A Microporous Co-MOF for Highly Selective CO2 ... - ACS Publications

May 14, 2019 - ... XXX, XXX−XXX. © XXXX American Chemical Society ... 22 mg of 2,6-NDC, 21 mg of L, and 5 mL of DMF were taken in a 15. mL Teflon-l...
0 downloads 0 Views 3MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

A Microporous Co-MOF for Highly Selective CO2 Sorption in High Loadings Involving Aryl C−H···OCO Interactions: Combined Simulation and Breakthrough Studies Arun Pal,† Santanu Chand,† David G. Madden,‡ Douglas Franz,§ Logan Ritter,§ Alexis Johnson,§ Brian Space,§ Teresa Curtin,‡,∥ and Madhab C. Das*,† †

Department of Chemistry, Indian Institute of Technology Kharagpur, Kharagpur-721302, WB, India Bernal Institute, University of Limerick, V94 T9PX, Limerick, Ireland § Department of Chemistry, University of South Florida, 4202 E. Fowler Ave., CHE205, Tampa, Florida 33620-5250, United States ∥ Chemical Sciences Department, University of Limerick, V94 T9PX, Limerick, Ireland

Downloaded via NOTTINGHAM TRENT UNIV on August 9, 2019 at 22:05:33 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: In the context of porous crystalline materials toward CO2 separation and capture, a new 2-fold interpenetrated 3D microporous Co-MOF, IITKGP-11 (IITKGP denotes Indian Institute of Technology Kharagpur), has been synthesized consisting of a 1D channel of ∼3.6 × 5.0 Å2 along the [101] direction with a cavity volume of 35.20%. This microporous framework with a BET surface area of 253 m2g−1 shows higher uptake of CO2 (under 1 bar, 3.35 and 2.70 mmol g−1 at 273 and 295 K, respectively), with high separation selectivities for CO2/N2 and CO2/CH4 gas mixtures under ambient conditions as estimated through IAST calculation. Moreover, real time dynamic breakthrough studies reveal the high adsorption selectivity toward CO2 for these binary mixed gases at 295 K and 1 bar. Besides high gas separation selectivity, capacity considerations in mixed gas phases are also important to check the performance of a given adsorbent. CO2 loading amounts in mixed gas phases are quite high as predicted through IAST calculation and experimentally determined from dynamic breakthrough studies. In order to get insight into the phenomena, GCMC simulation was performed demonstrating that the CO2 molecules are electrostatically trapped via interactions between oxygen on CO2 and hydrogen on pyridyl moieties of the spacers.



backbone of the framework for higher CO2 affinity.8,9 Similarly, the binding affinity of the frameworks toward CO2 molecules can be dramatically enhanced by the generation of unsaturated metal sites.9a,c However, maintaining the high separation selectivity and large storage capacity for CO2 adsorption is still an enormous challenge for synthetic porous materials. The strength of the framework−CO2 interactions is generally evaluated by the isosteric heats of adsorption (Qst) calculated from the single-component gas sorption isotherm data. Besides direct identification of CO2 crystallographically, GCMC simulations are performed to get insight into the CO2 sorption phenomena, identifying the plausible binding sites of CO2 with framework backbones. Often, CO2 molecules form quadrupole−quadrupole interactions among themselves.3a,10 In addition, N−H···O(CO2) and O−H···O(CO2) interactions are also responsible for high uptake capacity as demonstrated by the seminal works of Shimizu et al. and Chen et al., respectively.9b,11 However, electrostatic interaction involving

INTRODUCTION

It is widely accepted that the reason for the unusual abrupt rise of global temperature is due to the existence of increased greenhouse gas levels in the atmosphere, and carbon dioxide itself has the lion’s share of 65% among the greenhouse gases.1 As an alternative approach to the commonly used alkanolamine solutions for CO2 separation through chemisorption, porous solids are currently investigated as potential materials for an effective solution to the carbon capture challenge through physisorption.2 Metal−organic frameworks (MOFs)3 consisting of metal ions or clusters and organic linkers having permanent porosity, high surface area, tunable characters with useful thermal and chemical stability are investigated for their potential applications in diverse areas including gas adsorption and separation.4−7 For a MOF which is capable of superior carbon capture, it must have suitable cavity size, polarity, shape, and functionality for sorption of CO2. Selective CO2 sorption from its gas mixtures could be accomplished by modifications of the framework such as tuning the pore size through interpenetrations or varying the length of the organic ligands and/or immobilizing special functional groups on the © XXXX American Chemical Society

Received: May 14, 2019

A

DOI: 10.1021/acs.inorgchem.9b01402 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. Two crystallographically independent Co(II) centers in IITKGP-11 with surrounding environment (a). A representation of the overall framework showing a channel size of ∼3.6 × 5.0 Å2 along the [101] direction (b).

