Two Finite Binuclear [M2(μ2-OH)(COO)2] (M = Co ... - ACS Publications

Mar 17, 2017 - The volume is 5878.7 Å3 per unit cell, which accounts for approximately 65.3% of the cell volume. This result means that JLU-Liu37 pos...
0 downloads 8 Views 2MB Size
Article pubs.acs.org/IC

Two Finite Binuclear [M2(μ2‑OH)(COO)2] (M = Co, Ni) Based Highly Porous Metal−Organic Frameworks with High Performance for Gas Sorption and Separation Jiantang Li, Xiaolong Luo, Nian Zhao, Lirong Zhang,* Qisheng Huo, and Yunling Liu* State Key Laboratory of Inorganic Synthesis and Preparative Chemistry, College of Chemistry, Jilin University, Changchun 130012, P. R. China S Supporting Information *

ABSTRACT: Two highly porous MOFs, [Co2(μ2-OH)(bpdc)(Htpim)2][SiF 6]·3.5DMA·2.5CH3OH (JLU-Liu37, H2bpdc = biphenyl-4,4′-dicarboxylate, Htpim = 2,4,5-tri(4pyridyl)imidazole) and [Ni2(μ2-OH)(bpdc)(Htpim)2][SiF6]· 7.5DMA·6CH3OH (JLU-Liu38), have been solvothermally synthesized by using the mixed ligand strategy. Both of the compounds possess finite binuclear [M2(μ2-OH)(COO)2] (M = Co, Ni) secondary building units (SBUs) which formed with a polar functional group, μ2-OH. JLU-Liu37 and JLU-Liu38 exhibit notable adsorption capacities for CO2 and light hydrocarbons (CH4, C2H6, and C3H8). Moreover, both of the materials exhibit arrestive natural gas selective separation ability, especially for C3H8/CH4 (206 for an equimolar mixture under 1 bar and 298 K, for JLU-Liu37). Both of the MOFs may be effectively applied in the separation of industrial light hydrocarbons.



and outstanding selective separations of CO2 over CH4 or N2.37 Up to now, although lots of infinite one-periodic binuclear [M2(μ2-OH)(COO)2] SBUs have been synthesized, finite binuclear [M2(μ2-OH)(COO)2] nodes have been rarely reported.38−40 In order to obtain highly functionalized MOFs, some great success has been achieved by preparing MOFs with mixed ligand strategies. Two or more kinds of ligands may bring greater diversity in MOF structures. In terms of structural design, it is well-documented that size-matching or symmetrymatching ligands are suitable for constructing anticipated channels or cavities of coordination frameworks. Moreover, as previously described, “acid and base ligands are perfect partners that can compensate charge balance, coordination deficiency, repulsive vacuum, and weak interaction all at once.”41−48 In this regard, acid ligand biphenyl-4,4′-dicarboxylate (H2 bpdc) and base ligand 2,4,5-tri(4-pyridyl)imidazole (Htpim) with similar size and symmetry have been chosen. We successfully utilize these two ligands and cobalt or nickel nitrates to obtain two new MOFs, [Co2(μ2-OH)(bpdc)(Htpim) 2 ][SiF 6 ]·3.5DMA·2.5CH 3 OH (JLU-Liu37) and [Ni 2 (μ 2 -OH)(bpdc)(Htpim) 2 ][SiF 6 ]·7.5DMA·6CH 3 OH (JLU-Liu38). The two compounds are isoreticular with each other and possess unusual finite binuclear [M2(μ2-OH)(COO)2] (M = Co, Ni) SBUs and two mixed ligands. Compared to many other classical SBUs or molecular building

INTRODUCTION Nowadays, more and more attention has been paid to energy and environmental problems that seriously affect human survival. Therefore, how to efficiently utilize renewable energy, such as natural gas (NG), has become particularly important. However, the existence of small amounts of CO2 and light hydrocarbon impurities hinder the conversion rate and energy content. As a consequence, effective separation of CH4 from the feed gas mixture has become indispensable.1,2 During the past 2 decades, metal−organic frameworks (MOFs) have become a new generation of porous hybrid materials due to their high structural and functional tunability.3−6 Numerous MOFs have been proved to have excellent performance in gas adsorption and separation,7−9 luminescence,10,11 sensing,12,13 catalysis,14,15 magnetism,16−19 and so on. Since MOFs have been considered as a kind of promising material for gas storage and separation, several strategies have been explored for improving their properties, such as introduction of Lewis basic sites (LBSs),20−23 providing open metal sites (OMSs),24−28 pore space partition (PSP), and so on.29,30 However, some of the functionalization methods are likely to result in the instability of structures and the complexity of synthesis.31−33 As a unique and effective way, in situ introducing of polar functional groups (−NH2, −OH, −F, etc.) into SBUs has been proved to have an excellent influence on gas adsorption and separation abilities.34−36 For example, Hong and co-workers reported InOF-1, which owned chiral 41 In(OH)(CO2)2 helix chains and exhibited high CO2 uptake © 2017 American Chemical Society

