A water stable microporous metal–organic framework based on rod SBUs: synthesis, structure and adsorption properties

Jiaqi Yuan a, Li Mu b, Jiantang Li a, Lirong Zhang a, Guanghua Li a, Qisheng Huo a and Yunling Liu *a
aState Key Laboratory of Inorganic Synthesis and Preparative Chemistry, College of Chemistry, Jilin University, Changchun 130012, P. R. China. E-mail: yunling@jlu.edu.cn; Fax: +86 431 85168624; Tel: +86 431 85168614
bSchool of Bioscience and Technology, Changchun University, Changchun 130023, P. R. China

Received 15th January 2018 , Accepted 15th March 2018

First published on 16th March 2018


Abstract

A novel water stable porous metal–organic framework (MOF), [Cu3(OH)2(IBA)2(CH3COO)2]·DMA (1) (HIBA = 4-(imidazol-1-yl)-benzoic acid, DMA = N,N-dimethylacetamide), has been solvothermally synthesized and structurally characterized. The framework of compound 1 is constructed from rod SBUs with a new (5,8)-connected topology. Compound 1 exhibits exceptional chemical stability under various conditions (organic solvents, water, boiling water or pH = 1–8 acidic and basic aqueous solutions). Meanwhile, it exhibits good selectivity for CO2 over CH4 (6.6 for CO2/CH4 = 0.5/0.5 and 5.8 for CO2/CH4 = 0.05/0.95 under 1 bar at 298 K). The excellent stability of compound 1 is of extreme importance for its practical industrial applications.


Introduction

In the past two decades, porous metal–organic frameworks (MOFs), as an arising class of functional materials, which are made up of inorganic metal ions or clusters connected by organic linkers through coordination bonds, have gained considerable attention from both academia and industry.1 Owing to their tunable pore size, modifiable pore surface and milder synthesis conditions, they have exhibited a wide range of potential applications (gas adsorption and separation,2 catalysis,3 drug delivery,4 luminescence sensing,5 magnetism,6etc.).

As one of the primary greenhouse gases, carbon dioxide has attracted the attention of scientists all over the world not only for its possibility to cause global warming, but also because of the troubles it can cause when using natural gas since it can lower the energy conversion rate.7 Therefore, effective and selective adsorption of CO2 is significant for natural gas purification.8 MOFs, as a new class of sorbent materials, are quite qualified for such an application demand.

However, most of the earlier reported MOFs, and even some of the milestone MOFs, such as MOF-5,9 are sensitive to water conditions due to the instability of ligand–metal bonds.10 The structures of MOFs are vulnerable to the attack of water molecules, which might lead to phase changes, ligand displacement and structure decomposition.11 Compared with porous carbon materials, zeolites and metal oxides, the limited thermal and chemical stability (especially water stability) of MOFs hampers their popularity in industrial applications, as water or moisture commonly exists in most industrial processes.12 Consequently, water stable MOFs, such as UiO-66,13 NU-1000,14 DUT-51,15 and MIL-127,16 have been in great demand when considering these materials for adsorption applications. Although thousands of different MOFs have been reported in the literature, only a few of them have high chemical stability, especially water stability.

Generally, water stable MOFs can be classified into three categories: (1) metal azolate frameworks with N-donor ligands, such as zeolite-like imidazole frameworks (ZIFs);17 (2) metal carboxylate frameworks consisting of multivalent metal ions, such as MIL-101 and UiO-66;13,18 and (3) MOFs without open metal sites or decorated by a hydrophobic pore surface.11a However, only N-donor ligands, such as imidazole, can easily form mononuclear metal nodes, just like ZIFs which are normally constructed from mononuclear Zn(II), Co(II) and Cd(II). In addition, multivalent metal ions, such as Cr3+ and Zr4+, are usually costly and need complex synthesis. Thus, in order to obtain highly stable and at the same time structurally diverse MOFs, we choose a linear ligand: HIBA (HIBA = 4-(imidazol-1-yl)-benzoic acid), which contains both carboxylate and imidazole groups in consideration of the following points: (1) the multiple coordination mode of the carboxylate groups is able to form a variety of multinuclear metal nodes; (2) the N-coordination point from the imidazole group can effectively prevent the formation of open metal sites and commendably enhance the chemical and thermal stability of the framework. As a consequence, a novel water stable porous rod SBU-based metal–organic framework: [Cu3(OH)2(IBA)2(CH3COO)2]·DMA (1), has been successfully synthesized.19 Adjacent rod SBUs are connected with each other via the ligand to form a three-dimensional framework. Because of the introduction of the imidazole group and the stable rod SBUs, compound 1 has an exceptional chemical stability, especially water stability. The outstanding stability of compound 1 is necessary when considering it for practical adsorption applications.12 In addition, the adsorption behaviour of activated compound 1 for some small gases (CO2, CH4 and C2H6) has been analysed. Ideal adsorbed solution theory (IAST) calculations have also been performed to investigate the gas selectivity.

