CHEMSUSCHEM FULL PAPERS DOI: 10.1002/cssc.201402206

Porous Anionic Indium–Organic Framework with Enhanced Gas and Vapor Adsorption and Separation Ability Yuanbiao Huang, Zujin Lin, Hongru Fu, Fei Wang, Min Shen, Xusheng Wang, and Rong Cao*[a] A three-dimensional microporous anionic metal–organic framework (MOF) (Et4N)3[In3(TATB)4] (FJI-C1, H3TATB = 4,4’,4’’-s-triazine-2,4,6-triyltribenzoic acid) with large unit cell volume has been synthesized. Assisted by the organic cation group Et4N in the pores of the compound, FJI-C1 not only shows high adsorption uptakes of C2 and C3 hydrocarbons, but also exhibits highly selective separation of propane, acetylene, ethane, and ethylene from methane at room temperature. Furthermore, it also exhibits high separation selectivity for propane over C2 hy-

drocarbons and acetylene can be readily separated from their C2 hydrocarbons mixtures at low pressure due to the high selectivity for C2H2 in comparison to C2H4 and C2H6. In addition, FJI-C1 with hydrophilic internal pores surfaces shows highly efficient adsorption separation of polar molecules from nonpolar molecules. Notably, it exhibits high separation selectivity for benzene over cyclohexane due to the p–p interactions between benzene molecules and s-triazine rings of the porous MOF.

Introduction Metal–organic frameworks (MOFs) have been receiving intensive research interest because of their tunable nanospace and potential applications in gas storage, separation, and catalysis.[1, 2] Recently, anionic MOFs have began to attract attention because extraframework cations play important roles in gas sorption and catalysis.[3–7] Especially, the cationic species in the anionic frameworks enhance the adsorbent–adsorbate interactions through charge-induced forces.[3g, 4] However, only a small number of studies focused on simple inorganic cations, little research using organic cations in MOFs and its effect on the light hydrocarbons and vapors adsorption separation.[4] Compared to other metals, indium(III) as 6–8 coordinated metal node usually forms anionic MOFs.[5, 6] Eddaoudi et al. constructed a series of porous zeolite-like anionic MOFs topologically related to inorganic zeolites for gas adsorption and catalysis.[6, 7] The Bu group reported a family of porous indium–organic frameworks with a cage-within-cage structure and mixed inorganic building blocks with differing nuclearity.[3] Schrçder and co-workers found that partially interpenetrated MOF NOTT-202 (NOTT stands for Nottingham) based on indium has a unique capability for selective hysteretic sorption of CO2 and high adsorption of hydrogen through cationic exchanges.[8, 9] Our group developed a guest-dependent approach to obtain permanent pores in flexible indium(III) MOFs prepared by performing cation exchange for gas adsorption.[10] [a] Dr. Y. Huang, Z. Lin, H. Fu, Dr. F. Wang, M. Shen, X. Wang, Prof. Dr. R. Cao State Key Laboratory of Structural Chemistry Fujian Institute of Research on the Structure of Matter Chinese Academy of Sciences 155, Yangqiao Road West, Fuzhou, 350002 (PR China) E-mail: [email protected] Supporting Information for this article is available on the WWW under http://dx.doi.org/10.1002/cssc.201402206.

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

A trigonal planar ligand, benzenetricarboxylic acid (H3BTC), has been used most frequently to promote the formation of porous MOFs with dinuclear and polynuclear secondary building units (SBUs).[11] The volumetric expansion of BTC3 to benzenetribenzoic acid (H3BTB)[5c, 12] and 4,4’,4’’-s-triazine-2,4,6-triyltribenzoic acid (H3TATB)[13] generally results in MOFs with new structures with larger pore size and high surface area. Especially, compared to BTB3 , TATB3 is more likely to yield highly thermally stable MOFs with polynuclear SBUs and has superior selective gas separation, which is mainly attributed to the planarity of the ligand and its tendency to induce p–p stacking in MOFs due to the increased quadrupole moment.[13] With these considerations in mind, we used H3TATB and indium nitrate to prepare a porous anionic MOF, (Et4N)3[In3(TATB)4] (Et4N = tetraethylammonium) (FJI-C1). Herein, we report the synthesis, structure, gas and vapor adsorption, and separation properties of FJI-C1.

Results and Discussion Synthesis and Crystal structure of FJI-C1 FJI-C1 was successfully synthesized through the solvothermal reaction of H3TATB with indium(III) nitrate hydrate and tetraethylammonium chloride (Et4NCl) in N,N-diethylformamide (DEF) at 120 8C for 3 days. The anionic MOF FJI-C1 (Figures S1 and S2 in the Supporting Information) was isolated as colorless crystals with a high yield of 86 %. Its structure was determined to be (Et4N)3[In3(TATB)4]·(DEF)16·(H2O)11 on the basis of single-crystal X-ray diffraction (XRD) and 1H NMR, elemental, and thermogravimetric analysis (TGA). Single-crystal XRD studies show that FJI-C1 crystallizes in the trigonal space group R-3 with a large unit cell volume ChemSusChem 2014, 7, 2647 – 2653

2647

CHEMSUSCHEM FULL PAPERS

www.chemsuschem.org

(67 829.4(9) 3, Table S1). Each eight-coordinate indium(III) center is bridged by eight oxygen atoms from four different carboxylate groups of TATB3 serving as a tetrahedral four-connected node, while the ligand links of three separate indium(III) centers and thus acts as a three-connected node (Figure 1 a). Consequently, the planar 3-connected TATB3 ligand with tetrahedral 4-connected mononuclear In(COO)4 nodes generates a porous anionic framework with the (3,3,3,3,4,4)connected new topology (Figure 1 c). The free volume accessible to solvents and charge-balancing cations in FJI-C1 is 48 890.4 3 per unit cell, and the pore volume ratio is 72.0 % as calculated with the PLATON program (Figure 1 b).[14] The disor-