Received: January 17, 2017 Published: March 17, 2017 4141

DOI: 10.1021/acs.inorgchem.7b00156 Inorg. Chem. 2017, 56, 4141−4147

Article

Inorganic Chemistry blocks (MBB), MOFs constructed of finite [M2(μ2-OH)(COO)2] SBUs are relatively rare.49 Both materials exhibit high porosity and thermal stability. In addition, the two compounds demonstrate notable adsorption capacity for CO2 and light hydrocarbons. Moreover, the selectivity for CO2/CH4, C2H6/ CH4, and C3H8/CH4 were appraised by the ideal adsorbed solution theory (IAST).



each bpdc2− ligand links two inorganic SBUs, and each Htpim ligand links three inorganic SBUs. Figure 1a unambiguously

EXPERIMENTAL SECTION

Materials and Methods. All of the chemicals were obtained from commercial channels and used without further purification. Powder Xray diffraction (PXRD) data was tested on a Rigaku D/max-2550 diffractometer (Cu Kα radiation, λ = 1.5418 Å). Elemental analyses (C, H, and N) were collected by a vario MICRO cube (Elementar). The thermal gravimetric analyses (TGA) were performed on a TGA Q500 thermogravimetric analyzer (air, heating rate of 10 °C min−1). BET surface area was measured by N2 adsorption isotherms at 77 K with a Micrometrics ASAP 2420 instrument. The CO2 adsorption isotherm of JLU-Liu38 was measured at 273 and 298 K with a Micrometrics ASAP 2010. The CO2 adsorption isotherm of JLULiu37 was measured at 273 and 298 K with a Micrometrics ASAP 2020. CH4, C2H6, and C3H8 gas adsorption measurements were performed on a Micromeritics ASAP 2020 instrument. Synthesis of JLU-Liu37. A solid mixture of H2bpdc (0.02 g, 0.083 mmol), Htpim (0.047 g, 0.17 mmol), Co(NO3)2·6H2O (0.047 g, 0.16 mmol), DMA (10 mL), MeOH (10 mL), and HBF4 (0.9 mL) was sealed in a 40 mL vial and then heated at 100 °C for 24 h. Red block crystals were collected and washed with DMA and then dried in air [76% yield based on Co(NO3)2·6H2O]. Elemental analysis (wt %) for JLU-Liu37: C, 53.21; H, 5.034; N, 12.60. Found: C, 53.05; H, 5.077; N, 12.44. Synthesis of JLU-Liu38. The procedure was the same as that for JLU-Liu37, except that Co(NO3)2·6H2O was replaced by Ni(NO3)2· 6H2O (0.047 g, 0.16 mmol). Green block crystals were collected and washed with DMA and then dried in air [79% yield based on Ni(NO3)2·6H2O]. Elemental analysis (wt %) for JLU-Liu38: C, 52.66; H, 6.403; N, 12.50. Found: C, 52.48; H, 6.289; N, 12.66. X-ray Crystallography. Crystallographic data of the two compounds were collected on a Bruker Apex II CCD diffractometer using graphite-monochromated Mo Kα (λ = 0.71073 Å) radiation at room temperature. The structures of these two compounds were solved by direct methods and refined by full-matrix least-squares on F2 using SHELXL-2014. All the metal atoms were located first, and then the oxygen, carbon, nitrogen, fluorine, and silicon atoms of the compounds were subsequently found in difference Fourier maps. The hydrogen atoms of the ligands were placed geometrically. All metal, oxygen, carbon, and nitrogen atoms were refined anisotropically. All fluorine and silicon atoms were refined isotropically. The final formula was derived from crystallographic data combined with elemental and thermogravimetric analysis data. The detailed crystallographic data and selected bond lengths and angles for JLU-Liu37 and JLU-Liu38 are listed in Tables S1−S3 of the Supporting Information (SI). Topology information was calculated through TOPOS 4.0.50

Figure 1. Single-crystal structure of JLU-Liu37 and JLU-Liu38 (without the guest molecules): (a) binuclear [M2(μ2-OH)(COO)2] (M = Co and Ni) SBUs consisting of ligands and metal cores and the formation of a 3D framework from a double reinforcement of the structure by two ligands. (b) CPK model of the channel. (c) The new type of (3,8) connected topology.

