. (21) Liang, L.; Liu, C.; Jiang, F.; Chen, Q.; Zhang, L.; Xue, H.; Jiang, H.-L.; Qian, J.; Yuan, D.; Hong, M. Carbon dioxide capture and conversion by an acid-base resistant metal-organic framework. Nat. Commun. 2017, 8, 1233. (22) Wei, N.; Zuo, R.-X.; Zhang, Y.-Y.; Han, Z.-B.; Gu, X.-J. Robust high-connected rare-earth MOFs as efficient heterogeneous catalysts for CO2 conversion. Chem. Commun. 2017, 53, 3224−3227. (23) Li, B.; Chrzanowski, M.; Zhang, Y.; Ma, S. Applications of metal-organic frameworks featuring multi-functional sites. Coord. Chem. Rev. 2016, 307, 106−129. (24) Chen, J.; Zhong, M.; Tao, L.; Liu, L.; Jayakumar, S.; Li, C.; Li, H.; Yang, Q. The cooperation of porphyrin-based porous polymer and thermal-responsive ionic liquid for efficient CO2 cycloaddition reaction. Green Chem. 2018, 20, 903−911. (25) Xu, H.; Liu, X.-F.; Cao, C.-S.; Zhao, B.; Cheng, P.; He, L.-N. A porous Metal-Organic Framework assembled by [Cu30] nanocages: serving as recyclable catalysts for CO2 fixation with aziridines. Adv. Sci. 2016, 3, 1600048. (26) He, H.; Perman, J. A.; Zhu, G.; Ma, S. Metal-Organic Frameworks for CO2 chemical transformations. Small 2016, 12, 6309−6324. (27) Ding, M.; Jiang, H.-L. Incorporation of imidazolium-based poly(ionic liquid)s into a Metal-Organic Framework for CO2 capture and conversion. ACS Catal. 2018, 8, 3194−3201. (28) Aguila, B.; Sun, Q.; Wang, X.; O’Rourke, E.; Al-Enizi, A. M.; Nafady, A.; Ma, S. Lower activation energy for catalytic reactions through Host-Guest cooperation within Metal-Organic Frameworks. Angew. Chem. 2018, 130, 10264−10268. (29) Wang, X.; Gao, W.-Y.; Niu, Z.; Wojtas, L.; Perman, J. A.; Chen, Y.-S.; Li, Z.; Aguila, B.; Ma, S. A metal-metalloporphyrin framework based on an octatopic porphyrin ligand for chemical fixation of CO2 with aziridines. Chem. Commun. 2018, 54, 1170−1173. (30) Chen, Y. Z.; Zhang, R.; Jiao, L.; Jiang, H. L. Metal-organic framework-derived porous materials for catalysis. Coord. Chem. Rev. 2018, 362, 1−23. (31) Niu, Z.; BhagyaGunatilleke, W. D. C.; Sun, Q.; Lan, P. C.; Perman, J.; Ma, J. G.; Cheng, Y.; Aguila, B.; Ma, S. Metal-Organic Framework anchored with a lewispair as a new paradigm for catalysis. Chem. 2018, 4, 2587−2599. (32) Li, P.-Z.; Wang, X. J.; Liu, J.; Phang, H. S.; Li, Y.; Zhao, Y. Highly effective carbon fixation via catalytic conversion of CO2 by an acylamide-containing Metal-Organic Framework. Chem. Mater. 2017, 29, 9256−9261. (33) Ding, M.; Chen, S.; Liu, X.-Q.; Sun, L.-B.; Lu, J.; Jiang, H. L. Metal-Organic Framework-templated catalyst: synergy in multiple sites for catalytic CO2 fixation. ChemSusChem 2017, 10, 1898−1903. (34) Dong, J.; Cui, P.; Shi, P. F.; Cheng, P.; Zhao, B. Ultrastrongalkali-resisting lanthanide-zeolites assembled by [Ln60] nanocages. J. Am. Chem. Soc. 2015, 137, 15988−15991. (35) Maina, J. W.; Pozo-Gonzalo, C.; Kong, L.; Schütz, J.; Hill, M.; Dumée, L. F. Metal organic framework based catalysts for CO2 conversion. Mater. Horiz. 2017, 4, 345−361. (36) Feng, D.; Chung, W.-C.; Wei, Z.; Gu, Z.-Y.; Jiang, H.-L.; Chen, Y.-P.; Darensbourg, D. J.; Zhou, H.-C. Construction ofUltrastable Porphyrin Zr Metal-Organic Frameworks through Linker Elimination. J. Am. Chem. Soc. 2013, 135, 17105−17110.

