High CO2 Uptake Capacity and Selectivity in a Fascinating Nanotube

Dec 30, 2016 - Although lots of previously reported cases of the porous MOFs have been explored, it could be a more interesting challenge to build por...
2 downloads 7 Views 3MB Size
Article pubs.acs.org/IC

High CO2 Uptake Capacity and Selectivity in a Fascinating NanotubeBased Metal−Organic Framework Yun-Long Wu,† Jinjie Qian,‡ Guo-Ping Yang,*,† Fan Yang,† Yu-Tong Liang,† Wen-Yan Zhang,† and Yao-Yu Wang*,† †

Key Laboratory of Synthetic and Natural Functional Molecule Chemistry of the Ministry of Education, Shaanxi Key Laboratory of Physico-Inorganic Chemistry, College of Chemistry & Materials Science, Northwest University, Xi’an 710127, China ‡ College of Chemistry & Materials Engineering, Wenzhou University, Wenzhou 325035, China S Supporting Information *

ABSTRACT: An unusual porous metal−organic framework has been synthesized by using Pb(II) and rigid V-shaped 4,4′-(pyridine-3,5-diyl)diisophthalic acid (H4L). Structure analysis reveals that there exist 1D cylindrical 14.26 Å and triangular prism 10.69 × 10.69 × 10.69 Å3 nanotubes in the framework. Gas sorption behavior of the nanoporous MOF shows a relatively high capacity and selectivity of CO2 over CH4.

1. INTRODUCTION Methane, as a kind of clean energy, is the major component of many combustible gases, such as natural gas, biogas, coke oven gas, landfill gas, straw syngas, etc.1 The large amount of greenhouse gas emission from the combustion of organic fuels, as one of the most intractable environmental issues, has presented burning issues for modern scientists. The emphasis of the research is focused on the following two aspects. On one hand, the separation of CO2 from fuel gases can improve the combustion efficiency of fuel gases,2 and on the other hand, as one of the greenhouse gases, the storage of CO2 may avoid its emission to the environment. Many separation technologies, such as cryogenic distillation, membrane separation, chemical scrubbing, physical adsorption, etc., have been used to separate CO2 from fuel gases.3 Attractively, because of diversities of structures such as high surface area, tunable functional pore size, functionalized pore walls, etc., the porous metal−organic frameworks (MOFs) based on metal ions and organic linkers, as a new subclass of solid state porous materials, has shone brilliantly on the aspect of gas adsorption and separation (CO2, CH4, H2, etc.) in recent years.4 Structures of porous MOFs are mainly depended on organic linkers. The rigid organic ligands with more coordinated atoms (O or N) have been widely investigated for their rigid predictable coordination models.5 In the latest reports, two porous MOFs based on the 1,3,5-tri(1H-1,2,3-triazol-5-yl) benzene, Fe(II) ions, and Co(II) ions in CO scavenging and selective, tunable O2 binding have been reported by Long’s group.6 The large family of Zr-based MOFs which exhibit rich © XXXX American Chemical Society

structure types and outstanding stability based on the various rigid organic linkers have been systemically explored in molecule adsorption and separation by Li and Zhou et al.7 The porous MOFs constructed by V-shaped rigid multicarboxylate ligands in CO2 adsorption and separation have also been widely explored. Although lots of previously reported cases of the porous MOFs have been explored, it could be a more interesting challenge to build porous MOFs due to their intrinsic higher stability and may act as the potential functional materials in various fields, especially for gas storage and separation.8 Our group has long been interested in designing and synthesizing the porous MOFs materials in preferential adsorption of CO2.9 As an effective combination of our previous research work, we herein report the synthesis, crystal structure, and gas sorption behavior of a 3D uncommon porous MOF 1-[Pb2(L)]·2DMF·2H2O materials based on a rigid Vshaped organic ligand (H4L = 4,4′-(pyridine-3,5-diyl)diisophthalic acid; DMF = N,N′-dimethylformamide). Structural analysis shows that there exist two kinds of 1D inerratic nanotubes (1D cylindrical 14.26 Å and triangular prism 10.69 × 10.69 × 10.69 Å3 nanotubes) in the framework, which has rarely been observed in 3D porous MOFs. The gas sorption of 1a shows a relatively high capacity and selectivity of CO2 over CH4. Received: October 19, 2016

A

DOI: 10.1021/acs.inorgchem.6b02491 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