illustrates the formation of the 3D framework. On one hand, the finite binuclear nodes are woven by Htpim ligands to configure a 3D framework. On the other hand, the bpdc2− ligands connected each node once again to doubly support the framework and effectively increase the rigidity of the framework. JLU-Liu37 possesses one-dimensional channels, the diameter of which is 10 Å excluding the van der Waals radius. Moreover, an independent [SiF6]− ion, which is bonded to Htpim ligands through a hydrogen bond from imidazole hydrogen, is observed in each channel to balance the framework charge (Figure S3, SI). It is noted that the [SiF6]− ion originates from the glass vial, which reacts with the HF from the decomposition of HBF4. From another viewpoint of topology, the finite Co-based binuclear nodes can be regarded as eight-connected nodes and Htpim ligands can be regarded as three-connected nodes, and JLU-Liu37 adopts a new (3,8) connected topology with a Schläfli symbol of {42·5}2{44·510·69· 74·8} [Figures 1c and S1 (SI)]. The total solvent-accessible volume was calculated by PLATON. The volume is 5878.7 Å3 per unit cell, which accounts for approximately 65.3% of the cell volume. This result means that JLU-Liu37 possesses high porosity and potential applications in gas adsorption. The crystalline phase purities of JLU-Liu37 and JLU-Liu38 were confirmed by comparing experimental PXRD patterns with the simulated one from single-crystal data (Figure S5, SI). Thermogravimetric analysis (TGA) of JLU-Liu37 and JLULiu38 was carried out to investigate their thermal stabilities (Figure S6, SI). The results showed that the JLU-Liu37 and JLU-Liu38 exhibited similar TGA curves. JLU-Liu37 was observed to have a total weight loss of 23.7% before 310 °C, corresponding to the loss of guest molecules, then followed by a relatively steady plateau up to 393 °C. JLU-Liu38 was observed to have a total weight loss of 41.8% before 265 °C, corresponding to the loss of guest molecules, then followed by a relatively steady plateau up to 423 °C. Both of two compounds exhibit high thermal stability.



RESULTS AND DISCUSSION The X-ray crystallographic analysis reveals that JLU-Liu37 and JLU-Liu38 are isostructural with different metals. Here, JLULiu37 is selected for detailed discussion of the structure. JLULiu37 crystallizes in the orthorhombic crystal system with the space group Pnma. The [Co2(μ2-OH)(COO)2] SBU is formed by three pyridyl nitrogen atoms of three different Htpim ligands, two carboxylate oxygen atoms of two different bpdc2− ligands, and one oxygen atom from a bridging μ2-OH hydroxyl group, which is confirmed by bond valence sum (BVS) calculations (Tables S4 and S5, SI).51 Up to now, MOFs based on an eight-connected finite binuclear [Co2(μ2-OH)(COO)2] SBU have been rarely reported (Figure S4, SI). In addition, 4142

DOI: 10.1021/acs.inorgchem.7b00156 Inorg. Chem. 2017, 56, 4141−4147

Article

Inorganic Chemistry

Figure 2. N2 sorption isotherms for JLU-Liu37 (a) and JLU-Liu38 (b) at 77 K.

Figure 3. (a) CO2, (b) CH4, (c) C2H6, and (d) C3H8 gas sorption isotherms for JLU-Liu37 and JLU-Liu38 at 273 and 298 K under 1 bar.

small gas (CO2, CH4, C2H6, and C3H8) adsorption capacities. The CO2 uptake for JLU-Liu37 is 75.5 (14.8%) and 34.8 (6.8%) cm3/g at 273 and 298 K under 1 atm, respectively (Figure 3a). However, the CO2 uptake for JLU-Liu38 reaches 92.6 (18.2%) cm3/g at 273 K under 1 bar, which is higher than that of IFMC-1 (91.4 cm3/g) with nitrogen-rich triazole units under the same conditions.52 To investigate the interaction between CO2 and the framework, the isosteric CO2 adsorption enthalpies (Qst) on JLU-Liu37 and JLU-Liu38 are calculated (Figure S8, SI). At zero coverage, the enthalpy of CO2 adsorption for JLU-Liu37 and JLU-Liu38 are 22 and 24 kJ mol−1, which are comparable to that of LBS-modified HZIrF1R (23 kJ mol−1).53 This phenomenon indicates the strong van der Waals interactions with the functional framework. Moreover, JLU-Liu37 and JLU-Liu38 exhibit notable adsorption capacities for C2H6 (171 and 191 cm3 g−1 at 273 K, 99 and 111 cm3 g−1 at 298 K, respectively) and C3H8 (201 and 212 cm3 g−1 at 273 K, 178 and 188 cm3 g−1 at 298 K, respectively) (Figure 3c,d). These adsorption results are higher than those of many reported MOFs, such as UTSA-35a and BIF-24.54,55 Nevertheless, JLU-Liu37 and JLU-Liu38 adsorb a little amount of

To confirm the permanent porosity of JLU-Liu37 and JLULiu38, homogeneous microcrystalline samples were exchanged in acetone for 72 h. The solution was refreshed three times daily during this time period. The acetone solvent-exchanged sample was degassed under high vacuum at 100 °C for 12 h to obtain the fully evacuated framework, which can be verified by the TGA analysis (Figure S6, SI). The N2 adsorption for these two materials at 77 K exhibited almost the same reversible typeI isotherms, which are characteristic of microporous materials (Figure 2). The BET surface areas of JLU-Liu37 and JLULiu38 were calculated to be 1795 and 1784 m2 g−1, respectively. This phenomenon indicates that the difference in metal nodes does not affect the porosity of either material. The experimental micropore volumes of both materials are 0.87 and 0.88 cm3/g, which are close to the theoretical values of 0.93 and 0.91 cm3/g, respectively. Meanwhile, the pore size of both compounds is approximately 8.6−11 Å through density functional theory (DFT) analysis, which is conformed to the pore diameter measured in the crystal structure. The high porosity and the modified polar functional groups μ2-OH of these two materials inspired us to investigate their 4143

DOI: 10.1021/acs.inorgchem.7b00156 Inorg. Chem. 2017, 56, 4141−4147

Article

Inorganic Chemistry

Figure 4. Gas sorption isotherms of CO2, CH4, C2H6, and C3H8 along with the dual-site Langmuir−Freundlich (DSLF) fits (a, JLU-Liu37; b, JLULiu38); gas mixture adsorption selectivities are predicted by IAST at 298 K under 1 bar (c and e, JLU-Liu37; d and f, JLU-Liu38).