ACKNOWLEDGMENTS The authors gratefully acknowledge the financial support of the National Natural Science Foundation of China (Nos. 21771078, 21671074, and 21621001), the 111 Project (B17020), and the National Key Research and Development Program of China (2016YFB0701100).



REFERENCES

(1) Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm, Z. R.; Bae, T.-H.; Long, J. R. Carbon Dioxide Capture in Metal-Organic Frameworks. Chem. Rev. 2012, 112, 724− 781. (2) Li, J.-R.; Ma, Y.; McCarthy, M. C.; Sculley, J.; Yu, J.; Jeong, H.K.; Balbuena, P. B.; Zhou, H.-C. Carbon dioxide capture-related gas adsorption and separation in metal-organic frameworks. Coord. Chem. Rev. 2011, 255, 1791−1823. (3) Liu, J.; Thallapally, P. K.; McGrail, B. P.; Brown, D. R.; Liu, J. Progress in adsorption-based CO2 capture by metal-organic frameworks. Chem. Soc. Rev. 2012, 41, 2308−2322. (4) Himeda, Y.; Onozawa-Komatsuzaki, N.; Sugihara, H.; Kasuga, K. Recyclable catalyst for conversion of carbon dioxide into formateattributable to an oxyanion on the catalyst ligand. J. Am. Chem. Soc. 2005, 127, 13118−13119. (5) Darensbourg, D. J.; Holtcamp, M. W. Catalysts for the reactions of epoxides and carbon dioxide. Coord. Chem. Rev. 1996, 153, 155− 174. (6) Wang, W.; Wang, S.; Ma, X.; Gong, J. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 3703−3727. (7) Jiang, T.; Ma, X.; Zhou, Y.; Liang, S.; Zhang, J.; Han, B. Solventfree synthesis of substituted ureas from CO2 and amines with a functional ionic liquid as the catalyst. Green Chem. 2008, 10, 465− 469. (8) Fu, Y.; Sun, D.; Chen, Y.; Huang, R.; Ding, Z.; Fu, X.; Li, Z. An amine-functionalized titanium Metal-Organic Framework photocatalyst with visible-light-induced activity for CO2 reduction. Angew. Chem., Int. Ed. 2012, 51, 3364−3367. (9) Trickett, C. A.; Helal, A.; Al-Maythalony, B. A.; Yamani, Z. H.; Cordova, K. E.; Yaghi, O. M. The chemistry of metal−organic frameworks for CO2 capture, regeneration and conversion. Nat. Rev. Mater. 2017, 2, 17045. (10) Liang, J.; Xie, Y.-Q.; Wang, X.-S.; Wang, Q.; Liu, T.-T.; Huang, Y.-B.; Cao, R. An imidazolium-functionalized mesoporous cationic metal−organic framework for cooperative CO2 fixation into cyclic carbonate. Chem. Commun. 2018, 54, 342−345. (11) Lu, X.-B.; Darensbourg, D. J. Cobalt catalysts for the coupling of CO2 and epoxides to provide polycarbonates and cyclic carbonates. Chem. Soc. Rev. 2012, 41, 1462−1484. (12) North, M.; Pasquale, R.; Young, C. Synthesis of cyclic carbonates from epoxides and CO2. Green Chem. 2010, 12, 1514− 1539. (13) Ma, J.; Sun, N.; Zhang, X.; Zhao, N.; Xiao, F.; Wei, W.; Sun, Y. A short review of catalysis for CO2 conversion. Catal. Today 2009, 148, 221−231. (14) Xie, Y.; Wang, T.-T.; Liu, X.-H.; Zou, K.; Deng, W.-Q. Capture and conversion of CO2 at ambient conditions by a conjugated microporous polymer. Nat. Commun. 2013, 4, 1960. (15) Tu, W.; Zhou, Y.; Zou, Z. Photocatalytic Conversion of CO2 into Renewable Hydrocarbon Fuels: State-of-the-Art Accomplishment, Challenges, and Prospects. Adv. Mater. 2014, 26, 4607−4626. (16) Lin, S.; Diercks, C. S.; Zhang, Y.-B.; Kornienko, N.; Nichols, E. M.; Zhao, Y.; Paris, A. R.; Kim, D.; Yang, P.; Yaghi, O. M.; Chang, C. J. Covalent organic frameworks comprising cobalt porphyrins for catalytic CO2 reduction in water. Science 2015, 349, 1208. (17) Yamaguchi, K.; Ebitani, K.; Yoshida, T.; Yoshida, H.; Kaneda, K. Mg−Al mixed oxides as highly active acid-base catalysts for cycloaddition of carbon dioxide to epoxides. J. Am. Chem. Soc. 1999, 121, 4526−4527. F