Figure 2. Gas sorption isotherms for single component CO2, CH4, and N2 on activated IITKGP-11 at 273 K (a) and at 295 K (b; solid circles, adsorption data points; open circles, desorption data points).

relevant for industrial processes exhibiting high separation selectively of CO2 at 295 K and 1 bar. CO2 loading amounts in mixed gas phases are quite high as estimated through IAST calculation and experimentally validated through the dynamic breakthrough study. GCMC simulations show that the adsorbed CO2 molecules are electrostatically trapped through interactions between oxygen on CO2 and hydrogen on the pyridyl moieties.

aryl C−H and adsorbed CO2 molecules (C−H···OCO) are rarely reported.12 The gas mixture selectivities are predicted based on the wellknown ideal adsorbed solution theory (IAST) proposed by Myers and Prausnitz for the mixed gases within the desired pressure range, whereas the separation performances are practically verified by dynamic breakthrough experiments.13 Although there are considerable reports on IAST selectivity for MOF-based gas separations, reports on selectivity both predicted by IAST and validated through dynamic breakthrough experiments are scarcer. In addition, high gas separation selectivity, consideration of desired gas uptake capacity from its mixed gas phases, is also necessary to evaluate the performance of a particular adsorbent. It is thus important to evaluate the loading amounts of CO2 both by the theoretical calculations through the IAST method and experimentally as well through breakthrough study. Considering these aspects, we report herein a new 2-fold interpenetrated 3D microporous MOF, {[Co 2 (2,6NDC)2(L)2]·xG}n, IITKGP-11 (IITKGP denotes Indian Institute of Technology Kharagpur; G represents disordered solvents in pores) based on a rigid organic linker (H2NDC = 2,6-naphthalenedicarboxylic acid), a linear spacer L (1,4-bis(4pyridyl)-2, 3-diaza-1, 3-butadiene), and a Co(II) ion. The gas sorption study on activated IITKGP-11 shows higher uptake of CO2 with high IAST selectivities for CO2/N2 (flue gas mixture, 15:85) and CO2/CH4 (landfill gas mixture, 50:50) gas mixtures under ambient conditions. Moreover, we performed real time dynamic breakthrough experiments under conditions



EXPERIMENTAL SECTION

Materials. All solvents and chemicals were used as received. Spacer 1,4-bis(4-pyridyl)-2,3-diaza-1,3-butadiene (L) was synthesized according to our earlier report.14 {[Co2(2,6-NDC)2(L)2]·xG}n, IITKGP-11. 29 mg of Co(NO3)2.6H2O, 22 mg of 2,6-NDC, 21 mg of L, and 5 mL of DMF were taken in a 15 mL Teflon-lined autoclave followed by 10 min of sonication to get a homogeneous solution. The autoclave was then sealed into a stainless steel jacket and heated at 90 °C for 48 h, which was then cooled down to room temperature for 12 h. Reddish needle like crystals were obtained in 70% yield based on the metal. IR (cm−1): 3325.8(b), 1611.5(s), 1547.5(s), 1487.8(s), 1423.7(w), 1383.0(s), 1351.0(s), 1311.3(m), 1230.7(s), 1183.2(s), 1090.6(s), 1055(m), 1018.8(s), 950.3(s), 794.6(s), 686.4(s), 522.0(s).



RESULTS AND DISCUSSION IITKGP-11 crystallizes in a monoclinic crystal system with the centrosymmetric space group P21/n, and the asymmetric unit consists of two Co(II) ions, two 2,6-NDC2− ligands, and two spacers L. Figure 1a represents the surrounding environment of Co(II) centers where both the Co(II) ions acquire distorted B

DOI: 10.1021/acs.inorgchem.9b01402 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 3. Gas sorption selectivities and binary mixture gas sorption isotherms (a, b) at 273 K and (c, d) at 295 K estimated by IAST for CO2/N2 (15:85) and CO2/CH4 (50:50).