2. EXPERIMENTAL SECTION

3. RESULTS AND DISCUSSION Description of Complex 1. Colorless crystals of 1 are obtained under solvothermal reaction. The phase purity of the bulk product has been confirmed by the powder X-ray diffraction (PXRD) analysis. The resultant formula of 1 is defined as [Pb2(L)]·2DMF·2H2O, which is calculated from the combination of the calculated TGA curves, element analysis (EA), and squeeze void count electrons after the PLATON/ SQUEEZE process (Figure S7).11 Single crystal X-ray diffraction data analysis reveals that 1 crystallized in the hexagonal system with P63cm space group. Both the Pb(II) ions and the ligand are at crystallographic special positions with the total site occupancies of 1 and 0.5, respectively. As illustrated in Figure 1, two kinds of Pb(II) are

Materials and Instruments. All the starting materials and reagents used in the experiment were commercially obtained and used without further purification. The element analyses of C, H, and N were performed through a Perkin-Elemer 2400C Elemental Analyzer. Infrared spectra were carried out with KBr pellets in the range of 4000−400 cm−1 on a Bruker EQUINOX-55 FT-IR spectrometer. Thermogravimetric analyses (TGA) were determined with a NETZSCH STA 449C microanalyzer under a nitrogen stream at a heating rate of 5 °C min−1. Powder X-ray diffraction data (PXRD) were collected on a Bruker D8 ADVANCE X-ray powder diffractometer with Cu−Kα radiation (1.5418 Å) at room temperature. The sorption isotherms were measured by a Micrometrics ASAP 2020M. Synthesis of [Pb2(L)]·2DMF·2H2O (1). A mixture of Pb(NO3)2 (0.20 mmol, 0.0662 g), H4L (0.10 mmol, 0.0407 g), DMF (4 mL), H2O (2 mL), and three drops of HNO3 (V[HNO3]:V[H2O] = 1:3) was stirred at room temperature for 30 min; then the mixture was placed in a 15 mL Teflon-lined stainless steel vessel and heated at 145 °C for 3 days. After that, the reactor was cooling to room temperature, and the colorless needle-like crystals were obtained. Yield 51% (based on H4L). Anal. Calcd: C, 32.43; H, 2.72; N, 4.2. Found: C, 32.10; H, 2.53; N, 4.31. FT-IR (KBr, cm−1): 3432 (s), 1663 (s), 1550 (s), 1376 (s), 1101 (m), 817 (s), 770 (m), 698 (m), 464 (m). Crystallographic Data Collection and Refinement. The crystal data of 1 were recorded on a Bruker SMART APEX II CCD detector equipped with graphite monochromated Mo−Kα radiation (λ = 0.71073 Å) by the ϕ/ω scan technique. The crystal structure of 1 was refined anisotropically on F2 by the SHELXTL program and modified for absorption by using SADABS.10 Non-hydrogen atoms were refined with anisotropic thermal parameters. All hydrogen atoms of organic ligands were placed in calculated at idealized positions and refined with a riding on attached atoms with isotropic thermal parameters. The solvent molecules were disordered and not located, and the SQUEEZE method was applied to remove the contributions of disordered guest molecules. The final formula was confirmed by combining singlecrystal structure, TGA data, element analyses, and squeeze void count electrons after the SQUEEZE process. More details on the crystallographic studies are given in the Supporting Information. Crystallographic data and selected bonds lengths/angles are displayed in Tables 1 and S1. The crystal data (CCDC No. 1495355) is deposited at the Cambridge Crystallographic Data Center.

Figure 1. Coordination environment of center Pb(II) ions. Symmetry codes: #1: y, x, 1 + z; #2: x, y, 1 + z; #3: −x + y, y, 0.5 + z; #4: 1 − y, 1 − x, 0.5 + z; #5: 1 − x, 1 − y, 0.5 + z; #6: x, y, z; #7: y, x, z.