Hydrogen, methane, and carbon dioxide adsorption and separation

The porosity of the sample was confirmed by performing a nitrogen sorption experiment at 77 K. The activated sample exhibits a reversible typical type I behavior as expected for microporous materials (Figure 2), which is corresponds with the crystal structure. The Brunauer–Emmett–Teller (BET) and Langmuir surface areas of FJI-C1 are 1726.3 and 2392.6 m2 g 1, respectively, which are larger than those of In MOFs based on BTC3 (507.8 and 711.8 m2 g 1, respectively).[3] The pore distribution analysis by density functional theory (DFT) indicates that there is a narrow distribution of micropores at around 1.1 nm, which agrees well with a single-crystal structure. The high surface area of FJIC1 prompts us to study its gasuptake capacity, especially for hydrogen, methane, and CO2. The low pressure hydrogen sorption isotherms of the samples at 77 and 87 K are shown in Figure S17 a. The amount of H2 uptake at 77 and 87 K is 171.3 3 Figure 1. (a) Coordination mode of TATB in FJI-C1; (b) space-filling 3D model of frameworks in FJI-C1 showing (1.49 wt %) and 104.9 cm3 g 1 hexagonal channels; (c) (3,3,3,3,4,4)-connected new topology (hydrogen atoms, Et4N, DEF, and H2O have been omitted for clarity). (0.94 wt %), respectively, comparably to those of MOF-5 and MOF-74.[15] The isosteric enthalpy + dered positively charged organo cationic Et4N not only serves of adsorption (Qst) of compound FJI-C1 for hydrogen has been as extraframework charge-balancing species incorporated into calculated using virial methods according to adsorption FJI-C1, but also supports and enhances the stability and perdata.[16] The initial hydrogen adsorption enthalpy is [3g, 10] manent porosity of the main framework. 5.44 kJ mol 1. With an increase in hydrogen coverage, Qst deThe powder X-ray diffraction (PXRD) patterns of the resulting crystals are coincident with the simulated ones (Figures S3 and S4), which indicates a good purity and homogeneity of FJIC1. Although some of the Et4N + and the solvents (DEF and water) molecules within the channels are crystallographically disordered due to the large unit cell volume, the stoichiometry has been determined through integration analysis of 1H NMR spectroscopy (Figures S6 and S7), mass spectrum (Figure S8), TGA (Figure S9), IR spectroscopy (Figure S10), and elemental analysis.

Figure 2. N2 sorption isotherms at 77 K for FJI-C1. Inset: Pore size distribution calculated by DFT.

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2014, 7, 2647 – 2653

2648

CHEMSUSCHEM FULL PAPERS creases steadily (Figure S21). On increasing the pressure, the samples reveal an increasing hydrogen uptake with a maximum excess value reaching 3.31 wt % (29.7 g L 1) at 77 K and 32 bar (Figures S17 b and S18), which is higher than that of In MOF withh a flexible octacarboxylate ligand (3.05 wt %)[10b] and comparable to that of PCN-66 (PCN stands for porous coordination network, 29.6 g L 1, 45 bar)[17] under the same measurement conditions. Methane, as the main component of natural gas, is another interesting alternative for on-board fuel. To achieve an effective storage of methane, the adsorbents should have a high porosity. Therefore, methane-uptake capacities of the anionic MOF FJI-C1 have also been tested. As is seen from Figure S17 c, the adsorption capacity (1 bar) for methane is 15.1 and 9.7 cm3 g 1 at 273 and 298 K, respectively. Qst of FJI-C1 for methane is 11.4 kJ mol 1 at zero coverage (Figure S23). In the mediumpressure range (< 40 bar, Figures S17 d and S19), the capacity increases very fast. As the pressure is further increased, the uptake increases slowly and the maximum excess value reaches 12.0 wt % (150.2 v/v) at 298 K and 90 bar, which is higher than that of PCN-61 (145.0 v/v, 35 bar 298 k)[17] . The capture and storage of CO2 is becoming urgent because CO2 is one of the major greenhouse gases causing global warming.[9] The CO2 adsorption isotherms of FJI-C1 have been measured at 195, 273, and 298 K (Figure S20 a). The adsorption capacity for CO2 is 64.0 and 41.2 cm3 g 1 at 273 and 298 K, respectively. Interestingly, when the temperature decreases to 195 K, the CO2 uptake capacity is as high as 427.5 cm3 g 1 (Figure S20 a). The high CO2 adsorption capacity is comparable to those of ZIFs (zeolitic imidazolate framework) and MOFs based on InIII under the same measurement conditions.[3g, 9, 10, 18] Qst for CO2 is 20.7 kJ mol 1 at zero coverage (Figure S22). The high CO2 uptake might be due to the interaction between Et4N + and CO2.[3, 9, 10] Compared to CH4 uptake, FJI-C1 shows a much larger capacity for CO2 uptake at room temperature and 1 bar (Figure S20b). Therefore, the results demonstrate that the material has the ability to selectively adsorb CO2 in the presence of CH4. Adsorption selectivity was calculated using the ideal ad-