CH4 (18 and 17 cm3 g−1 at 273 K, 10 and 8 cm3 g−1 at 298 K, respectively) (Figure 3b). These results confirm that JLULiu37 and JLU-Liu38 exhibit similar hydrocarbon adsorption capacities with the following trend: C3H8 > C2H6 > CH4. This phenomenon can be attributed to the framework having a higher affinity for the larger and highly polarizable molecules.56,57 This conclusion highlights JLU-Liu37 and JLULiu38 as promising materials for selective separation of these small hydrocarbons (C2H6/CH4, C3H8/CH4) according to the number of carbon atoms. The selectivities of JLU-Liu37 and JLU-Liu38 for C2H6 and C3H8 over CH4 were calculated by IAST.58−60 By using the dual-site Langmuir−Freundlich equation to fit the data (Figure 4), the models fit the isotherms at 298 K very well (R2 > 0.999). The fitting parameters were then used to predict the multicomponent adsorption with IAST (Table S6, SI). At 298 K and 1 bar, the selectivities of JLU-Liu37 and JLU-Liu38 for equimolar C2H6/CH4 binary mixtures are 11 and 15, respectively. This result is comparable to those of UTSA-33a and UTSA-34b.61,62 The C3H8/CH4 selectivities of JLU-Liu37 and JLU-Liu38 are 206 and 98, respectively, which are much higher than the very high value (>80) for UTSA-35a.55 In order to investigate the very high C3H8/CH4 selectivity of JLULiu37, the isosteric heats (Qst) of JLU-Liu37 for CH4 and

C3H8 were calculated (Figures S9 and S11, SI). At zerocoverage, the Qst values of CH4 and C3H8 for JLU-Liu37 are 17.5 and 29.2 kJ mol−1. Thus, the high selectivity may be attributed to the steric effects and host−guest interactions. The high C2H6/CH4 and C3H8/CH4 selectivities suggest that JLULiu37 and JLU-Liu38 may be effectively applied in the separation of industrial light hydrocarbons. In addition, the selectivity of CO2 over CH4 at a general feed composition of landfill gas (50%:50%) and a typical feed composition for natural gas purification (5%:95%) was also calculated (Figure 4c,d). At 298 K and 1 bar, the CO2/CH4 selectivity of JLU-Liu37 is 3.8 and 3.8 and for JLU-Liu38 is 5.6 and 5.4, respectively, which are comparable to those of ZJU-60 (5.5 at equimolar, 298 K, and 1 bar) and MOF-177 (4.4 at equimolar, 298 K, and 1 bar).63,64 These notable results indicate that JLU-Liu37 and JLU-Liu38 have good potential applications for natural gas purification.



CONCLUSION In summary, by utilizing the mixed ligand strategy, two isoreticular MOFs, JLU-Liu37 and JLU-Liu38, have been successfully synthesized. Both of the compounds feature a new topology and finite binuclear [M2(μ2-OH)(COO)2] (M = Co, Ni) SBUs which formed with a polar functional group μ2-OH. 4144