six- and seven-coordinated geometric configurations, respectively. Pb1 is seven-connected by six O atoms of four carboxylate groups from four different L4− ligands and one N atom from the other L4− ligand, giving the distorted pentagonal bipyramid geometry. Pb2 adopts distorted pentagonal pyramid geometry by six O atoms of four different carboxylate groups from two independent L4− ligands. In 1, the carboxylate groups of the fully deprotonated L4− adopt one bridging tridentate (η2μ2χ3) (Figure S1) coordination models to bind with the Pb(II) ions, which produce a 1D chain motif (Figure 2c). Further, the C atoms of the carboxylate groups act as the extending nodes to sustain these 1D chains into a 3D porous structure (Figure 2d). Interestingly, there exist two 1D nanotubes (1D cylindrical with diameter 14.26 Å and triangular prism with sizes 10.69 × 10.69 × 10.69 Å3) (Figures 2a,b and S2), which is rarely reported. Topologically, the whole framework can be simplified as a new trinodal (3,4,7)coordinated topological net with the (43·63)(43)(46·612·83) point symbol (Figures 2f and S4). After the hypothetical removal of the noncoordinated and disordered solvent molecules, the potential solvent area volume of 1a-[Pb2(L)] is determined to be ∼50.5% (2596 Å3) out of the total volume of the unit cell (5143 Å3) by the PLATON program. To obtain the guest-free phase of 1a, 1 was soaked in CH3OH for 7 days and subsequent heating at 150 °C under vacuum for 6 h, which is evidenced by the TGA curve of 1a (Figure S5). The FT-IR of 1a (Figure S6) shows that the lack of characteristic CO vibration of DMF reveals that the

Table 1. Crystallographic Data and Structure Refinement for Complex 1

a

formula

C21H9NO8Pb2

formula weight crystal system space group a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) V (Å3) Z Dc/g cm−3 F(000) GOOF Rint a R1 [I > 2σ] b wR2 (all data)

817.67 hexagonal P63cm 21.346(7) 21.346(7) 13.033(4) 90 90 120 5143(4) 6 1.584 2220 1.009 0.1065 R1 = 0.0326 wR2 = 0.0699

R1 = ∑||Fo| − |Fc||/∑|Fo|. bwR2 = [∑w(Fo2 − Fc2)2/∑w(Fo2)2]1/2. B

DOI: 10.1021/acs.inorgchem.6b02491 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 2. (a) 1D cylindrical and (b) triangular prism nanotubes in 1. (c) 1D chain in 1. (d) A space-filling view of 1. (e) View of porous structure of 1. (f) Topological net of 1.

Figure 3. (a) The adsorption isotherms of 1a (CO2 and CH4) at 195 K. (b) The adsorption isotherms of CO2 and CH4 at 273 and 298 K.

different temperatures (CO2 and CH4 at 195, 273, and 298 K). On the basis of the CO2 adsorption isotherm, the Brunauer− Emmett−Teller (BET) and Langmuir surface areas of 1a are evaluated to be 669.3 and 800.5 m2 g−1, respectively. The experimental pore size distribution (PSD) curve based on the NLDFT model is shown in Figure S11. As shown in Figure 3a, the CO2 adsorption displays a type-I isotherm and reaches 198.3 cm3 g−1 (38.9 wt %) at 195 K. However, the CH4 uptake is up to 89.21 cm3 g−1 (6.4 wt %), which is much lower than the value of CO2 under the same conditions. Such adsorption selectivity is more significant at high temperature. Thus, the adsorption studies of the CO2 and CH4 were carried out at the high temperature (273 and 298 K). As shown in Figure 3b, the CO2 and CH4 uptake are 95.8 cm3

disordered solvent has been removed from the nanotubes. The structure integrity of 1a has also been verified by the PXRD experiment (Figure S8). Sorption Properties. The permanent porosity of 1a has been verified by the N2 adsorption isotherm at 77 K. The desolvated solid products exhibit a type-I isotherm (Figure S9), which confirms retention of microporosity upon removal of guest solvent molecules with a saturated N2 adsorption of 23 cm−3 g−1. H2 sorption measurements at 77 K have also been experimented to check the hydrogen storage performances (Figure S10). The special nanotube-based large space in 1a has been employed to give a detailed exploration of the potential application in CO2/CH4 separation, and adsorption studies of CO2 and CH4 have been investigated on desolvated 1a at C

DOI: 10.1021/acs.inorgchem.6b02491 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 4. IAST adsorption selectivity of 1a for equimolar mixture of CO2 and CH4 at 273 K (a) and 298 K (b).