www.chemsuschem.org sorption solution theory (IAST) based upon experimental CO2 and CH4 isotherms. The adsorption selectivity is defined as Si/j = (q1/q2)/(p1/p2), where qi is the amount of i adsorbed and pi is the partial pressure of i in the mixture. At 1 bar, the predicted CO2/CH4 selectivity is 5.89 at 298 K (Figure S24) for equimolar gas-phase mixtures, which is two times higher than that of MOF-5[19] and ZIF-8.[20] Light-hydrocarbon adsorption and separation Because of the similar sizes and volatilities of the light hydrocarbons, separation of these gas mixtures must be performed at low temperatures and high pressures when using traditional cryogenic distillation, causing large energy costs for the chemical industry. Selective adsorption separation of light hydrocarbons is one of the most promising alternative energy- and cost-efficient methods.[21–24] The high porosity and CO2 uptake of FJI-C1 allowed us to evaluate the performance of the anionic framework for light hydrocarbon adsorption. Atmospheric pressure sorption measurements of light hydrocarbons (CH4, C2H2, C2H4, C2H6, and C3H8) were carried out at 273 K and room temperature (Figure 3). At 1 bar and 273 K the adsorption capacities for C3H8 (160.9 cm3 g 1, 316.1 mg g 1), C2H2 (135.9 cm3 g 1, 157.7 mg g 1), C2H6 (123.6 cm3 g 1, 165.5 mg g 1), and C2H4 (85.2 cm3 g 1, 106.5 mg g 1) are very high. As the temperature is increased to room temperature, although its adsorption uptake decreases, FJI-C1 also exhibits notable adsorption capacities for C3H8 (141.9 cm3 g 1, 278.7 mg g 1), C2H2 (93.8 cm3 g 1, 108.9 mg g 1), C2H6 (87.4 cm3 g 1, 117.1 mg g 1), and C2H4 (64.0 cm3 g 1, 1 80.0 mg g ). The adsorption uptake is higher than that of reported MOFs, such as ZIF-8,[24c] UTSA-5a (UTSA stands for University of Texas at San Antonio),[22h] UTSA-33a,[22b] UTSA-35a,[22e] UTSA-36a,[22d] ZJU-30 (ZJU stands for Zhejiang University),[22g] ZJU-48a,[22f] and BIF-24 (BIF stands for boron imidazolate framework).[24e] The high adsorption capacities for light hydrocarbons can be ascribed to the stronger interactions of those gas molecules with the anionic framework and the ethyl groups of the cationic Et4N + in the pores, which provides dis-

Figure 3. Adsorption isotherms of light hydrocarbons (methane, ethylene, ethane, acetylene and propane) on FJI-C1 at (a) 273 and (b) 298 K.

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2014, 7, 2647 – 2653

2649

CHEMSUSCHEM FULL PAPERS persion-force interactions with light hydrocarbons,[4b,c] whereas the organic counter cations selectively block the channels and act as pore gates, providing an opportunity to restrict, trap, or accommodate those gas molecules.[9] Additionally, the s-triazine rings and benzene rings of the compound FJI-C1 may provide p–p interactions with acetylene and ethene to improve their adsorption. Interestingly, the material adsorbs systematically larger amounts of C3 and C2 hydrocarbons in comparison to methane (C1; 9.7 cm3 g 1, 6.9 mg g 1, 298 K, Figure 3). The results demonstrate that FJI-C1 shows adsorption capacities in the following order: C3 > C2 > C1. The uptake trend is different from that of the Fe-MOF-74, for which the adsorption capacities for ethylene and acetylene are higher than for propane due to the olefins and alkynes being coordinated to the open FeII sites.[23] The separation selectivities, based on pure component molar loadings, of C3H8/CH4, C2H2/CH4, C2H6/CH4, and C2H4/CH4 at 298 K are 14.6, 9.7, 9.0, and 6.6, respectively, which are systematically higher than that of UTSA-33a[22b] and UTSA-36a.[22d] Figure 4 b shows the IAST calculations of the adsorption selectivity for binary equimolar C3H8/CH4, C2H2/CH4, C2H6/CH4, and C2H4/ CH4 mixtures at 298 K on compound FJI-C1. The selectivities of C2H4 with respect to CH4 are in excess of 9.8 for a range of pressures up to 100 kPa. Especially the adsorption selectivities

www.chemsuschem.org for the mixtures of C2H2/CH4 and C3H8/CH4 are very high, exceeding 39.3 and 78.7, respectively. It thus indicates that the separation of these binary mixtures is even easier. To understand such high separation selectivity, Qst for C3H8, C2H2, C2H6, C2H4, and CH4 on FJI-C1 is calculated and presented in Figure 4 a. Qst of CH4 (11.4 kJ mol 1) is significantly lower than that of C2H4 (16.3 kJ mol 1), C2H6 (21.7 kJ mol 1), C2H2 (28.9 kJ mol 1), and C3H8 (26.8 kJ mol 1) at zero coverage. Although the Qst values for C3H8 and C2H2 are very similar to each other, the adsorption selectivity according to the IAST calculations for C3H8/C2H2 is high (up to 4.1) at 20–100 kPa (Figure 4 c). With increase of the Qst values, the selectivities for C3H8/C2H2 (1.5–4.1), C3H8/C2H6 (5.0–6.8), C3H8/C2H4 (8.0–9.0), and C3H8/CH4 (78.7–471) also increase at room temperature. The selective adsorption of acetylene over ethylene and ethane is particularly of practical interest for the separation of gas mixtures in numerous industrial applications.[22] Separation of C2 light hydrocarbons is one of the most energy-intensive processes because they are close in molecular sizes. In addition, the purification of ethylene by removing the 1 % acetylene present is essential in producing high-quality polymers.[21a] Interestingly, FJI-C1 shows a high adsorption selectivity towards C2H2/C2H4 (3.7–7.0) and C2H2/C2H6 (2.8–5.8) at room temperature at low pressures (< 10 kPa, Figure 4 d).

Figure 4. (a) Isosteric heats of light hydrocarbon (methane, ethylene, ethane, acetylene and propane) adsorption on FJI-C1. Adsorption selectivities on FJI-C1 at 298 K calculated using IAST for equimolar mixtures of (b) C3H8/CH4, C2H2/CH4, C2H6/CH4, and C2H4/CH4 ; (c) C3H8/CH4, C3H8/C2H4, C3H8/C2H4, and C3H8/C2H2 ; and (d) C2H2/C2H4, C2H2/C2H6 and C2H6/C2H4.