DOI: 10.1021/acs.inorgchem.7b00156 Inorg. Chem. 2017, 56, 4141−4147

Article

Inorganic Chemistry

Chen, B. Pore Chemistry and Size Control in Hybrid Porous Materials for Acetylene Capture From Ethylene. Science 2016, 353, 141−144. (10) Wang, Z.; Qin, L.; Chen, J.; Zheng, H. H-Bonding Interactions Induced Two Isostructural Cd(II) Metal-Organic Frameworks Showing Different Selective Detection of Nitroaromatic Explosives. Inorg. Chem. 2016, 55, 10999−11005. (11) Hu, Z.; Lustig, W. P.; Zhang, J.; Zheng, C.; Wang, H.; Teat, S. J.; Gong, Q.; Rudd, N. D.; Li, J. Effective Detection of Mycotoxins by a Highly Luminescent Metal-Organic Framework. J. Am. Chem. Soc. 2015, 137, 16209−16215. (12) Hao, Z.; Song, X.; Zhu, M.; Meng, X.; Zhao, S.; Su, S.; Yang, W.; Song, S.; Zhang, H. One-Dimensional Channel-Structured Eu-MOF for Sensing Small Organic Molecules and Cu2+ Ion. J. Mater. Chem. A 2013, 1, 11043. (13) Pal, T. K.; Chatterjee, N.; Bharadwaj, P. K. Linker-Induced Structural Diversity and Photophysical Property of MOFs for Selective and Sensitive Detection of Nitroaromatics. Inorg. Chem. 2016, 55, 1741−1747. (14) Mon, M.; Ferrando-Soria, J.; Grancha, T.; Fortea-Pérez, F. R.; Gascon, J.; Leyva-Pérez, A.; Armentano, D.; Pardo, E. Selective Gold Recovery and Catalysis in a Highly Flexible Methionine-Decorated Metal-Organic Framework. J. Am. Chem. Soc. 2016, 138, 7864−7867. (15) Peters, A. W.; Li, Z.; Farha, O. K.; Hupp, J. T. Toward Inexpensive Photocatalytic Hydrogen Evolution: A Nickel Sulfide Catalyst Supported on a High-Stability Metal-Organic Framework. ACS Appl. Mater. Interfaces 2016, 8, 20675−20681. (16) Mon, M.; Pascual-Á lvarez, A.; Grancha, T.; Cano, J.; FerrandoSoria, J.; Lloret, F.; Gascon, J.; Pasán, J.; Armentano, D.; Pardo, E. Solid-State Molecular Nanomagnet Inclusion into a Magnetic MetalOrganic Framework: Interplay of the Magnetic Properties. Chem. - Eur. J. 2016, 22, 539−545. (17) Chen, Q.; Xue, W.; Lin, J.; Wei, Y.; Yin, Z.; Zeng, M.; Kurmoo, M.; Chen, X. Windmill Co4{Co4(μ4-O)} with 16 Divergent Branches Forming a Family of Metal-Organic Frameworks: Organic Metrics Control Topology, Gas Sorption, and Magnetism. Chem. - Eur. J. 2016, 22, 12088−12094. (18) Chen, Q.; Xue, W.; Lin, J. B.; Lin, R. B.; Zeng, M. H.; Chen, X. M. Highly-Connected, Porous Coordination Polymers Based on [M4(μ3-OH)2] (M = CoII and NiII) Clusters: Different Networks, Adsorption and Magnetic Properties. Dalton Trans. 2012, 41, 4199− 4206. (19) Zeng, M. H.; Yin, Z.; Tan, Y. X.; Zhang, W. X.; He, Y. P.; Kurmoo, M. Nanoporous Cobalt(II) MOF Exhibiting Four Magnetic Ground States and Changes in Gas Sorption upon Post-Synthetic Modification. J. Am. Chem. Soc. 2014, 136, 4680−4688. (20) Qiao, Z. W.; Wang, N. Y.; Jiang, J. W.; Zhou, J. Design of Amine-Functionalized Metal-Organic Frameworks for CO2 Separation: the More Amine, the Better? Chem. Commun. 2016, 52, 974− 977. (21) Lee, C.; Huang, H.; Liu, Y.; Luo, T.; Lee, G.; Peng, S.; Jiang, J.; Chao, I.; Lu, K. Cooperative Effect of Unsheltered Amide Groups on CO2 Adsorption Inside Open-Ended Channels of a Zinc(II)-Organic Framework. Inorg. Chem. 2013, 52, 3962−3968. (22) Zheng, B.; Bai, J.; Duan, J.; Wojtas, L.; Zaworotko, M. J. Enhanced CO2 Binding Affinity of a High-Uptake rht-Type MetalOrganic Framework Decorated with Acylamide Groups. J. Am. Chem. Soc. 2011, 133, 748−751. (23) Planas, N.; Dzubak, A. L.; Poloni, R.; Lin, L.; McManus, A.; McDonald, T. M.; Neaton, J. B.; Long, J. R.; Smit, B.; Gagliardi, L. The Mechanism of Carbon Dioxide Adsorption in an AlkylamineFunctionalized Metal-Organic Framework. J. Am. Chem. Soc. 2013, 135, 7402−7405. (24) He, H.; Sun, F.; Aguila, B.; Perman, J. A.; Ma, S.; Zhu, G. A Bifunctional Metal-Organic Framework Featuring the Combination of Open Metal Sites and Lewis Basic Sites for Selective Gas Adsorption and Heterogeneous Cascade Catalysis. J. Mater. Chem. A 2016, 4, 15240−15246. (25) Wang, X.; Ma, S.; Forster, P. M.; Yuan, D.; Eckert, J.; López, J. J.; Murphy, B. J.; Parise, J. B.; Zhou, H. Enhancing H2 Uptake by

These two materials exhibit similar high porosity and notable gas sorption ability for CO2 and light hydrocarbons. Meanwhile, IAST calculation reveals that JLU-Liu37 and JLU-Liu38 display excellent selective separation for C2H6/CH4 and C3H8/ CH4. Finally, these burgeoning highly porous crystalline materials may be effectively applied in the field of separation of industrial light hydrocarbons.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b00156. Procedures to calculate selectivity by IAST, structure information, PXRD, TGA, Qst, crystal data and structure refinement (CCDC 1527949 and 1527950), and BVS calculation (PDF) Crystal data for JLU-Liu37 in CIF format (CIF) Crystal data for JLU-Liu38 in CIF format (CIF)