g−1 (18.8 wt %) and 25.7 cm3 g−1 (1.8 wt %) at 273 K and 1 bar, 56.9 cm3 g−1 (11.2 wt %) and 10.5 cm3 g−1 (0.75 wt %) at 298 K and 1 bar, respectively. The CO2 uptake is nearly 4 and 5.4 times larger than that of CH4 at 273 and 298 K, respectively, which may be attributed to the great differences in diameters between the gas molecules (CO2 3.30 Å and CH4 3.80 Å) and the sieving effect of the narrow apertures.2c12 In addition, the CO2 uptake of 1a (18.8 and 11.2 wt %) at 273 and 298 K is comparable and even superior to the recently reported microporous complexes,13 displaying that the appropriate pore size may be more attractive for CO2 under low pressure. Due to the significant differences in the gas uptake between CO2 and CH4 at 273 and 298 K, the selectivity for CO2/CH4 of 1a was predicted by using the ideal adsorbed solution theory (ISAT) (Figure S12) at room temperature.14 As shown in Figure 4a,b, the predicated selectivities for CO2/CH4 (the component of the gas phase are set as 50/50) have been explored at 273 and 298 K, respectively. It is obvious to see that the high selectivity for CO2/CH4 can be obtained. The calculating selectivity of the CO2/CH4 largely depends on the temperature and pressure. The initial values of the selectivity for CO2/CH4 are 3.07 (298 K) and 26.3 (273 K) and the values arrive to 7 (298 K) and 118.8 (273 K) at 1 bar from an equimolar gas-phase mixture, which is better than most of the known MOFs under the same conditions and even comparable to the MOFs with high selectivity that are modified with open metal sites (OMSs) or polar groups (Table S2).15 The significant sorption behavior of 1 closely belongs to the potential solvent area volume with the appropriate pore size matching well with the kinetic diameters of CO2 and the open center metal sites produce the specific affinity for CO2.16 Importantly, the above studies may supply theoretical prediction in capturing CO2 and the highly selective uptake of CO2 from equimolar gas-phase mixtures, implying that 1a may be used in industrial application and acted as a potential candidate for postcombustion CO2 capture application from natural gas. The high selectivity for CO2 adsorption may be caused by the effect of the OMSs (the opened Pb(II) sites that are uncoordinated by H2O or organic molecule) located within the wall of the porous surface, and the OMSs may strengthen the electrostatic interactions between the porous surface and CO2 molecules. Moreover, the smaller kinetic diameter of CO2 (3.30 Å) compared to CH4 (3.80 Å) promotes CO2 molecules being more easily adsorbed by 1a. To further investigate the affinity of 1a for CO2, the adsorption heat of CO2 (Qst) has been

simulated by the virial equation based on the sorption isotherms at 273 and 298 K (Figure S13). The initial average Qst is 53.9 kJ mol−1, which is superior to those of reported MOFs with special active sites, such as OMSs and uncoordinated atoms (O, N) located on the surface of the porous complexes.3a,17 However, the values of Qst show a gradual decrease with the increase of the uptake of CO2. As the uptakes of CO2 attain to 56.8 cm3 g−1, the Qst reaches to 29.1 kJ mol−1 (Figure 5). The high Qst of CO2 is mainly ascribed to the

Figure 5. Isosteric heat of CO2 adsorption for 1a by the virial equation from the adsorption isotherms at 273 and 298 K.

opened Pb(II) sites (located within the porous surface) that make the frameworks with high polar enhance the framework− CO2 interactions, resulting in the high uptake capacity and selectivity for CO2.

4. CONCLUSIONS In conclusion, an unusual porous MOF 1 with two kinds of 1D inerratic nanotubes has been synthesized by using Pb(II) and rigid V-shaped 4,4′-(pyridine-3,5-diyl)diisophthalic acid (H4L). Gas adsorption experiments of 1a reveal that the high selectivity of CO2 from the equimolar mixtures of CO2 and CH 4 may make it act as a potential candidate for postcombustion CO2 capture application from natural gas. The present study may supply a facile route to design and synthesize the MOFs not only with novel structures but also with intriguing properties as promising functional materials. D