DOI: 10.1021/acs.inorgchem.9b00701 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (37) Canivet, J.; Fateeva, A.; Guo, Y.; Coasne, B.; Farrusseng, D. Water adsorption in MOFs: fundamentals and applications. Chem. Soc. Rev. 2014, 43, 5594−5617. (38) Gascon, J.; Kapteijn, F. Metal-Organic Framework membraneshigh potential, bright future. Angew. Chem., Int. Ed. 2010, 49, 1530− 1532. (39) Keskin, S.; Van Heest, T. M.; Sholl, D. S. Can Metal-Organic Framework materials play a useful role in large-scale carbon dioxide separations. ChemSusChem 2010, 3, 879−891. (40) Bai, Y.; Dou, Y.; Xie, L.-H.; Rutledge, W.; Li, J.-R.; Zhou, H.-C. Zr-based metal−organic frameworks: design, synthesis, structure, and applications. Chem. Soc. Rev. 2016, 45, 2327−2367. (41) Li, P. Z.; Wang, X. J.; Liu, J.; Lim, J. S.; Zou, R.; Zhao, Y. A triazole-containing Metal-Organic Framework as a highly effective and substrate size-dependent catalyst for CO2 conversion. J. Am. Chem. Soc. 2016, 138, 2142−2145. (42) Gu, J.-M.; Kim, W.-S.; Huh, S. Size-dependent catalysis by DABCO-functionalized Zn-MOF with one-dimensional channels. Dalton Trans. 2011, 40, 10826−10829. (43) Ugale, B.; Dhankhar, S. S.; Nagaraja, C. M. Construction of 3fold-interpenetrated three-dimensional Metal-Organic Frameworks of Nickel(II) for highly efficient capture and conversion of carbon dioxide. Inorg. Chem. 2016, 55, 9757−9766. (44) Fan, Y.; Li, J.; Ren, Y.; Jiang, H. A Ni(salen)-based MetalOrganic Framework: synthesis, structure, and catalytic performance for CO2 cycloaddition with epoxides. Eur. J. Inorg. Chem. 2017, 2017, 4982−4989. (45) Zhao, D.; Liu, X.-H.; Guo, J.-H.; Xu, H.-J.; Zhao, Y.; Lu, Y.; Sun, W.-Y. Porous Metal-Organic Frameworks with chelating multiaminesites for selective adsorption and chemical conversion of carbon dioxide. Inorg. Chem. 2018, 57, 2695−2704. (46) Kim, J.; Kim, S.-N.; Jang, H.-G.; Seo, G.; Ahn, W.-S. CO2 cycloaddition of styrene oxide over MOF catalysts. Appl. Catal., A 2013, 453, 175−180. (47) Mo, Z.-W.; Zhou, H.-L.; Zhou, D.-D.; Lin, R.-B.; Liao, P.-Q.; He, C.-T.; Zhang, W.-X.; Chen, X.-M.; Zhang, J.-P. Mesoporous Metal-Organic Frameworks with exceptionally high working capacities for adsorption heat transformation. Adv. Mater. 2018, 30, 1704350. (48) Klein, N.; Senkovska, I.; Gedrich, K.; Stoeck, U.; Henschel, A.; Mueller, U.; Kaskel, S. A mesoporous Metal-Organic Framework. Angew. Chem., Int. Ed. 2009, 48, 9954−9957. (49) Feng, D.; Gu, Z.-Y.; Li, J.-R.; Jiang, H.-L.; Wei, Z.; Zhou, H.-C. Zirconium-metalloporphyrin PCN-222: mesoporous Metal-Organic Frameworks with ultrahigh stability as biomimetic catalysts. Angew. Chem., Int. Ed. 2012, 51, 10307−10310. (50) Feng, D.; Wang, K.; Su, J.; Liu, T.-F.; Park, J.; Wei, Z.; Bosch, M.; Yakovenko, A.; Zou, X.; Zhou, H.