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2014, 7, 2647 – 2653

2650

CHEMSUSCHEM FULL PAPERS

www.chemsuschem.org

Vapor adsorption and separation Separation of organic solvents is of great importance for industry and the environment. Traditional distillation approaches for these challenging separations are inefficient and cause high energy costs. Adsorption separation, as an alternative method, has been proven to be a feasible approach because of high selectivity and regeneration.[25] Compared with conventional microporous adsorbent materials such as activated carbon or zeolites, MOFs have advantages for the separation of specific vapors owing to their tuned pore sizes and large surface areas.[26, 27] However, the effective separation of polar and nonpolar molecules is still a challenge when using MOFs because most of the vapor separation is ascribed to the molecule shape selectivity.[26a] The ligand TATB3 of FJI-C1 has three nitrogen atoms in the s-triazine ring, which allows the peripheral benzene rings to be co-planar with the central ring. This not only ensures a nearly complete delocalization of p electrons and strong ligand–adsorbate p–p interaction to improve the adsorption, but also provides a hydrophilic environment on the internal surfaces of the pores to enhance the separation of the polar and nonpolar molecules. Inspired by the structural features of FJI-C1, we anticipate that it has potential for the separation of organic solvents. Therefore, the vapor sorption isotherms of water, methanol, ethanol, cyclohexane, benzene, and toluene were measured. Interestingly, as shown in Figure 5, this reveals that FJI-C1 has much larger uptake capacities of polar molecule vapors than of nonpolar molecules at low pressure, except for benzene molecules. The adsorption amounts of ethanol, water, and methanol vary incrementally from 5.71 to 19.28 wt %, which are higher than those for MOFs based on flexible tetrapodal ligands.[26a] Thus, FJI-C1 shows a strong affinity to polar molecules, which was attributed to the surfaces of internal hydrophilic pores. Although benzene is a nonpolar molecule, it is surprising that FJI-C1 exhibits a higher benzene adsorption uptake (13.38 wt %) than that of ethanol (5.71 wt %), even higher than that of water at low pressure (11.83 wt %, P/Po < 0.70). It should be noted that the benzene molecule has a delocalized p electron system. As is known, the p electrons of the s-triazine rings in FJI-C1 are nearly completely delocalized, which could provide p–p interactions that cause higher benzene uptake. A similar phenomenon is also found in the p-conjugated nonpolar molecule toluene (1.57 wt %), of which the amount of vapor adsorption is much higher than that the of non-conjugated nonpolar cyclohexane molecule (0.11 wt %). However, the toluene adsorption uptake is lower than that of benzene and ethanol. This may be ascribed to the size of the toluene molecule (molecular area 55.3 2) being larger than that of benzene (48.2 2).[26a] The separation of C6H6 and C6H12 in the petrochemical industry is another key challenge because of their very similar boiling points (C6H6 : 80.1 8C; C6H12 : 80.7 8C).[28] It is difficult to separate benzene and cyclohexane using distillation, which is a highly demanding and energy-consuming process. Interest 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Figure 5. Adsorption isotherms of FJI-C1 for methanol, water, benzene, ethanol, toluene, and cyclohexane at 298 K; Po is the saturated vapor pressure at 298 K.

ingly, FJI-C1 shows a very high adsorption selectivity for C6H6 (13.38 wt %) over C6H12 (0.12 wt %, Figure 5) although both are nonpolar molecules. The size of C6H6 (48.2 2) is larger than that of C6H12 (47.5 2); however, the former has a delocalized p electron system and the acidity of C(sp2) H is relatively stronger than that of C(sp3) H. Therefore, the benzene molecules can provide host–guest interactions with the C6H6 rings and striazine rings of FJI-C1, which results in a higher uptake.

Conclusions We have successfully synthesized the highly porous anionic MOF (Et4N)3[In3(TATB)4] (FJI-C1; TATB3 = 4,4’,4’’-s-triazine-2,4,6triyltribenzoate) based on a tricarboxylate ligand TATB3 , which contains the organic cation Et4N + in the pores. The material with large unit cell volumes and high surface areas shows high CO2 adsorption uptake and moderate hydrogen and CH4 adsorption capacities. Furthermore, it not only exhibits high C2 and C3 hydrocarbon adsorption uptake, but also shows high separation selectivity for C2 and C3 hydrocarbons over CH4 ; C3H8 towards C2 hydrocarbons; and C2H2 over C2H4 and C2H6. In addition, FJI-C1 has a large adsorption uptake of polar vapors and high selectivity for polar vapors over nonpolar molecules. Surprisingly, it not only gives exceptional high benzene uptake in comparison to some polar molecules (e.g., ethanol), but also shows high separation selectivity for benzene over cyclohexane owing to p–p interactions between benzene molecules and the s-triazine rings of the porous material.

Experimental Section General procedures H3TATB was synthesized according to literature.[13e,h] Unless otherwise stated all other chemicals (98%) were commercial available from Sigma-Aldrich company and used without further purification. FTIR spectra were recorded on a Magna 750 FTIR spectrometer with samples prepared as KBr pellets in the range of 450– ChemSusChem 2014, 7, 2647 – 2653