Experimental

Materials and methods

All chemicals were obtained from commercial sources and used without further purification. Powder X-ray diffraction (PXRD) data were collected using a Rigaku D/max-2550 diffractometer with Cu-Kα radiation (λ = 1.5418 Å). Elemental analyses (C, H, and N) were performed by using a vario MICRO (Elementar, Germany). Thermal gravimetric analyses (TGA) were performed on a TGA Q500 thermogravimetric analyzer used in air with a heating rate of 10 °C min−1.

Synthesis of compound 1

A mixture of Cu(NO3)2·3H2O (0.010 g, 0.041 mmol), HIBA (0.004 g, 0.021 mmol), DMA (1.5 mL), H2O (0.5 mL), and HNO3 (300 μL, 2.2 mL HNO3 in 10 mL DMF) was added into a 20 mL glass vial. The glass vial was sonicated for 10 minutes until a clear and transparent solution was obtained. The sealed vial was heated at 105 °C for 36 hours. Green stick crystals were obtained, and then collected, washed with DMA, and dried in air. Elemental analysis (%) calc. for C28H34Cu3N6O10: C 40.98; H 3.86; N 8.92, found: C 41.73; H 4.22; N 10.43. The experimental powder X-ray diffraction (XRD) pattern of compound 1 agrees well with the simulated one based on the single-crystal X-ray data, indicating that compound 1 is a pure phase (Fig. S1).

X-ray crystallography

Crystallographic data for compound 1 was collected using a Bruker Apex II CCD diffractometer with graphite-monochromated Mo-Kα (λ = 0.71073 Å) radiation at room temperature. The structure was solved by direct methods and refined by full-matrix least-squares on F2 using SHELXTL-NT version 5.1.20 All metal atoms were located first, and then the oxygen and carbon atoms of the compound were subsequently found in difference Fourier maps. The hydrogen atoms of the ligand were placed geometrically. All non-hydrogen atoms were refined anisotropically. The final formula was derived from the crystallographic data combined with the elemental and thermogravimetric analysis data. The detailed crystallographic data and selected bond lengths and angles for compound 1 are listed in Tables S1 and S2, respectively. Topology information for compound 1 was calculated using TOPOS 4.0.21

Gas adsorption measurements

N2, CO2, CH4 and C2H6 gas adsorption measurements were performed on Micromeritics ASAP 2420, Micromeritics ASAP 2010 and Micromeritics 3-Flex instruments. Before gas adsorption measurements, the microcrystalline samples of compound 1 were exchanged with fresh ethanol 8 times for 3 days to completely remove the nonvolatile solvent molecules, which can be proved by TGA (Fig. S5). The samples were activated by drying under dynamic vacuum at 80 °C for 30 minutes. Before gas adsorption measurements, the samples were dried again by using the ‘outgas’ function of the surface area analyzer for 10 hours at 120 °C.