AUTHOR INFORMATION

Corresponding Authors

*L.Z. e-mail: [email protected]. *Y.L. e-mail: [email protected]. ORCID

Jiantang Li: 0000-0002-8963-5402 Yunling Liu: 0000-0001-5040-6816 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (No. 21373095, 21371067, and 21621001)



REFERENCES

(1) Arenillas, A.; Smith, K. M.; Drage, T. C.; Snape, C. E. CO2 Capture Using some Fly Ash-Derived Carbon Materials. Fuel 2005, 84, 2204−2210. (2) Wang, J.; Huang, L.; Yang, R.; Zhang, Z.; Wu, J.; Gao, Y.; Wang, Q.; O’Hare, D.; Zhong, Z. Recent Advances in Solid Sorbents for CO2 Capture and New Development Trends. Energy Environ. Sci. 2014, 7, 3478−3518. (3) Guillerm, V.; Kim, D.; Eubank, J. F.; Luebke, R.; Liu, X.; Adil, K.; Lah, M. S.; Eddaoudi, M. A Supermolecular Building Approach for the Design and Construction of Metal-Organic Frameworks. Chem. Soc. Rev. 2014, 43, 6141−72. (4) Jiang, J.; Zhao, Y.; Yaghi, O. M. Covalent Chemistry Beyond Molecules. J. Am. Chem. Soc. 2016, 138, 3255−3265. (5) Qian, J.; Jiang, F.; Su, K.; Li, Q.; Yuan, D.; Hong, M. SelfAssembly of Polyhedral Indium−Organic Nanocages. Inorg. Chem. 2014, 53, 12228−12230. (6) Lu, W.; Wei, Z.; Gu, Z. Y.; Liu, T. F.; Park, J.; Park, J.; Tian, J.; Zhang, M.; Zhang, Q.; Gentle, T., III; Bosch, M.; Zhou, H. C. Tuning the Structure and Function of Metal-Organic Frameworks Via Linker Design. Chem. Soc. Rev. 2014, 43, 5561−5593. (7) Liu, B.; Wu, W.; Hou, L.; Li, Z.; Wang, Y. Two Nanocage-Based Metal-Organic Frameworks with Mixed-Cluster SBUs and CO2 Sorption Selectivity. Inorg. Chem. 2015, 54, 8937−8942. (8) Cadiau, A.; Adil, K.; Bhatt, P. M.; Belmabkhout, Y.; Eddaoudi, M. A Metal-Organic Framework-Based Splitter for Separating Propylene From Propane. Science 2016, 353, 137−40. (9) Cui, X.; Chen, K.; Xing, H.; Yang, Q.; Krishna, R.; Bao, Z.; Wu, H.; Zhou, W.; Dong, X.; Han, Y.; Li, B.; Ren, Q.; Zaworotko, M. J.; 4145