2651

CHEMSUSCHEM FULL PAPERS 4000 cm 1. PXRD patterns were recorded on a Rigaku-Dmax2500 diffractometer using CuKa radiation (l = 0.154 nm). The simulated powder patterns were calculated using Mercury 2.0. Carbon, hydrogen, and nitrogen content was determined using a Vario MICRO E III elemental analyzer. Thermogravimetric analyses were performed on an SDT Q600 unit (TA Instruments) at a heating rate of 10 8C min 1 under a nitrogen atmosphere. A crystal of FJI-C1 was sealed in a nuclear magnetic tube and treated with 0.6 mL [D6]DMSO (dimethoxy sulfoxide) and one drop of DCl/D2O (37 wt%). The 1H NMR spectra of FJI-C1 after hydrolysis were recorded at ambient temperature on a BRUKER AVANCE III spectrometer. The chemical shifts were referenced to tetramethylsilane in the solvent signal in [D6]DMSO. The mass spectrum of FJI-C1 after hydrolysis was tested in ion trap mass spectrometry with DECAX-30000 LCQ Deca XP (ThermoFinnigan). The single-component gas adsorption isotherms were measured on an accelerated surface area and porosimetry 2020 System (Micrometitics). The high pressure gas adsorption isotherms for measured and CH4 were measured on a HTP1 V (Hiden). The organic vapor sorption isotherms were measured on the Intelligent Gravimetric Sorption Analyser IGA100B (Hiden Corporation). The FJI-C1 sample was exchanged with CH2Cl2 three times in three days, followed by degassing at 50 8C for 12 h prior to gas and vapor sorption measurements.

Crystal structure determination of FJI-C1 Single-crystal XRD pattern were collected at 173 K using an Oxford Cryosystems low temperature device attached to an Oxford Diffraction SuperNova dual wavelength diffractometer equipped with an Atlas CCD (charge-coupled device) detector and operating mirror monochromated CuKa radiation mode (l = 1.54184 ). A FJI-C1 crystal was fixed in a loop (Figure S1). Absorption corrections were applied using the SADABS (area detector absorption) programm. Structures were solved by direct methods using SHELX-97 and were refined by full-matrix least-squares on F2 using SHELX-97. Non-hydrogen atoms were refined using anisotropic displacement parameters during the final cycles. Hydrogen atoms were placed in calculated positions with isotropic displacement parameters set to 1.2  Ueq of the attached atom. Some of the Et4N + and the solvents (DEF and water) molecules in FJI-C1 are highly disordered, and attempts to locate and refine these solvent peaks were unsuccessful. Contributions to scattering from some Et4N + and the solvents (DEF and water) molecules were removed using the SQUEEZE option of PLATON. CCDC 970484 contains the supplementary crystallographic data for FJI-C1. These data can be obtained free of charge from the Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.