Results and discussion

Crystal structure descriptions

Single-crystal X-ray diffraction analysis shows that compound 1 crystallizes in the monoclinic crystal system with the space group C2/c. Structural analysis reveals that there exist novel 1D infinite rod-shaped metal chain SBUs in the crystal structure. As shown in Fig. 1a, there are two types of Cu(II) atoms in the rod SBU. One is five-coordinated by two μ3-OH, two carboxylate groups and one nitrogen atom. The other is six-coordinated by two μ3-OH and four carboxylate groups. Saturated coordinated Cu(II) atoms are combined to form a rod-shaped metal chain, which can be simplified as a series of edge-sharing octahedra.19a It is worth noting that the saturated metal ions can effectively improve the stability of the structure. The angular ligand HIBA links the metal chains together to form a 3D framework. As shown in Fig. 1b and c, along the [001] direction, there exists a one-dimensional channel with a diameter of 6.2 Å × 5.3 Å considering the van der Waals radius. The position of four carboxylate carbon atoms and two nitrogen atoms can be considered to be located at the vertices of a distorted octahedron as shown in Fig. S10. These edge-sharing octahedra consist of infinite rod SBUs, which are connected with each other through the linear ligand HIBA to form a 3D framework. The position of the N atom can be treated as a 5-c node and the C atom as an 8-c node. Thus, the whole structure can be described as a (5,8)-connected binodal network, which belongs to a new topology with a Schläfli symbol of (34·43·52·6) (38·49·56·65). The topological characteristics exhibited by the tiles are illustrated in Fig. 1d. The total solvent-accessible volume has been calculated using PLATON. The volume is equal to 3241.3 Å3 per unit cell, which takes up approximately 24.7% of the cell volume, exhibiting porosity and offering possibilities for gas adsorption.
image file: c8ce00064f-f1.tif
Fig. 1 Single-crystal structure of compound 1: (a) topology simplification of the metal chain and the ligand; (b) ball and stick model of the 3D framework of compound 1 viewed along the [001] direction; (c) space-filling view of the structure of compound 1 along the [001] direction; (d) topological features of compound 1 displayed by tiles and face symbols for yellow and green are (38·4·2·62·82) and (38).

Thermogravimetric and stability analysis

Thermogravimetric analysis (TGA) measurement indicated that compound 1 exhibits a three-step weight loss (Fig. S5). The first slight weight loss of 6.8% was observed prior to 200 °C, which can be attributed to the removal of guest molecules (DMA) (calcd 6.2%).22 The weight loss of 58.7% between 250 and 400 °C can be associated with the removal of organic ligands and structure collapse (calcd 59.2%). The last remaining weight loss at 500 °C of 29.7% is attributed to the formation of CuO (calcd 29.8%).

In order to explore the chemical stability of compound 1, first, the as-synthesized samples of compound 1 were soaked in water at room temperature and 105 °C for 3 days, respectively. The corresponding experimental PXRD patterns indicate that the samples still retain their high crystallinity and agree well with those simulated from the single-crystal structure data (Fig. 2a), demonstrating the extraordinary water stability of compound 1. Furthermore, the stability of compound 1 in acidic and basic aqueous solutions was investigated. As shown in Fig. 2b, the framework of compound 1 can be retained for at least 24 hours in pH 1.0 HCl and pH 8.0 NaOH solutions, and only a part collapsed after being soaked for over 48 hours. Also, compound 1 shows high stability in common organic solvents, including EtOH, CH3CN, CH3COCH3 and CH2Cl2 (Fig. 2c). Finally, in order to confirm the thermal stability of compound 1, temperature-dependent PXRD was conducted. As Fig. 2d shows, the samples of compound 1 can remain crystalline up to 200 °C. When it was heated to 250 °C, the framework of compound 1 partly collapsed. Combined with the above research, compound 1 has an excellent chemical and thermal stability. Compared with many well-known Cu-based MOFs, such as HKUST-1,23 MOF-14,24etc., compound 1 performs better in terms of chemical (especially water) stability. The notable chemical stability of compound 1 towards water, acidic and basic conditions indicates that it is a promising material for practical industrial applications.25


image file: c8ce00064f-f2.tif
Fig. 2 PXRD patterns of compound 1: after treatment in water at room temperature and 105 °C (a); after treatment under acidic and basic conditions (b); after being soaked in different organic solvents for one week (c); the temperature-dependent PXRD pattern (d).

Gas adsorption and separation behaviors

To verify the porosity of compound 1, N2 adsorption measurement was carried out at 77 K. Fig. S2 shows a near type-I sorption isotherm for N2 at 77 K, with an adsorption amount of 70 cm3 g−1 at 1 bar. The Brunauer–Emmett–Teller (BET) surface area was calculated to be 197 m2 g−1, while the Langmuir surface area was 274 m2 g−1. The micropore volume is 0.11 cm3 g−1, which is close to the theoretical value of 0.17 cm3 g−1. In order to prove the stability of compound 1, N2 adsorption measurement of the water-treated samples was also carried out (Fig. S11). The N2 isotherm type is similar to the untreated one. The adsorption amount of N2 is 45.5 cm3 g−1 at 77 K under 1 bar, which is relatively less than the amount of untreated samples. As shown in Fig. 2a, there are some little differences in the PXRD patterns between compound 1 and the water-treated samples, indicating that the framework of compound 1 has been only partly collapsed.