DOI: 10.1021/acs.inorgchem.6b02491 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry



frameworks. Nature 2015, 527, 503−507. (b) Li, J.-R.; Sculley, J.; Zhou, H.-C. Metal-Organic Frameworks for Separations. Chem. Rev. 2012, 112, 869−932. (c) Li, N.; Xu, J.; Feng, R.; Hu, T.-L.; Bu, X.-H. Governing metal−organic frameworks towards high stability. Chem. Commun. 2016, 52, 8501−8513. (d) Zhao, J.; Dong, W.-W.; Wu, Y.-P; Wang, Y.-N.; Wang, C.; Li, D.-S.; Zhang, Q.-C. Two (3,6)-connected porous metal-organic frameworks based on linear trinuclear [Co3(COO)6] and paddlewheel dinuclear [Cu2(COO)4] SBUs: gas adsorption, photocatalytic behaviour, and magnetic properties. J. Mater. Chem. A 2015, 3, 6962−6969. (e) Lu, X.-Q.; Jin, D.-L.; Wei, S.X.; Wang, Z.-J.; An, C.-H.; Guo, W.-Y. Strategies to enhance CO2 capture and separation based on engineering absorbent materials. J. Mater. Chem. A 2015, 3, 12118−12132. (5) (a) Li, M.; Li, D.; O’Keeffe, M.; Yaghi, O. M. Topological Analysis of Metal−Organic Frameworks with Polytopic Linkers and/or Multiple Building Units and the Minimal Transitivity Principle. Chem. Rev. 2014, 114, 1343−1370. (b) Jiang, J.-C.; Zhao, Y.-B.; Yaghi, O. M. Covalent Chemistry beyond Molecules. J. Am. Chem. Soc. 2016, 138, 3255−3265. (c) Zhao, J.; Wang, Y.-N.; Dong, W.-W.; Wu, Y.-P.; Li, D.-S.; Zhang, Q.-C. A Robust Luminescent Tb(III)-MOF with Lewis Basic Pyridyl Sites for the Highly Sensitive Detection of Metal Ions and Small Molecules. Inorg. Chem. 2016, 55, 3265−3271. (d) O'Keeffe, M.; Yaghi, O. M. Deconstructing the crystal structures of metalorganic frameworks and related materials into their underlying nets. Chem. Rev. 2012, 112, 675−702. (e) Zhao, J.; Wang, Y.; Dong, W.; Wu, Y.; Li, D.; Liu, B.; Zhang, Q. A new surfactant-introduction strategy for separating the pure single-phase of metal−organic frameworks. Chem. Commun. 2015, 51, 9479−9482. (6) (a) Reed, D. A.; Xiao, D. J.; Gonzalez, M. I.; Darago, L. E.; Herm, Z. R.; Grandjean, F.; Long, J. R. Reversible CO Scavenging via Adsorbate-Dependent Spin State Transitions in an Iron(II)− Triazolate Metal−Organic Framework. J. Am. Chem. Soc. 2016, 138, 5594−5602. (b) Xiao, D. J.; Gonzalez, M. I.; Darago, L. E.; Vogiatzis, K. D.; Haldoupis, E.; Gagliardi, L.; Long, J. R. Selective, Tunable O2 Binding in Cobalt(II)−Triazolate/Pyrazolate Metal−Organic Frameworks. J. Am. Chem. Soc. 2016, 138, 7161−7170. (7) Bai, Y.; Dou, Y.-B; Xie, L.-H.; Rutledge, W.; Li, J.-R.; Zhou, H.-C. Zr-based metal-organic frameworks: design, synthesis, structure, and applications. Chem. Soc. Rev. 2016, 45, 2327−2367. (8) (a) Alezi, D.; Belmabkhout, Y.; Suyetin, M.; Bhatt, P. M.; Weselinski, Ł. J.; Solovyeva, V.; Adil, K.; Spanopoulos, I.; Trikalitis, P. N.; Emwas, A.-H.; Eddaoudi, M. MOF Crystal Chemistry Paving the Way to Gas Storage Needs: Aluminum-Based soc-MOF for CH4, O2, and CO2 Storage. J. Am. Chem. Soc. 2015, 137, 13308−13318. (b) Wang, F.; Fu, H.-R.; Kang, Y.; Zhang, J. A new approach towards zeolitic tetrazolate-imidazolate frameworks (ZTIFs) with uncoordinated N-heteroatom sites for high CO2 uptake. Chem. Commun. 2014, 50, 12065−12068. (c) Li, Y.-W.; Li, J.-R.; Wang, L.-F.; Zhou, B.-Y.; Chen, Q.; Bu, X.-H. Microporous metal−organic frameworks with open metal sites as sorbents for selective gas adsorption and fluorescence sensors for metal ions. J. Mater. Chem. A 2013, 1, 495− 499. (d) Ling, Y.; Jiao, J.; Zhang, M.; Liu, H.; Bai, D.; Feng, Y.; He, Y. A porous lanthanide metal−organic framework based on a flexible cyclotriphosphazenefunctionalized hexacarboxylate exhibiting selective gas adsorption. CrystEngComm 2016, 18, 6254−6261. (9) (a) Li, J.; Yang, G.-P.; Wei, S.-L.; Gao, R.-C.; Bai, N.-N.; Wang, Y.-Y. Two Microporous Metal-Organic Frameworks with Suitable Pore Size Displaying the High CO2/CH4 Selectivity. Cryst. Growth Des. 2015, 15, 5382−5387. (b) Liu, B.; Wu, W.-P.; Hou, L.; Wang, Y.-Y. Four uncommon nanocage-based Ln-MOFs: highly selective luminescent sensing for Cu2+ ions and selective CO2 capture. Chem. Commun. 2014, 50, 8731−8734. (10) Sheldrick, G. M. SADABS: Empirical Absorption Correction Program; University of Göttingen: Göttingen, Germany, 1997. (11) Spek, A. L. Single-crystal structure validation with the program PLATON. J. Appl. Crystallogr. 2003, 36, 7−13. (12) Kong, L.; Zou, R.; Bi, W.; Zhong, R.; Mu, W.; Liu, J.; Han, R. P. S.; Zou, R. Selective adsorption of CO2/CH4 and CO2/N2 within a