Synthesis of FJI-C1 Ligand H3TATB (0.6 mmol), In(NO3)3·x H2O (0.4 mmol) and tetraethylammonium chloride (0.4 mmol) were dissolved in DEF (6 mL). The mixture was placed in a 20 mL vial and heated at 120 8C for 3 days and then cooled to RT within one day. After washing with ethanol (three times), colorless crystals were obtained with about 86 % yield based on InIII. IR (KBr): n˜ = 3399 (br), 3051 (w), 2984 (w), 1710(m), 1609 (m), 1565 (m), 1356 (s), 1262 (m), 1168 (m), 1132 (m), 1102 (w), 1016 (m), 881 (m), 815 (m), 774 (m), 731 (m), 700 (m), 503 cm 1 (w); Elemental analysis calcd. C 55.80, H 7.16, N 10.09, found: C 51.11, H 6.29, N 8.37 %  2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org Acknowledgements We acknowledge the financial support from the 973 Program (2011CB932504 and 2013CB933200), NSFC (21273238 and 21221001, 21331006), the NSF of Fujian Province (2014J05022) and Chunmiao Project of Haixi Institute of Chinese Academy of Sciences (CMZX-2014-004). We also acknowledge Prof. Jian Zhang, Jianxin Chen, Xiaoying Huang, and Daqiang Yuan for the crystal structures determination and Prof. Jian Zhang and Daqiang Yuan for the hydrocarbon sorption measurements. Keywords: adsorption · anionic materials · indium · light hydrocarbons · metal–organic frameworks [1] a) J.-R. Li, J. Sculley, H.-C. Zhou, Chem. Rev. 2012, 112, 869; b) M. O’Keeffe, O. M. Yaghi, Chem. Rev. 2012, 112, 675; c) L. Ma, C. Abney, W. Lin, Chem. Soc. Rev. 2009, 38, 1248; d) O. K. Farha, J. T. Hupp, Acc. Chem. Res. 2010, 43, 1166; e) A. Corma, H. Garca, F. X. L. i. Xamena, Chem. Rev. 2010, 110, 4606. [2] a) J. Park, Z. U. Wang, L.-B. Sun, Y.-P. Chen, H.-C. Zhou, J. Am. Chem. Soc. 2012, 134, 20110; b) J.-R. Li, J. Yu, W. Lu, L.-B. Sun, J. Sculley, P. B. Balbuena, H.-C. Zhou, Nat. Commun. 2013, 4, 1538; c) J. Park, D. Yuan, K. T. Pham, J.-R. Li, A. Yakovenko, H.-C. Zhou, J. Am. Chem. Soc. 2012, 134, 99; d) T.-F. Liu, Y.-P. Chen, A. A. Yakovenko, H.-C. Zhou, J. Am. Chem. Soc. 2012, 134, 17358; e) Y. Liu, J.-R. Li, W. M. Verdegaal, T.-F. Liu, H.-C. Zhou, Chem. Eur. J. 2013, 19, 5637. [3] a) S.-T. Zheng, C. Mao, T. Wu, S. Lee, P. Feng, X. Bu, J. Am. Chem. Soc. 2012, 134, 11936; b) S.-T. Zheng, T. Wu, C. Chou, A. Fuhr, P. Feng, X. Bu, J. Am. Chem. Soc. 2012, 134, 4517; c) S.-T. Zheng, T. Wu, F. Zuo, C. Chou, P. Feng, X. Bu, J. Am. Chem. Soc. 2012, 134, 1934; d) S.-T. Zheng, F. Zuo, T. Wu, B. Irfanoglu, C. Chou, R. A. Nieto, P. Feng, X. Bu, Angew. Chem. Int. Ed. 2011, 50, 1849; Angew. Chem. 2011, 123, 1889; e) S.-T. Zheng, J. J. Bu, T. Wu, C. Chou, P. Feng, X. Bu, Angew. Chem. Int. Ed. 2011, 50, 8858; Angew. Chem. 2011, 123, 9020; f) S.-T. Zheng, J. T. Bu, Y. Li, T. Wu, F. Zuo, P. Feng, X. Bu, J. Am. Chem. Soc. 2010, 132, 17062; g) S. Chen, J. Zhang, T. Wu, P. Feng, X. Bu, J. Am. Chem. Soc. 2009, 131, 16027. [4] a) D. Peralta, G. Chaplais, A. Simon-Masseron, K. Barthelet, C. Chizallet, A.-A. Quoineaud, G. D. Pirngruber, J. Am. Chem. Soc. 2012, 134, 8115; b) E. Quartapelle Procopio, F. Linares, C. Montoro, V. Colombo, A. Maspero, E. Barea, J. A. R. Navarro, Angew. Chem. Int. Ed. 2010, 49, 7308; Angew. Chem. 2010, 122, 7466; c) M.-H. Choi, H. J. Park, D. H. Hong, M. P. Suh, Chem. Eur. J. 2013, 19, 17432. [5] a) X. Gu, Z. Lua, Q. Xu, Chem. Commun. 2010, 46, 7400; b) K. C. Stylianou, R. Heck, S. Y. Chong, J. Bacsa, J. T. A. Jones, Y. Z. Khimyak, D. Bradshaw, M. J. Rosseinsky, J. Am. Chem. Soc. 2010, 132, 4119; c) J. Yu, Y. Cui, C. Wu, Y. Yang, Z. Wang, M. O’Keeffe, B. Chen, G. Qian, Angew. Chem. Int. Ed. 2012, 51, 10542; Angew. Chem. 2012, 124, 10694. [6] Y. Liu, V. Ch. Kravtsov, M. Eddaoudi, Angew. Chem. Int. Ed. 2008, 47, 8446; Angew. Chem. 2008, 120, 8574. [7] a) Y. L. Liu, V. Ch. Kravtsov, R. Larsen, M. Eddaoudi, Chem. Commun. 2006, 1488; b) Y. Liu, V. Ch. Kravtsov, R. D. Walsh, P. Poddar, H. Srikanth, M. Eddaoudi, Chem. Commun. 2004, 2806; c) S. Wang, T. Zhao, G. Li, L. Wojtas, Q. Huo, M. Eddaoudi, Y. Liu, J. Am. Chem. Soc. 2010, 132, 18038; d) M. H. Alkordi, J. A. Brant, L. Wojtas, V. Ch. Kravtsov, A. J. Cairns, M. Eddaoudi, J. Am. Chem. Soc. 2009, 131, 17753; e) D. F. Sava, V. Ch. Kravtsov, F. Nouar, L. Wojtas, J. F. Eubank, M. Eddaoudi, J. Am. Chem. Soc. 2008, 130, 3768; f) M. H. Alkordi, Y. Liu, R. W. Larsen, J. F. Eubank, M. Eddaoudi, J. Am. Chem. Soc. 2008, 130, 12639; g) D. F. Sava, V. Ch. Kravtsov, J. Eckert, J. F. Eubank, F. Nouar, M. Eddaoudi, J. Am. Chem. Soc. 2009, 131, 10394; h) F. Nouar, J. Eckert, J. F. Eubank, P. Forster, M. Eddaoudi, J. Am. Chem. Soc. 2009, 131, 2864; i) Y. Liu, V. Ch. Kravtsov, D. A. Beauchamp, J. F. Eubank, M. Eddaoudi, J. Am. Chem. Soc. 2005, 127, 7266. [8] a) S. Yang, X. Lin, W. Lewis, M. Suyetin, E. Bichoutskaia, J. E. Parker, C. C. Tang, D. R. Allan, P. J. Rizkallah, P. Hubberstey, N. R. Champness, K. M. Thomas, A. J. Blake, M. Schrçder, Nat. Mater. 2012, 11, 710; b) S. Yang, L. Liu, J. Sun, K. M. Thomas, A. J. Davies, M. W. George, A. J. Blake, A. H. Hill, A. N. Fitch, C. C. Tang, M. Schrçder, J. Am. Chem. Soc. 2013, 135,

ChemSusChem 2014, 7, 2647 – 2653

2652

CHEMSUSCHEM FULL PAPERS

[9]

[10]

[11] [12] [13]

[14] [15] [16] [17] [18] [19] [20]

[21]

[22]