The CO2 adsorption performance was explored and the adsorption isotherms are shown in Fig. 3a. The amount of CO2 uptake for compound 1 is 35 and 26 cm3 g−1 at 273 and 298 K under 1 bar, respectively. To investigate the affinity of the framework of compound 1 to CO2, the isosteric CO2 adsorption enthalpy (Qst) on compound 1 was calculated using the virial model (Fig. 3b). Compound 1 shows a near-zero coverage with a Qst value of 29.6 kJ mol−1. This phenomenon indicates the strong van der Waals interactions between the framework and the gas molecules at low pressure.


image file: c8ce00064f-f3.tif
Fig. 3 CO2 (a), CH4 (c), and C2H6 (d) adsorption isotherms for compound 1 at 273 K and 298 K under 1 bar; Qst of CO2 for compound 1 (b).

Due to its good stability, compound 1 has potential storage and selective separation applications for CO2 and small hydrocarbons (CH4, C2H6, etc.), which are the main components of natural gas. The adsorption isotherms of CH4 and C2H6 are evaluated at 273 and 298 K under 1 bar. The maximum adsorption capacities for CH4 are 13 and 6 cm3 g−1 and for C2H6 are 28 and 25 cm3 g−1, respectively (Fig. 3c and d). The adsorption capacity for C2H6 is higher than that for CH4. This adsorption capacity trend of C2H6 > CH4 is consistent with those reported in some previous studies.26 As estimated from the sorption isotherms at 273 and 298 K, the Qst values at zero coverage are 30 and 31 kJ mol−1 for CH4 and C2H6 adsorption, respectively (Fig. S6 and S7).

In order to evaluate the practical separation ability of compound 1 for CO2, the gas selectivity for CO2/CH4 (5% and 95%, 50% and 50%) and C2H6/CH4 was calculated via IAST, which is a common method used to calculate binary mixture gas adsorption based on experimental single-component isotherms.27 By using the dual-site Langmuir–Freundlich equation to fit the data (Fig. 4a and c), we successfully made the models fit the isotherms at 298 K very well (R2 > 0.9999).28 Then, the fitting parameters are used to calculate the multi-component adsorption with IAST (Table S3). As shown in Fig. 4b, the selectivity of compound 1 for CO2 over CH4 according to the experimental data is 6.6 and 5.8 at 298 K and 1 bar, which surpass the values for MOF-5, JLU-Liu37 (3.8),29 ZJU-60 (5–5.5),30 SNNU-22 (4.5) and many reported carbon materials under the same measurement conditions.19b Also, the potential separation application of typical small hydrocarbons is investigated for compound 1 using IAST. As shown in Fig. 4d, the selectivity for C2H6 over CH4 for equimolar mixture is 19 at 298 K and 1 bar.


image file: c8ce00064f-f4.tif
Fig. 4 CO2, CH4 and C2H6 adsorption isotherms at 298 K along with the dual-site Langmuir–Freundlich (DSLF) fits (a and c); gas mixture adsorption selectivity is predicted using IAST at 298 K and 100 kPa for compound 1 (b and d).

Conclusions

In summary, we have successfully solvothermally synthesized a water stable microporous MOF material with a (5,8)-connected rod-packing network topology. Compound 1 exhibits satisfactory chemical stability, especially water stability, as well as certain adsorption performance for small gases and good selectivity for CO2/CH4. The excellent stability of compound 1 is essential for its practical industrial applications in the field of gas separation in the near future.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (No. 21771078 and 21621001).