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.6b02491. Additional tables, selected bond lengths and angles, structure figures, FT-IR, TGA, PXRD patterns, and the detailed calculations of sorption (PDF) X-ray crystallographic data for 1 (CIF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (G.-P.Y.). *E-mail: [email protected] (Y.-Y.W.). ORCID

Yao-Yu Wang: 0000-0002-0800-7093 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We thank the NSFC (Grant Nos. 21201139, 21371142, 21531007, and 21601137), the China Postdoctoral Science Foundation (Grant No. 2016M600807), and the Postdoctoral Science Foundation of Northwest University (Grant No. 334100049) for financial support and Dr. Feng-Feng Wang for crystallographic data refinement (Beijing City Key laboratory of Polymorphic Drugs, Center of Pharmaceutical Polymorphs, Institute of Materia Medica, Chinese Academy of Medical Sciences and Peking Union Medical College, Beijing 100050, China).



REFERENCES

(1) (a) Zhang, Z.-J.; Yao, Z.-Z.; Xiang, S.-C.; Chen, B.-L. Perspective of microporous metal-organic frameworks for CO2 capture and separation. Energy Environ. Sci. 2014, 7, 2868−2899. (b) Oh, Y.; Le, V.-D.; Maiti, U. N.; Hwang, J. O.; Park, W. J.; Lim, J.; Lee, K. E.; Bae, Y.-S.; Kim, Y.-H.; Kim, S.-O. Selective and Regenerative Carbon Dioxide Capture by Highly Polarizing Porous Carbon Nitride. ACS Nano 2015, 9, 9148−9157. (c) Qian, J.-J.; Jiang, F. L.; Zhang, L. J.; Su, K. Z.; Pan, J.; Li, Q. P.; Yuan, D. Q.; Hong, M.-C. Unusual pore structure and sorption behaviour in a hexanodal zinc−organic framework material. Chem. Commun. 2014, 50, 1678−1681. (2) (a) He, Y.-B.; Zhou, W.; Qian, G.-D.; Chen, B.-L. Methane storage in Metal-Organic Frameworks. Chem. Soc. Rev. 2014, 43, 5657−5678. (b) Zhou, H.-C.; Long, J. R.; Yaghi, O. M. Introduction to Metal−Organic Frameworks. Chem. Rev. 2012, 112, 673−674. (c) Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm, Z. R.; Bae, T.-H.; Long, J. R. Carbon Dioxide Capture in Metal-Organic Frameworks. Chem. Rev. 2012, 112, 724− 781. (3) (a) Zhang, P.; Tong, J.-J.; Jee, Y.; Huang, K. Stabilizing a hightemperature electrochemical silver-carbonate CO2 capture membrane by atomic layer deposition of a ZrO2 overcoat. Chem. Commun. 2016, 52, 9817−9820. (b) Park, J.; Yuan, D.; Pham, K. T.; Li, J.-R.; Yakovenko, A.; Zhou, H.-C. Reversible Alteration of CO2 Adsorption upon Photochemical or Thermal Treatment in a Metal-Organic Framework. J. Am. Chem. Soc. 2012, 134, 99−102. (c) Chen, Y.-Q.; Qu, Y.-K.; Li, G.-R.; Zhuang, Z.-Z.; Chang, Z.; Hu, T.-L.; Xu, J.; Bu, X.H. Zn(II)-Benzotriazolate Clusters Based Amide Functionalized Porous Coordination Polymers with High CO2 Adsorption Selectivity. Inorg. Chem. 2014, 53, 8842−8844. (4) (a) Cho, H. S.; Deng, H.-X.; Miyasaka, K.; Dong, Z.-Y.; Cho, M.; Neimark, A. V.; Kang, J. K.; Yaghi, O. M.; Terasaki, O. Extra adsorption and adsorbate superlattice formation in metal-organic E