DOI: 10.1021/acs.inorgchem.7b00156 Inorg. Chem. 2017, 56, 4141−4147

Article

Inorganic Chemistry “Close-Packing” Alignment of Open Copper Sites in Metal-Organic Frameworks. Angew. Chem., Int. Ed. 2008, 47, 7263−7266. (26) Liu, B.; Yao, S.; Shi, C.; Li, G.; Huo, Q.; Liu, Y. Significant Enhancement of Gas Uptake Capacity and Selectivity Via the Judicious Increase of Open Metal Sites and Lewis Basic Sites within Two Polyhedron-Based Metal-Organic Frameworks. Chem. Commun. 2016, 52, 3223−3226. (27) Caskey, S. R.; Wong-Foy, A. G.; Matzger, A. J. Dramatic Tuning of Carbon Dioxide Uptake via Metal Substitution in a Coordination Polymer with Cylindrical Pores. J. Am. Chem. Soc. 2008, 130, 10870− 10871. (28) Li, J.; Kuppler, R. J.; Zhou, H. Selective Gas Adsorption and Separation in Metal-Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1477−1504. (29) Zhai, Q. G.; Bu, X.; Zhao, X.; Li, D. S.; Feng, P. Pore Space Partition in Metal-Organic Frameworks. Acc. Chem. Res. 2017, 50, 407−417. (30) Zhai, Q. G.; Bu, X.; Mao, C.; Zhao, X.; Daemen, L.; Cheng, Y.; Ramirez-Cuesta, A. J.; Feng, P. An Ultra-Tunable Platform for Molecular Engineering of High-Performance Crystalline Porous Materials. Nat. Commun. 2016, 7, 13645−13643. (31) Kandiah, M.; Nilsen, M. H.; Usseglio, S.; Jakobsen, S.; Olsbye, U.; Tilset, M.; Larabi, C.; Quadrelli, E. A.; Bonino, F.; Lillerud, K. P. Synthesis and Stability of Tagged UiO-66 Zr-MOFs. Chem. Mater. 2010, 22, 6632−6640. (32) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’Keeffe, M.; Yaghi, O. M. Systematic Design of Pore Size and Functionality in Isoreticular MOFs and their Application in Methane Storage. Science 2002, 295, 469−472. (33) Cohen, S. M. Postsynthetic Methods for the Functionalization of Metal-Organic Frameworks. Chem. Rev. 2012, 112, 970−1000. (34) Zhang, Z.; Zhao, Y.; Gong, Q.; Li, Z.; Li, J. MOFs for CO2 Capture and Separation From Flue Gas Mixtures: The Effect of Multifunctional Sites On their Adsorption Capacity and Selectivity. Chem. Commun. 2013, 49, 653−661. (35) Yang, S.; Sun, J.; Ramirez-Cuesta, A. J.; Callear, S. K.; David, W. I. F.; Anderson, D. P.; Newby, R.; Blake, A. J.; Parker, J. E.; Tang, C. C.; Schröder, M. Selectivity and Direct Visualization of Carbon Dioxide and Sulfur Dioxide in a Decorated Porous Host. Nat. Chem. 2012, 4, 887−894. (36) Ibarra, I. A.; Yang, S.; Lin, X.; Blake, A. J.; Rizkallah, P. J.; Nowell, H.; Allan, D. R.; Champness, N. R.; Hubberstey, P.; Schröder, M. Highly Porous and Robust Scandium-Based Metal-Organic Frameworks for Hydrogen Storage. Chem. Commun. 2011, 47, 8304−8306. (37) Qian, J.; Jiang, F.; Yuan, D.; Wu, M.; Zhang, S.; Zhang, L.; Hong, M. Highly Selective Carbon Dioxide Adsorption in a WaterStable Indium-Organic Framework Materialw. Chem. Commun. 2012, 48, 9696−9698. (38) Zhang, L.; Qian, J.; Yang, W.; Kuang, X.; Zhang, J.; Cui, Y.; Wu, W.; Wu, X.; Lu, C.; Chen, W. A (3,8)-Connected Metal-Organic Framework with a Unique Binuclear [Ni2(μ2-OH)(COO)2] Node for High H2 and CO2 Adsorption Capacities. J. Mater. Chem. A 2015, 3, 15399−15402. (39) Zheng, S. T.; Mao, C.; Wu, T.; Lee, S.; Feng, P.; Bu, X. Generalized Synthesis of Zeolite-Type Metal-Organic Frameworks Encapsulating Immobilized Transition-Metal Clusters. J. Am. Chem. Soc. 2012, 134, 11936−11939. (40) Zheng, S. T.; Wu, T.; Irfanoglu, B.; Zuo, F.; Feng, P.; Bu, X. Multicomponent Self-Assembly of a Nested Co24@Co48 MetalOrganic Polyhedral Framework. Angew. Chem., Int. Ed. 2011, 50, 8034−8037. (41) Du, M.; Li, C.; Liu, C.; Fang, S. Design and Construction of Coordination Polymers with Mixed-Ligand Synthetic Strategy. Coord. Chem. Rev. 2013, 257, 1282−1305. (42) Ren, G.; Chang, Z.; Xu, J.; Hu, Z.; Liu, Y.; Xu, Y.; Bu, X. Construction of a Polyhedron Decorated MOF with a Unique Network through the Combination of Two Classic Secondary Building Units. Chem. Commun. 2016, 52, 2079−2082.