Notes and references

  1. (a) H. C. Zhou, J. R. Long and O. M. Yaghi, Chem. Rev., 2012, 112, 673 CrossRef CAS PubMed ; (b) H. C. Zhou and S. Kitagawa, Chem. Soc. Rev., 2014, 43, 5415 RSC ; (c) B. Li, H. M. Wen, Y. Cui, W. Zhou, G. Qian and B. Chen, Adv. Mater., 2016, 28, 8819 CrossRef CAS PubMed ; (d) Y. Cui, B. Li, H. He, W. Zhou, B. Chen and G. Qian, Acc. Chem. Res., 2016, 49, 483 CrossRef CAS PubMed ; (e) Q. G. Zhai, X. Bu, C. Mao, X. Zhao, L. Daemen, Y. Cheng, A. J. Ramirez-Cuesta and P. Feng, Nat. Commun., 2016, 7, 13645 CrossRef CAS PubMed ; (f) B. Li, H. M. Wen, Y. Cui, W. Zhou, G. Qian Rogow, J. A. Mason, T. M. McDonald, E. D. Bloch, Z. R. Herm, T. H. Bae and J. R. Long, Chem. Rev., 2012, 112, 724–781 CrossRef PubMed ; (g) X. Luo, Y. Cao, T. Wang, G. Li, Y. Yang, Z. Xu, J. Zhang, Q. Huo, Y. Liu and M. Eddaoudi, J. Am. Chem. Soc., 2016, 138, 786–789 CrossRef CAS PubMed .
  2. (a) K. Adil, Y. Belmabkhout, R. S. Pillai, A. Cadiau, P. M. Bhatt, A. H. Assen, G. Maurin and M. Eddaoudi, Chem. Soc. Rev., 2017, 46, 3402–3430 RSC ; (b) R. Banerjee, A. Phan, B. Wang, C. Knobler, H. Furukawa, M. O'Keeffe and O. M. Yaghi, Science, 2008, 319, 939–943 CrossRef CAS PubMed ; (c) J.-R. Li, J. Sculley and H.-C. Zhou, Chem. Rev., 2012, 112, 869–932 CrossRef CAS PubMed ; (d) Z. J. Zhang, Y. G. Zhao, Q. H. Gong, Z. Li and J. Li, Chem. Commun., 2013, 49, 653–661 RSC ; (e) M. P. Suh, H. J. Park, T. K. Prasad and D.-W. Lim, Chem. Rev., 2012, 112, 782–835 CrossRef CAS PubMed ; (f) X.-F. Wang, Y.-B. Zhang, H. Huang, J.-P. Zhang and X.-M. Chen, Cryst. Growth Des., 2008, 8, 4559–4563 CrossRef CAS ; (g) X.-F. Wang, Y.-K. Sun and S.-H. Wang, Microporous Mesoporous Mater., 2013, 181, 262–269 CrossRef CAS .
  3. (a) J. Y. Lee, O. K. Farha, J. Roberts, K. A. Scheidt, S. B. T. Nguyen and J. T. Hupp, Chem. Soc. Rev., 2009, 38, 1450–1459 RSC ; (b) L. Ma, C. Abney and W. Lin, Chem. Soc. Rev., 2009, 38, 1248–1256 RSC ; (c) M. Yoon, R. Srirambalaji and K. Kim, Chem. Rev., 2012, 112, 1196–1231 CrossRef CAS PubMed ; (d) Z. Weng, Y. Wu, M. Wang, J. Jiang, K. Yang, S. Huo, X.-F. Wang, Q. Ma, G. W. Brudvig, V. S. Batista, Y. Liang, Z. Feng and H. Wang, Nat. Commun., 2018, 9, 415 CrossRef PubMed .
  4. (a) N. Ahmad, H. A. Younus, A. H. Chughtai and F. Verpoort, Chem. Soc. Rev., 2015, 44, 9–25 RSC ; (b) P. Horcajada, R. Gref, T. Baati, P. K. Allan, G. Maurin, P. Couvreur, G. Ferey, R. E. Morris and C. Serre, Chem. Rev., 2012, 112, 1232–1268 CrossRef CAS PubMed ; (c) J. Wang, D. Ma, W. Liao, S. Li, M. Huang, H. Liu, Y. Wang, R. Xie and J. Xu, CrystEngComm, 2017, 19, 5244–5250 RSC .
  5. (a) W. P. Lustig, S. Mukherjee, N. D. Rudd, A. V. Desai, J. Li and S. K. Ghosh, Chem. Soc. Rev., 2017, 46, 3242–3285 RSC ; (b) L. Basabe-Desmonts, D. N. Reinhoudt and M. Crego-Calama, Chem. Soc. Rev., 2007, 36, 993–1017 RSC ; (c) A. J. Lan, K. H. Li, H. H. Wu, D. H. Olson, T. J. Emge, W. Ki, M. C. Hong and J. Li, Angew. Chem., Int. Ed., 2009, 48, 2334–2338 CrossRef CAS PubMed .