4954; c) S. Yang, X. Lin, A. J. Blake, G. S. Walker, P. Hubberstey, N. R. Champness, M. Schrçder, Nat. Chem. 2009, 1, 487. a) S. Yang, J. Sun, A. J. Ramirez-Cuesta, S. K. Callear, W. I. F. David, D. Anderson, R. Newby, A. J. Blake, J. E. Parker, C. C. Tang, M. Schrçder, Nat. Chem. 2012, 4, 887; b) S. Yang, X. Lin, A. J. Blake, K. M. Thomas, P. Hubberstey, N. R. Champness, M. Schrçder, Chem. Commun. 2008, 6108; c) S. Yang, G. S. B. Martin, J. J. Titman, A. J. Blake, D. R. Allan, N. R. Champness, M. Schrçder, Inorg. Chem. 2011, 50, 9374. a) Z. Lin, T. Liu, Y. Huang, J. L, R. Cao, Chem. Eur. J. 2012, 18, 7896; b) Z. Lin, Y. Huang, T. Liu, X. Li, R. Cao, Inorg. Chem. 2013, 52, 3127; c) Z. Lin, L. Han, D. Wu, Y. Huang, R. Cao, Cryst. Growth Des. 2013, 13, 255; d) T. Liu, J. L, C. Tian, M. Cao, Z. Lin, R. Cao, Inorg. Chem. 2011, 50, 2264. D. J. Tranchemontagne, J. L. Mendoza-Corts, M. O’Keeffe, O. M. Yaghi, Chem. Soc. Rev. 2009, 38, 1257. C. J. Doonan, W. Morris, H. Furukawa, O. M. Yaghi, J. Am. Chem. Soc. 2009, 131, 9492. a) S. Ma, D. Yuan, X.-S. Wang, H.-C. Zhou, Inorg. Chem. 2009, 48, 2072; b) D. Sun, S. Ma, Y. Ke, T. M. Petersen, H.-C. Zhou, Chem. Commun. 2005, 2663; c) S. Ma, D. Sun, M. Ambrogio, J. A. Fillinger, S. Parkin, H.-C. Zhou, J. Am. Chem. Soc. 2007, 129, 1858; d) D. Sun, Y. Ke, D. J. Collins, G. A. Lorigan, H.-C. Zhou, Inorg. Chem. 2007, 46, 2725; e) S. Ma, H.-C. Zhou, J. Am. Chem. Soc. 2006, 128, 11734; f) S. Ma, X.-S. Wang, D. Yuan, H.-C. Zhou, Angew. Chem. Int. Ed. 2008, 47, 4130; Angew. Chem. 2008, 120, 4198; g) D. Sun, S. Ma, Y. Ke, D. J. Collins, H.-C. Zhou, J. Am. Chem. Soc. 2006, 128, 3896; h) Y. K. Park, S. B. Choi, H. Kim, K. Kim, B.-H. Won, K. Choi, J.-S. Choi, W.-S. Ahn, N. Won, S. Kim, D. H. Jung, S.-H. Choi, G.-H. Kim, S.-S. Cha, Y. H. Jhon, J. K. Yang, J. Kim, Angew. Chem. Int. Ed. 2007, 46, 8230; Angew. Chem. 2007, 119, 8378; i) H. Furukawa, J. Uribe-Romo, J. Kim, M. O’Keeffe, O. M. Yaghi, Inorg. Chem. 2011, 50, 9147; j) H. Zhang, N. Li, C. Tian, T. Liu, F. Du, P. Lin, Z. Li, S. Du, Cryst. Growth Des. 2012, 12, 670; k) W. Gao, F. Xing, D. Zhou, M. Shao, S. Zhu, Inorg. Chem. Commun. 2011, 14, 601. P. van der Sluis, A. L. Spek, Acta Crystallogr. Sect. A 1990, A46, 194. J. L. C. Rowsell, O. M. Yaghi, J. Am. Chem. Soc. 2006, 128, 1304. C. R. Reid, K. M. Thomas, Langmuir 1999, 15, 3206. D. Yuan, D. Zhao, D. Sun, H.-C. Zhou, Angew. Chem. Int. Ed. 2010, 49, 5357; Angew. Chem. 2010, 122, 5485. R. Banerjee, A. Phan, B. Wang, C. Knobler, H. Furukawa, M. O’Keeffe, O. M. Yaghi, Science 2008, 319, 939. A. Demessence, D. M. D’Alessandro, M. L. Foo, J. R. Long, J. Am. Chem. Soc. 2009, 131, 8784. a) S. Keskin, T. M. van Heest, D. S. Sholl, ChemSusChem 2010, 3, 879; b) J. M. Simmons, H. Wu, W. Zhou, T. Yildirim, Energy Environ. Sci. 2011, 4, 2177. a) N. Nijem, H. Wu, P. Canepa, A. Marti, K. J. Balkus, Jr., T. Thonhauser, J. Li, Y. J. Chabal, J. Am. Chem. Soc. 2012, 134, 15201; b) K. Li, D. H. Olson, J. Seidel, T. J. Emge, H. Gong, H. Zeng, J. Li, J. Am. Chem. Soc. 2009, 131, 10368; c) Z. Zhang, Y. Zhao, Q. Gong, Z. Li, J. Li, Chem. Commun. 2013, 49, 653. a) Y. He, R. Krishna, B. Chen, Energy Environ. Sci. 2012, 5, 9107; b) Y. He, Z. Zhang, S. Xiang, F. R. Fronczek, R. Krishna, B. Chen, Chem. Eur. J. 2012, 18, 613; c) Y. He, Z. Zhang, S. Xiang, H. Wu, F. R. Fronczek, W. Zhou, R. Krishna, M. O’Keeffe, B. Chen, Chem. Eur. J. 2012, 18, 1901; d) M. C. Das, H. Xu, S. Xiang, Z. Zhang, H. D. Arman, G. Qian, B. Chen, Chem. Eur. J. 2011, 17, 7817; e) Y. He, Z. Zhang, S. Xiang, F. R. Fronczek, R. Krishna, B. Chen, Chem. Commun. 2012, 48, 6493; f) H. Xu, J. Cai, S. Xiang, Z. Zhang, C. Wu, X. Rao, Y. Cui, Y. Yang, R. Krishna, B. Chen, G.