Figure 4. Breakthrough curves for separation of (a) flue gas mixture (15% CO2/85% N2) and (b) landfill gas mixture (50% CO2/50% CH4) at 1 bar and 295 K.

whereas both carboxylates of another independent 2,6-NDC2− unit display a μ2-η1:η1 bidentate bridging mode forming the {Co2(COO)2} unit. These units are connected through both the ligands to form an overall 2-fold interpenetrated 3D framework (Figure S1a, Supporting Information). The topological analysis15 shows that the structure of this MOF can be understood as a sqc104 topology with a point symbol of

octahedral configuration, the equatorial sites are surrounded by four O atoms from three 2,6-NDC2− ligands, and the axial sites are ligated by two N atoms from two spacers L. Co1 and Co2 centers are linked via two carboxylate groups of two 2,6NDC2− ligands in a bridging fashion to make a {Co2(COO)2} unit with a Co1···Co2 distance of 4.149 Å. Both carboxylates of one 2,6-NDC2− ligand show a bidentate chelating mode, C

DOI: 10.1021/acs.inorgchem.9b01402 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

253 m2g−1 (Figure S6, Supporting Information). The BET surface area was also calculated to be 238 m2 g−1 from CO2 sorption measurement at 195 K (Figure S7, Supporting Information). Horvath−Kawazoe (HK) differential pore volume vs pore width plot obtained from 195 K CO2 sorption analysis reveals that this MOF contains pores in the range of 3.47−5.12 Å with median pore width of 3.77 Å, which is fairly consistent with the single crystal structural analysis (Figure S8, Supporting Information). Single-component gas sorption isotherms of CO2, CH4, and N2 performed at 273 and 295 K are presented in Figure 2. Under 1 bar, CO2 uptake capacity at 273 and 295 K is 75.10 cm3 g−1 (3.35 mmol g−1) and 60.50 cm3 g−1 (2.70 mmol g−1), respectively. Most importantly, it shows remarkably high uptake of CO2 at 0.15 bar, similar to the CO2 partial pressure in flue gas mixtures (1.77 mmol g−1 at 273 K, 1.10 mmol g−1 at 295 K) and at 0.5 bar, similar to the CO2 partial pressure in landfill gas mixtures (2.81 mmol g−1 at 273 K, 2.06 mmol g−1 at 295 K) indicating IITKGP-11 as an efficient material for mixed gas adsorption-based separation studies. On the contrary, it (under 1 bar) takes up a comparatively lesser amount of CH4, 34.20 cm3 g−1 (1.53 mmol g−1) at 273 K and 23.10 cm3 g−1 (1.03 mmol g−1) at 295 K, and N2, 9.80 cm3 g−1 (0.44 mmol g−1) at 273 K and 4.50 cm3 g−1 (0.20 mmol g−1) at 295 K. Hydrogen uptake was found to be 75 cm3 g−1 (0.67 wt %) from the sorption isotherm at 77 K under 1 bar (Figure S9, Supporting Information). To explore the affinity of the activated framework of IITKGP-11 toward adsorbed gas molecules, the isosteric heat of adsorption (Qst) was evaluated using the Clausius− Clapeyron equation using their single component gas sorption isotherms at 273 and 295 K (Figure S11, Supporting Information). The Qst values at near zero coverage of CO2, CH4, and N2 are 25, 13, and 8 kJ mol−1, respectively, which shows the gas binding affinity in the descending order of CO2> CH4> N2. Qst of CO2 is in a similar range to those of some of the well-known MOFs: Cu24(TPBTM)8 (26.3 kJ mol−1),17a NOTT-125 (25.35 kJ mol−1),17b NOTT-140 (24.7 kJ mol−1),17c HHU-5 (25.6 kJ mol−1),17b IITKGP-5 (22.6 kJ mol−1),17d SIFSIX-2-Cu (22 kJ mol−1),17e CuBTTri (21 kJ mol−1),17f UTSA-25a (22.2 kJ mol−1),11 UTSA-34b (25.4 kJ

Table 1. Selectivity Data and CO2 Loading Amount in Mixed Gas Phase from Both IAST Calculations and Breakthrough Experiments at 295 K under 1 Bar gas mixture at 295 K CO2/N2 (15:85) CO2/ CH4 (50:50)

IAST selectivity

selectivity from breakthrough study

CO2 loading in mixed gas phase from IAST calculation (mmol g−1)

CO2 loading in mixed gas phase from breakthrough study (mmol g−1)