  6. (a) E. Coronado and G. Minguez Espallargas, Chem. Soc. Rev., 2013, 42, 1525–1539 RSC ; (b) J. Wang, X. Jing, Y. Cao, G. Li, Q. Huo and Y. Liu, CrystEngComm, 2015, 17, 604 RSC ; (c) X.-W. Wu, F. Pan, D. Zhang, G.-X. Jin and J.-P. Ma, CrystEngComm, 2017, 19, 5864–5872 RSC .
  7. (a) K. Sumida, D. L. Rogow, J. A. Mason, T. M. McDonald, E. D. Bloch, Z. R. Herm, T.-H. Bae and J. R. Long, Chem. Rev., 2012, 112, 724–781 CrossRef CAS PubMed ; (b) J.-R. Li, R. J. Kuppler and H.-C. Zhou, Chem. Soc. Rev., 2009, 38, 1477 RSC ; (c) R. E. Morris and P. S. Wheatley, Angew. Chem., Int. Ed., 2008, 47, 4966 CrossRef CAS PubMed ; (d) J. Yu, L.-H. Xie, J.-R. Li, Y. Ma, J. M. Seminario and P. B. Balbuena, Chem. Rev., 2017, 117, 9674–9754 CrossRef CAS PubMed .
  8. (a) S. Yao, X. Sun, B. Liu, R. Krishna, G. Li, Q. Huo and Y. Liu, J. Mater. Chem. A, 2016, 4, 15081–15087 RSC ; (b) D. Wang, T. Zhao, G. Li, Q. Huo and Y. Liu, Dalton Trans., 2014, 43, 2365–2368 RSC .
  9. H. Li, M. Eddaoudi, M. O'Keeffe and O. M. Yaghi, Nature, 1999, 402, 276 CrossRef CAS .
  10. (a) L. Huang, H. Wang, J. Chen, Z. Wang, J. Sun, D. Zhao and Y. Yan, Microporous Mesoporous Mater., 2003, 58, 105 CrossRef CAS ; (b) P. M. Schoenecker, C. G. Carson, H. Jasuja, C. J. J. Flemming and K. S. Walton, Ind. Eng. Chem. Res., 2012, 51, 6513 CrossRef CAS .
  11. (a) C. Wang, X. Liu, N. K. Demir, J. P. Chen and K. Li, Chem. Soc. Rev., 2016, 45, 5107–5134 RSC ; (b) J. Canivet, A. Fateeva, Y. Guo, B. Coasne and D. Farrusseng, Chem. Soc. Rev., 2014, 43, 5594 RSC ; (c) A. J. Howarth, Y. Liu, P. Li, Z. Li, T. C. Wang, J. T. Hupp and O. K. Farha, Nat. Rev. Mater., 2016, 1, 15018 CrossRef CAS .
  12. N. C. Burtch, H. Jasuja and K. S. Walton, Chem. Rev., 2014, 114, 10575–10612 CrossRef CAS PubMed .
  13. M. Kandiah, M. H. Nilsen, S. Usseglio, S. Jakobsen, U. Olsbye, M. Tilset, C. Larabi, E. A. Quadrelli, F. Bonino and K. P. Lillerud, Chem. Mater., 2010, 22, 6632–6640 CrossRef CAS .
  14. J. E. Mondloch, W. Bury, D. Fairen-Jimenez, S. Kwon, E. J. Demarco, M. H. Weston, A. A. Sarjeant, S. T. Nguyen, P. C. Stuir, O. K. Farha and J. T. Hupp, J. Am. Chem. Soc., 2013, 135, 10294–10297 CrossRef CAS PubMed .
  15. V. Bon, V. Senkovska and S. Kaskel, Chem. Commun., 2012, 48, 8407–8409 RSC .
  16. D. Cunha, M. B. Yahia, S. Hall, S. R. Miller, H. Chevreau, E. ElkaÏm, G. Maurin, P. Horcajada and C. Serre, Chem. Mater., 2013, 25, 2767–2776 CrossRef CAS .
  17. (a) K. S. Park, Z. Ni, A. P. Cote, J. Y. Choi, R. Huang, F. J. Uribe-Romo, H. K. Chae, M. O'Keeffe and O. M. Yaghi, Proc. Natl. Acad. Sci. U. S. A., 2006, 103, 10186–10191 CrossRef CAS PubMed ; (b) F. Wang, H.-R. Fu, Y. Kang and J. Zhang, Chem. Commun., 2014, 50, 12065–12068 RSC ; (c) F. Wang, Z.-S. Liu, H. Yang, Y.-X. Tan and J. Zhang, Angew. Chem., Int. Ed., 2011, 50, 450–453 CrossRef CAS PubMed .