-C. A highly stable zeotypemesoporous zirconium Metal-Organic Framework with ultralargepores. Angew. Chem., Int. Ed. 2015, 54, 149−154. (51) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, K. P. A new zirconium inorganic building brick forming Metal Organic Frameworks with exceptional stability. J. Am. Chem. Soc. 2008, 130, 13850−13851. (52) Furukawa, H.; Gándara, F.; Zhang, Y.-B.; Jiang, J.; Queen, W. L.; Hudson, M. R.; Yaghi, O. M. Water adsorption in porous MetalOrganic Frameworks and related materials. J. Am. Chem. Soc. 2014, 136, 4369−4381. (53) Jiang, H.-L.; Feng, D.; Liu, T.-F.; Li, J.-R.; Zhou, H.-C. Pore surface engineering with controlled loadings of functional groups via click chemistry in highly stable Metal-Organic Frameworks. J. Am. Chem. Soc. 2012, 134, 14690−14693. (54) Schaate, A.; Roy, P.; Preuße, T.; Lohmeier, S. J.; Godt, A.; Behrens, P. Porous Interpenetrated Zirconium-Organic Frameworks (PIZOFs): A chemically versatile family of Metal-Organic Frameworks. Chem. - Eur. J. 2011, 17, 9320−9325. (55) Garibay, S. J.; Cohen, S. M. Isoreticular synthesis and modification of frameworks with the UiO-66 topology. Chem. Commun. 2010, 46, 7700−7702.

(56) Choi, K. M.; Jeong, H. M.; Park, J. H.; Zhang, Y.-B.; Kang, J. K.; Yaghi, O. M. Supercapacitors of nanocrystalline Metal-Organic Frameworks. ACS Nano 2014, 8, 7451−7457. (57) Wang, B.; Huang, H.; Lv, X.-L.; Xie, Y.; Li, M.; Li, J.-R. Tuning CO2 selective adsorption over N2 and CH4 in UiO-67 analogues through ligand functionalization. Inorg. Chem. 2014, 53, 9254−9259. (58) Nguyen, P. T. K.; Nguyen, H. T. D.; Nguyen, H. N.; Trickett, C. A.; Ton, Q. T.; Gutiérrez-Puebla, E.; Monge, M. A.; Cordova, K. E.; Gándara, F. New Metal-Organic Frameworks for chemical fixation of CO2. ACS Appl. Mater. Interfaces 2018, 10, 733−744. (59) Qin, J.-S.; Yuan, S.; Lollar, C.; Pang, J.; Alsalme, A.; Zhou, H.C. Stable metal-organic frameworks as a host platform for catalysis and biomimetics. Chem. Commun. 2018, 54, 4231−4249. (60) Mondloch, J. E.; Katz, M. J.; Isley III, W. C.; Ghosh, P.; Liao, P.; Bury, W.; Wagner, G. W.; Hall, M. G.; DeCoste, J. B.; Peterson, G. W.; Snurr, R. Q.; Cramer, C. J.; Hupp, J. T.; Farha, O. K. Destruction of chemical warfare agents using metal-organic frameworks. Nat. Mater. 2015, 14, 512. (61) Huang, N.; Yuan, S.; Drake, H.; Yang, X.; Pang, J.; Qin, J.; Li, J.; Zhang, Y.; Wang, Q.; Jiang, D.; Zhou, H.-C. Systematic engineering of single substitution in Zirconium Metal-Organic Frameworks toward high-performance catalysis. J. Am. Chem. Soc. 2017, 139, 18590− 18597. (62) Noh, H.; Cui, Y.; Peters, A. W.; Pahls, D. R.; Ortuño, M. A.; Vermeulen, N. A.; Cramer, C. J.; Gagliardi, L.; Hupp, J. T.; Farha, O. K. An exceptionally stable Metal-Organic Framework supported molybdenum(VI) oxide catalyst for cyclohexene epoxidation. J. Am. Chem. Soc. 2016, 138, 14720−14726. (63) Zhu, J.; Usov, P. M.; Xu, W.; Celis-Salazar, P. J.; Lin, S.; Kessinger, M. C.; Landaverde-Alvarado, C.; Cai, M.; May, A. M.; Slebodnick, C.; Zhu, D.; Senanayake, S. D.; Morris, A. J. A New Class of metal-cyclam-based Zirconium Metal-Organic Frameworks for CO2 adsorption and chemical fixation. J. Am. Chem. Soc. 2018, 140, 993−1003. (64) Sheldrick, G. M. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3−8. (65) Blatov, V. A.; Shevchenko, A. P.; Proserpio, D. M. Applied Topological Analysis of Crystal Structures with the Program Package ToposPro. Cryst. Growth Des. 2014, 14, 3576−3586. (66) Yuan, S.; Zou, L.; Li, H.; Chen, Y.-P.; Qin, J.; Zhang, Q.; Lu, W.; Hall, M. B.; Zhou, H.-C. Flexible Zirconium Metal-Organic Frameworks as bioinspired switchable catalysts. Angew. Chem., Int. Ed. 2016, 55, 10776−10780. (67) Gao, C. Y.; Ai, J.; Tian, H. R.; Wu, D.; Sun, Z.-M. An ultrastable zirconium-phosphonate framework as bifunctional catalyst for highly active CO2 chemical transformation. Chem. Commun. 2017, 53, 1293−1296. (68) Guillerm, V.; Weseliński, Ł. J.; Belmabkhout, Y.; Cairns, A. J.; D’Elia, V.; Wojtas, Ł.; Adil, K.; Eddaoudi, M. Discovery and introduction of a (3,18)-connected net as an ideal blueprint for the design of metal−organic frameworks. Nat. Chem. 2014, 6, 673. (69) Beyzavi, M. H.; Klet, R. C.; Tussupbayev, S.; Borycz, J.; Vermeulen, N. A.; Cramer, C. J.; Stoddart, J. F.; Hupp, J. T.; Farha, O. K. A Hafnium-Based Metal-Organic Framework as an Efficient and Multifunctional Catalyst for Facile CO2 Fixation and Regioselective and Enantioretentive Epoxide Activation. J. Am. Chem. Soc. 2014, 136, 15861−15864. (70) Li, J.; Ren, Y.; Yue, C.; Fan, Y.; Qi, C.; Jiang, H. Highly Stable Chiral Zirconium−Metallosalen Frameworks for CO2 Conversion and Asymmetric C-H Azidation. ACS Appl. Mater. Interfaces 2018, 10, 36047−36057. (71) Liang, J.; Chen, R.-P.; Wang, X.-Y.; Liu, T.-T.; Wang, X.-S.; Huang, Y.-B.; Cao, R. Postsynthetic ionization of an imidazolecontaining metal-organic framework for the cycloaddition of carbon dioxide and epoxides. Chem. Sci. 2017, 8, 1570−1575. (72) Zhou, Z.; He, C.; Xiu, J.; Yang, L.; Duan, C. Metal-Organic Polymers Containing Discrete Single-Walled Nanotube as a Heterogeneous Catalyst for the Cycloaddition of Carbon Dioxide to Epoxides. J. Am. Chem. Soc. 2015, 137, 15066−15069. G

DOI: 10.1021/acs.inorgchem.9b00701 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (73) Kim, J.; Kim, S.-N.; Jang, H.-G.; Seo, G.; Ahn, W.-S. CO2Cycloaddition of Styrene Oxide over MOF Catalysts. Appl. Catal., A 2013, 453, 175−180. (74) North, M.; Pasquale, R. Mechanism of Cyclic Carbonate Synthesis from Epoxides and CO2. Angew. Chem. 2009, 121, 2990− 2992. (75) Liu, L.; Wang, S.-M.; Han, Z.-B.; Ding, M.; Yuan, D.-Q.; Jiang, H.-L. Exceptionally Robust In-Based Metal-Organic Framework for Highly Efficient Carbon Dioxide Capture and Conversion. Inorg. Chem. 2016, 55, 3558−3565. (76) Wei, N.; Zhang, Y.; Liu, L.; Han, Z.-B.; Yuan, D.-Q. PentanuclearYb(III) cluster-based metal-organic frameworks as heterogeneous catalysts for CO2 conversion. Appl. Catal., B 2017, 219, 603−610. (77) He, H.; Sun, Q.; Gao, W.; Perman, J. A.; Sun, F.; Zhu, G.; Aguila, B.; Forrest, K.; Space, B.; Ma, S. A Stable Metal-Organic Framework Featuring a Local Buffer Environment for Carbon Dioxide Fixation. Angew. Chem., Int. Ed. 2018, 57, 4657−4662.

H

DOI: 10.1021/acs.inorgchem.9b00701 Inorg. Chem. XXXX, XXX, XXX−XXX