149.07

117

1.02

0.92

7.91

39

1.92

1.73

{36.418.53.6} (Figure S1a, Supporting Information). Figure 1b represents the packing diagram showing a 1D channel of ∼3.6 × 5.0 Å2 along the [101] direction (including van der Waals radii). A careful investigation of the structure reveals the presence of two π···π stacking interactions between terminal pyridyl moieties of spacer L with a centroid-to-centroid distance of 4.104 and 4.245 Å (Figure S1b, Supporting Information). The overall 3D network is further stabilized by several noncovalent interactions (C−H···N and C−H···O) with 2,6-NDC2− ligands and spacer L (Table S3, Supporting Information). Guest accessible void volume was assessed to be 35.17% of the unit cell volume (2030.4 Å3 per unit cell volume of 5773.4 Å3) from PLATON analysis.16 As-synthesized PXRD data could be well matched with the simulated pattern, which confirmed the bulk phase purity of the sample (Figure S2, Supporting Information). The TGA study shows that the first weight loss occurs at 205 °C, which corresponds to the loss of lattice solvents. After that, the material is stable up to 330 °C followed by destruction of the structure upon continued heating (Figure S5, Supporting Information). To verify the permanent porosity of the sample, gas sorption measurements were conducted with the activated sample. The as-synthesized sample was immersed in dry CHCl3 for 2 days and subsequently degassed at 333 K under a vacuum to get the activated framework. PXRD analysis revealed that the activated sample maintained its original crystallinity (Figure S3, Supporting Information). It takes up to 99 cm3 g−1 of N2 at 77 K under 1 bar of pressure showing a BET surface area of

Figure 5. (a) Multiple cycles of CO2 adsorption isotherms for IITKGP-11 at 295 K. (b) Multiple cycles of breakthrough experiments for CO2/N2 (15:85) and CO2/CH4 (50:50) gas mixtures. D

DOI: 10.1021/acs.inorgchem.9b01402 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 6. Interaction between CO2 oxygen and H atoms in IITKGP-11 for the binding site, showing internuclear distances. Co1, orange; Co2, cyan; C, gray; H, white; N, blue; O, red.

mol−1),11 UTSA-90a (20.5 kJ mol−1),17g IITKGP-6 (23 kJ mol−1),14 ZJNU-54 (24.7 kJ mol−1),17b and HHU-3 (24.6 kJ mol−1).17b In order to investigate the separation ability for CO2, the gas selectivities of CO2/N2 (15:85) and CO2/CH4 (50:50) were calculated via dual-site IAST, a common route for calculation of binary mixture gas sorption based on experimental single component gas sorption isotherms.13 The IAST calculation for the estimation of loadings of the components for numerous binary gas mixtures gives a more precise value in a wide range of MOFs and zeolites than typical configurational-bias MC simulations of mixture sorption studies.18 Single component isotherm data were fitted with excellent correlation coefficients employing the dual-site Langmuir−Freundlich equation (DSFL; Figures S12−S17, Supporting Information). Then, the multicomponent gas sorption isotherms with IAST were calculated with the help of DSLF fitting parameters (Table S4, Supporting Information). The results of IAST calculations are presented in Figure 3 and Table S5, Supporting Information. Under 1 bar, the IAST selectivity of this MOF for the CO2/N2 mixture is 52.27 at 273 K and 149.07 at 295 K; for the CO2/ CH4 mixture, it is 9.63 at 273 K and 7.91 at 295 K (Figure 3). CO2/N2 selectivity keeps rising at 295 K, and the exact reason is still not clear. However, similar phenomena were also observed by Chen et al. and us.19 These IAST selectivity values are comparable and/or higher than those of previously reported well-known MOFs: SIFSIX-2-Cu, MOF-177, CuBTTri, en-Cu-BTTri, NOTT-202a, HKUST-1, TIFSIX-1-Cu, UTSA-33a, UTSA-72a, Zn-MOF-74, PCN-88, ZIFs (ZIF-70, 81, 95, 100), zeolites (MFI, 13X), and activated carbons (ACs; Table S5, Supporting Information).11,17e,f,20 This result indicates the ability of IITKGP-11 to act as a physical absorbent for flue gas and landfill gas mixture separation under ambient conditions.