  18. G. Ferey, C. Mellot-Draznieks, C. Serre, F. Millange, J. Dutour, S. Surble and I. Margiolaki, Science, 2005, 309, 5743 CrossRef PubMed .
  19. (a) A. Schoedel, M. Li, D. Li, M. O'Keeffe and O. M. Yaghi, Chem. Rev., 2016, 116, 12466–12535 CrossRef CAS PubMed ; (b) J.-W. Zhang, M.-C. Hu, S.-N. Li, Y.-C. Jiang and Q.-G. Zhai, Dalton Trans., 2017, 46, 836 RSC .
  20. G. M. Sheldrick, SHELXTL-NT, version 5.1, Bruker AXS Inc., Madison, WI, 1997 Search PubMed .
  21. V. A. Blatov, A. P. Shevchenko and D. M. Proserpio, Cryst. Growth Des., 2014, 14, 3576–3586 CAS .
  22. S. S. Nagarkar, S. M. Unni, A. Sharma, S. Kurungot and S. K. Ghosh, Angew. Chem., Int. Ed., 2014, 53, 2638–2642 CrossRef CAS PubMed .
  23. J. L. C. Rowsell and O. M. Yaghi, J. Am. Chem. Soc., 2006, 128, 1304–1315 CrossRef CAS PubMed .
  24. B. Chen, M. Eddaoudi, S. T. Hyde, M. O'Keeffe and O. M. Yaghi, Science, 2001, 291, 1021–1023 CrossRef CAS PubMed .
  25. D.-X. Xue, Y. Belmabkhout, O. Shekhah, H. Jiang, K. Adil, A. J. Cairns and M. Eddaoudi, J. Am. Chem. Soc., 2015, 137, 5034–5040 CrossRef CAS PubMed .
  26. (a) Y. He, Z. Zhang, S. Xiang, F. R. Fronczek, R. Krishna and B. Chen, Chem. Commun., 2012, 48, 6493–6495 RSC ; (b) Y. He, W. Zhou, R. Krishna and B. Chen, Chem. Commun., 2012, 48, 11813–11831 RSC ; (c) H. X. Zhang, H. R. Fu, H. Y. Li, J. Zhang and X. Bu, Chem. – Eur. J., 2013, 19, 11527–11530 CrossRef CAS PubMed .
  27. (a) N. U. Qadir, S. A. M. Said and H. M. Bahaidarah, Microporous Mesoporous Mater., 2015, 201, 61 CrossRef CAS ; (b) J. L. C. Rowsell and O. M. Yaghi, J. Am. Chem. Soc., 2006, 128, 1304–1315 CrossRef CAS PubMed ; (c) A. L. Myers and J. M. Prausnitz, AIChE J., 1965, 11, 121 CrossRef CAS .
  28. (a) W. Lu, D. Yuan, J. Sculley, D. Zhao, R. Krishna and H. C. Zhou, J. Am. Chem. Soc., 2011, 133, 18126–18129 CrossRef CAS PubMed ; (b) S. D. Burd, S. Ma, J. A. Perman, B. J. Sikora, R. Q. Snurr, P. K. Thallapally, J. Tian, L. Wojtas and M. J. Zaworotko, J. Am. Chem. Soc., 2012, 134, 3663–3666 CrossRef CAS PubMed ; (c) H. Zhao, Z. Jin, H. Su, J. Zhang, X. Yao, H. Zhao and G. Zhu, Chem. Commun., 2013, 49, 2780–2782 RSC .
  29. J. Li, X. Luo, N. Zhao, L. Zhang, Q. Huo and Y. Liu, Inorg. Chem., 2017, 56, 4141–4147 CrossRef CAS PubMed .
  30. X. Duan, Q. Zhang, J. Cai, Y. Yang, Y. Cui, Y. He, C. Wu, R. Krishna, B. Chen and G. Qian, J. Mater. Chem. A, 2014, 2, 2628–2633 CAS .

Footnote

Electronic supplementary information (ESI) available: Materials and methods, crystal data and structure refinement, structure information, XRD, TGA, gas adsorption and adsorption selectivity. CCDC 1517751. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c8ce00064f

This journal is © The Royal Society of Chemistry 2018