(43) Wang, F.; Tang, Y.; Zhang, J. Achievement of Bulky Homochirality in Zeolitic Imidazolate-Related Frameworks. Inorg. Chem. 2015, 54, 11064−11066. (44) Wang, H.; Xu, J.; Zhang, D.; Chen, Q.; Wen, R.; Chang, Z.; Bu, X. Crystalline Capsules: Metal-Organic Frameworks Locked by SizeMatching Ligand Bolts. Angew. Chem., Int. Ed. 2015, 54, 5966−5970. (45) Luo, X.; Cao, Y.; Wang, T.; Li, G.; Li, J.; Yang, Y.; Xu, Z.; Zhang, J.; Huo, Q.; Liu, Y.; Eddaoudi, M. Host-Guest Chirality Interplay: A Mutually Induced Formation of a Chiral ZMOF and its Double-Helix Polymer Guests. J. Am. Chem. Soc. 2016, 138, 786−789. (46) Yin, Z.; Zhou, Y. L.; Zeng, M. H.; Kurmoo, M. The Concept of Mixed Organic Ligands in Metal-Organic Frameworks: Design, Tuning and Functions. Dalton Trans. 2015, 44, 5258−5275. (47) Wang, X.; Gao, W. Y.; Luan, J.; Wojtas, L.; Ma, S. An Effective Strategy to Boost the Robustness of Metal-Organic Frameworks via Introduction of Size-Matching Ligand Braces. Chem. Commun. 2016, 52, 1971−1974. (48) Yuan, S.; Qin, J. S.; Zou, L.; Chen, Y. P.; Wang, X.; Zhang, Q.; Zhou, H. C. Thermodynamically Guided Synthesis of Mixed-Linker Zr-MOFs with Enhanced Tunability. J. Am. Chem. Soc. 2016, 138, 6636−6642. (49) Tranchemontagne, D. J.; Mendoza-Cortés, J. L.; O’Keeffe, M.; Yaghi, O. M. Secondary Building Units, Nets and Bonding in the Chemistry of Metal-Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1257−1283. (50) 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. (51) Brown, I. D. The Chemical Bond in Inorganic Chemistry: The Bond Valence Model; Oxford University Press, 2002. (52) Qin, J.; Du, D.; Li, W.; Zhang, J.; Li, S.; Su, Z.; Wang, X.; Xu, Q.; Shao, K.; Lan, Y. N-Rich Zeolite-Like Metal-Organic Framework with Sodalite Topology: High CO2 Uptake, Selective Gas Adsorption and Efficient Drug Delivery. Chem. Sci. 2012, 3, 2114. (53) Wang, F.; Tang, Y. H.; Zhang, J. Achievement of Bulky Homochirality in Zeolitic Imidazolate-Related Frameworks. Inorg. Chem. 2015, 54, 11064−11066. (54) Zhang, H.; Fu, H.; Li, H.; Zhang, J.; Bu, X. Porous ctn-Type Boron Imidazolate Framework for Gas Storage and Separation. Chem. - Eur. J. 2013, 19, 11527−11530. (55) He, Y.; Zhang, Z.; Xiang, S.; Fronczek, F. R.; Krishna, R.; Chen, B. A Robust Doubly Interpenetrated Metal-Organic Framework Constructed From a Novel Aromatic Tricarboxylate for Highly Selective Separation of Small Hydrocarbons. Chem. Commun. 2012, 48, 6493. (56) Ma, H. F.; Liu, Q. Y.; Wang, Y. L.; Yin, S. G. A Water-Stable Anionic Metal−Organic Framework Constructed from Columnar Zinc-Adeninate Units for Highly Selective Light Hydrocarbon Separation and Efficient Separation of Organic Dyes. Inorg. Chem. 2017, 56, 2919−2925. (57) Kim, H.; Park, J.; Jung, Y. The Binding Nature of Light Hydrocarbons on Fe/MOF-74 for Gas Separation. Phys. Chem. Chem. Phys. 2013, 15, 19644−19650. (58) Burd, S. D.; Ma, S.; Perman, J. A.; Sikora, B. J.; Snurr, R. Q.; Thallapally, P. K.; Tian, J.; Wojtas, L.; Zaworotko, M. J. Highly Selective Carbon Dioxide Uptake by [Cu(bpy-n)2(SiF6)] (bpy-1 = 4,4′-Bipyridine; bpy-2 = 1,2-Bis(4-pyridyl)ethene). J. Am. Chem. Soc. 2012, 134, 3663−3666. (59) Lu, W.; Yuan, D.; Sculley, J.; Zhao, D.; Krishna, R.; Zhou, H. Sulfonate-Grafted Porous Polymer Networks for Preferential CO2 Adsorption at Low Pressure. J. Am. Chem. Soc. 2011, 133, 18126− 18129. (60) Zhao, H.; Jin, Z.; Su, H.; Zhang, J.; Yao, X.; Zhao, H.; Zhu, G. Target Synthesis of a Novel Porous Aromatic Framework and its Highly Selective Separation of CO2/CH4. Chem. Commun. 2013, 49, 2780. (61) He, Y.; Zhang, Z.; Xiang, S.; Fronczek, F. R.; Krishna, R.; Chen, B. A Microporous Metal-Organic Framework for Highly Selective 4146

DOI: 10.1021/acs.inorgchem.7b00156 Inorg. Chem. 2017, 56, 4141−4147

Article

Inorganic Chemistry Separation of Acetylene, Ethylene, and Ethane from Methane at Room Temperature. Chem. - Eur. J. 2012, 18, 613−619. (62) He, Y.; Zhang, Z.; Xiang, S.; Wu, H.; Fronczek, F. R.; Zhou, W.; Krishna, R.; O’Keeffe, M.; Chen, B. High Separation Capacity and Selectivity of C2 Hydrocarbons over Methane within a Microporous Metal-Organic Framework at Room Temperature. Chem. - Eur. J. 2012, 18, 1901−1904. (63) Duan, X.; Zhang, Q.; Cai, J.; Yang, Y.; Cui, Y.; He, Y.; Wu, C.; Krishna, R.; Chen, B.; Qian, G. A New Metal-Organic Framework with Potential for Adsorptive Separation of Methane From Carbon Dioxide, Acetylene, Ethylene, and Ethane Established by Simulated Breakthrough Experiments. J. Mater. Chem. A 2014, 2, 2628. (64) Saha, D.; Bao, Z.; Jia, F.; Deng, S. Adsorption of CO2, CH4, N2O, and N2 on MOF-5, MOF-177, and Zeolite 5A. Environ. Sci. Technol. 2010, 44, 1820−1826.

4147

DOI: 10.1021/acs.inorgchem.7b00156 Inorg. Chem. 2017, 56, 4141−4147