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

[23]

[24]

[25] [26]

[27]

[28]

Qian, J. Mater. Chem. A 2013, 1, 9916; g) J. Cai, J. Yu, H. Xu, Y. He, X. Duan, Y. Cui, C. Wu, B. Chen, G. Qian, Cryst. Growth Des. 2013, 13, 2094; h) G. Chen, Z. Zhang, S. Xiang, B. Chen, CrystEngComm 2013, 15, 5232; i) Y. Hu, S. Xiang, W. Zhang, Z. Zhang, L. Wang, J. Bai, B. Chen, Chem. Commun. 2009, 7551; j) S. Xiang, Z. Zhang, C.-G. Zhao, K. Hong, X. Zhao, D.-R. Ding, M.-H. Xie, C.-D. Wu, M. C. Das, R. Gill, K. M. Thomas, B. Chen, Nat. Commun. 2011, 2, 204; k) H. Wu, Q. Gong, D. H. Olson, J. Li, Chem. Rev. 2012, 112, 836; l) Z. R. Herm, E. D. Bloch, J. R. Long, Chem. Mater. 2014, 26, 323. a) S. J. Geier, J. A. Mason, E. D. Bloch, W. L. Queen, M. R. Hudson, C. M. Brown, J. R. Long, Chem. Sci. 2013, 4, 2054; b) E. D. Bloch, W. L. Queen, R. Krishna, J. M. Zadrozny, C. M. Brown, J. R. Long, Science 2012, 335, 1606. a) S. Horike, K. Kishida, Y. Watanabe, Y. Inubushi, D. Umeyama, M. Sugimoto, T. Fukushima, M. Inukai, S. Kitagawa, J. Am. Chem. Soc. 2012, 134, 9852; b) S. Wuttke, P. Bazin, A. Vimont, C. Serre, Y.-K. Seo, Y. K. Hwang, J.-S. Chang, G. Frey, M. Daturi, Chem. Eur. J. 2012, 18, 11959; c) U. Bçhme, B. Barth, C. Paula, A. Kuhnt, W. Schwieger, A. Mundstock, J. Caro, M. Hartman, Langmuir 2013, 29, 8592; d) P. Verma, X. Xu, D. G. Truhlar, J. Phys. Chem. C 2013, 117, 12648; e) H.-X. Zhang, H.-R. Fu, H.-Y. Li, J. Zhang, X. Bu, Chem. Eur. J. 2013, 19, 11527; f) P. L. Llewellyn, P. Horcajada, G. Maurin, T. Devic, N. Rosenbach, S. Bourrelly, C. Serre, D. Vincent, S. Loera-Serna, Y. Filinchuk, G. Frey, J. Am. Chem. Soc. 2009, 131, 13002; g) S. Couck, E. Gobechiya, C. E. A. Kirschhock, P. Serra-Crespo, J. Juan-AlcaÇiz, A. M. Joaristi, E. Stavitski, J. Gascon, F. Kapteijn, G. V. Baron, J. F. M. Denayer, ChemSusChem 2012, 5, 740. M. Minceva, A. E. Rodrigues, Ind. Eng. Chem. Res. 2002, 41, 3454. a) T. Liu, J. L, X. Lin, R. Cao, Chem. Commun. 2010, 46, 8439; b) J.-K. Sun, M. Ji, C. Chen, W.-G. Wang, P. Wang, R.-P. Chen, J. Zhang, Chem. Commun. 2013, 49, 1624; c) I. A. Ibarra, J. W. Yoon, J.-S. Chang, S. K. Lee, V. M. Lynch, S. M. Humphrey, Inorg. Chem. 2012, 51, 12242; d) S. Bourrelly, B. Moulin, A. Rivera, G. Maurin, S. Devautour-Vinot, C. Serre, T. Devic, P. Horcajada, A. Vimont, G. Clet, M. Daturi, J.-C. Lavalley, S. Loera-Serna, R. Denoyel, P. L. Llewellyn, G. Frey, J. Am. Chem. Soc. 2010, 132, 9488; e) P. Trens, H. Belarbi, C. Shepherd, P. Gonzalez, N. A. Ramsahye, U. Lee, Y.-K. Seo, J.-S. Chang, J. Phys. Chem. C 2012, 116, 25824; f) M. T. Luebbers, T. Wu, L. Shen, R. I. Masel, Langmuir 2010, 26, 11319; g) C. Yang, U. Kaipa, Q. Z. Mather, X. Wang, V. Nesterov, A. F. Venero, M. A. Omary, J. Am. Chem. Soc. 2011, 133, 18094; h) A. Shigematsu, T. Yamada, H. Kitagawa, J. Am. Chem. Soc. 2012, 134, 13145. a) V. Finsy, C. E. A. Kirschhock, G. Vedts, M. Maes, L. Alaerts, D. E. De Vos, G. V. Baron, J. F. M. Denayer, Chem. Eur. J. 2009, 15, 7724; b) L. Alaerts, C. E. A. Kirschhock, M. Maes, M. A. van der Veen, V. Finsy, A. Depla, J. A. Martens, G. V. Baron, P. A. Jacobs, J. F. M. Denayer, D. E. De Vos, Angew. Chem. Int. Ed. 2007, 46, 4293; Angew. Chem. 2007, 119, 4371; c) Z. Jin, H.-Y. Zhao, X.-J. Zhao, Q.-R. Fang, J. R. Long, G.-S. Zhu, Chem. Commun. 2010, 46, 8612; d) T. Remy, L. Ma, M. Maes, D. E. De Vos, G. V. Baron, J. F. M. Denayer, Ind. Eng. Chem. Res. 2012, 51, 14824; e) Q.-K. Liu, J.-P. Ma, Y.-B. Dong, J. Am. Chem. Soc. 2010, 132, 7005; f) M. P. M. Nicolau, P. S. Brcia, J. M. Gallegos, J. A. C. Silva, A. E. Rodrigues, B. Chen, J. Phys. Chem. C 2009, 113, 13173. Y.-X. Tan, Y.-P. He, J. Zhang, ChemSusChem 2012, 5, 1597.

Received: March 21, 2014 Published online on July 11, 2014

ChemSusChem 2014, 7, 2647 – 2653

2653

Porous anionic indium-organic framework with enhanced gas and vapor adsorption and separation ability.

A three-dimensional microporous anionic metal-organic framework (MOF) (Et4N)3[In3(TATB)4] (FJI-C1, H3TATB=4,4',4''-s-triazine-2,4,6-triyltribenzoic ac...
583KB Sizes 0 Downloads 5 Views