DOI: 10.1021/acs.inorgchem.6b02491 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry charged metal-organic framework. J. Mater. Chem. A 2014, 2, 17771− 17778. (13) (a) Li, L.-J.; Liao, P.-Q.; He, C.-T.; Wei, Y.-S.; Zhou, H.-L.; Lin, J.-M.; Li, X.-Y.; Zhang, J.-P. Grafting alkylamine in UiO-66 by chargeassisted coordination bonds for carbon dioxide capture from highhumidity flue gas. J. Mater. Chem. A 2015, 3, 21849−21855. (b) Millward, A. R.; Yaghi, O. M. Metal-Organic Frameworks with Exceptionally High Capacity for Storage of Carbon Dioxide at Room Temperature. J. Am. Chem. Soc. 2005, 127, 17998−17999. (c) Cui, P.; Ma, Y.-G.; Li, H.-H.; Zhao, B.; Li, J.-R.; Cheng, P.; Balbuena, P. B.; Zhou, H.-C. Multipoint Interactions Enhanced CO2 Uptake: A Zeolite-like Zinc-Tetrazole Framework with 24-Nuclear Zinc Cages. J. Am. Chem. Soc. 2012, 134, 18892−18895. (14) Myers, A. L.; Prausnitz, J. M. Thermodynamics of mixed-gas adsorption. AIChE J. 1965, 11, 121−127. (15) (a) Prasad, T. K.; Suh, M. P. Control of Interpenetration and Gas-Sorption Properties of Metal-Organic Frameworks by a Simple Change in Ligand Design. Chem. - Eur. J. 2012, 18, 8673−8680. (b) Qian, J.; Jiang, F.; Su, K.; Pan, J.; Zhang, L.; Li, X.; Yuan, D.; Hong, M. Sorption behaviour in a unique 3,12-connected zinc-organic framework with 2.4 nm cages. J. Mater. Chem. A 2013, 1, 10631− 10634. (c) Wang, D.; Zhao, T.; Cao, Y.; Yao, S.; Li, G.; Huo, Q.; Liu, Y. High performance gas adsorption and separation of natural gas in two microporous metal−organic frameworks with ternary building units. Chem. Commun. 2014, 50, 8648−8650. (16) (a) Zhao, J.; Dong, W.-W.; Wu, Y.-P.; Wang, Y.-N.; Wang, C.; Li, D.-S.; Zhang, Q.-C. Two (3,6)-connected porous metal-organic frameworks based on linear trinuclear [Co3(COO)6] and paddlewheel dinuclear [Cu2(COO)4] SBUs: gas adsorption, photocatalytic behaviour, and magnetic properties. J. Mater. Chem. A 2015, 3, 6962−6969. (b) Zhao, D.; Timmons, D. J.; Yuan, D.; Zhou, H.-C. Tuning the Topology and Functionality of Metal-Organic Frameworks by Ligand Design. Acc. Chem. Res. 2011, 44, 123−133. (17) (a) Tang, K.; Yun, R.; Lu, Z.; Du, L.; Zhang, M.; Wang, Q.; Liu, H. High CO2/N2 Selectivity and H2 Adsorption of a Novel Porous Yttrium Metal-Organic Framework Based on N, N′,N″-Tris(isophthalyl)-1,3,5-benzenetricarboxamide. Cryst. Growth Des. 2013, 13, 1382−1385. (b) Wang, B.; Huang, H.; Lv, X.-L.; Xie, Y.; Li, M.; Li, J.-R. Tuning CO2 Selective Adsorption over N2 and CH4 in UiO-67 Analogues through Ligand Functionalization. Inorg. Chem. 2014, 53, 9254−9259.

F

DOI: 10.1021/acs.inorgchem.6b02491 Inorg. Chem. XXXX, XXX, XXX−XXX