In addition to gas separation selectivity, in a pressure-swing adsorption (PSA) unit, the capacity of the absorbed amount of the desired gas from its gaseous mixture is also a very important parameter to justify the performance of any given adsorbent. The CO2 loading, as calculated from the IAST method under the equilibrium conditions of the binary gas mixtures of CO2/N2 and CO2/CH4 at 273 and 295 K up to 100 kPa, is depicted in Figure 3. The CO2 loading amounts from the binary gas mixture of CO2/N2 (15:85) are 1.67 and 1.02 mmol g−1 at 273 and 295 K, respectively, under 100 kPa. The loading amount of 1.02 mmol g−1 at 295 K is much higher than for MOF-177 (0.16 mmol g−1 at 296 K, 100 kPa), zeolite MFI (0.26 mmol g−1 at 296 K, 100 kPa) and ZIF-78 (0.76 mmol g−1 at 296 K, 100 kPa).11 For the case of a binary gas mixture of CO2/CH4 (50:50), these values are 2.62 and 1.92 mmol g−1 at 273 and 295 K, respectively, which are comparable to IITKGP-5 (2.72 and 1.83 mmol g−1 at 273 and 295 K, 100 kPa)17d but higher than IITKGP-6 (1.54 and 0.77 mmol g−1 at 273 and 295 K, 100 kPa)14 and IITKGP-8 (0.55 mmol g−1 at 295 K, 100 kPa).21 In order to acquire an experimental separation selectivity, real time dynamic breakthrough experiments were performed on an activated sample at 295 K at 1 bar. Before the breakthrough experiment, the adsorbent bed was purged under He gas with a flow rate of 25 mLmin−1 at 60 °C for 3 h (see SI for details). Measurements of breakthrough curves of a 15:85 (v/v) mixture of CO2/N2 (flue gas) and a 50/50 (v/v) mixture of CO2/CH4 (landfill gas) passed throughout the chromatographic column packed with activated sample (Figure S18, Supporting Information) at 1 bar and 295 K revealed the selective retention of CO2 and the passing of N2 and CH4 through this material (Figure 4). CO2 breakthrough takes place approximately at 64 and 37 min after dosing the gas mixtures CO2/N2 and CO2/CH4, respectively, while N2 and CH4 breakthrough take place within a few seconds, indicating E

DOI: 10.1021/acs.inorgchem.9b01402 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

calculated to be −32.80 kJ mol−1, which is in good agreement with the experimental Qst obtained from the Clausius− Clapeyron equation (Figure S25, Supporting Information).

high selectivity toward CO2. The increase in CH4 concentration at ca. 30 min can be attributed to roll-up caused by either the increase of the column temperature due to the exothermic adsorption of CO2 or the dead volume of the system.22 CO2 breakthrough time for CO2/N2 (15:85) and CO2/CH4 (50:50) gas mixtures is higher than that of benchmark MOFs, SIFSIX-2-Cu-i (∼300 s for CO2/N2 (10:90) and ∼200 s for CO2/CH4 (50:50)),17e SIFSIX-3-Zn (∼2000 s for CO2/N2 (10:90) and ∼380 s for CO2/CH4 (50:50)),17e and ZIFs such as ZIF-95 (∼200 s for CO2/N2 (50:50) and ∼280 s for CO2/CH4 (50:50))23 and ZIF-100 (∼150 s for CO2/N2 (50:50) and ∼135 s for CO2/CH4 (50:50)).23 Around 0.92 mmol g−1 of CO2 for the CO2/N2 mixture and 1.73 mmol g−1 of CO2 for the CO2/CH4 mixture were adsorbed by the activated sample under these dynamic conditions. However, these values are 1.02 and 1.92 mmol g−1 for CO2/N2 and CO2/CH4, respectively, at 295 K as obtained from IAST calculations (Table 1). To check the durability of the activated sample IITKGP-11 for practical application, we have performed single component multiple cycle CO2 adsorption isotherms, which showed that the CO2 uptake capacity remained the same over five cycles (Figure 5a). The material retained its crystallinity as confirmed by PXRD measurement (Figure S4, Supporting Information). The adsorption efficiency of this MOF was also validated by multiple breakthrough cyclic experiments under each gas flow showing its consistent CO2 uptake capacity from both the mixed gas phases (Figure 5b). The result clearly indicates retention in adsorption capacity over four cycles and revealed the complete restoration of MOF just by degassing at room temperature. SEM images revealed that the morphology of the sample remains the same after multiple CO2 adsorption cycles (Figure S10, Supporting Information). Grand Canonical Monte Carlo (GCMC) simulations were performed in order to get insight into the interactions of MOF with adsorbed CO2 molecules at the molecular level (see SI). The strongest binding for CO2 occurs near Co1, which by our calculations exhibits the highest partial charge (and thus lowest electron density), and there is a low-energy sorption region between naphthalene linkers in the crystal. At low pressure and room temperature (Figures S20 and S21), CO2 loads into square linear channels. The most accessible and occupied regions occur on the sans-N linker between the naphthalene groups, with an indication of π-stacking interactions (Figure S21, Supporting Information). With higher pressure (Figure S22, Supporting Information), the region between the naphthalene groups becomes more accessible. A similar occurrence happens at higher temperature, suggesting the importance of dynamic effects near ambient temperatures. For simulated annealing calculations, a single gas molecule was inserted into the 2 × 2 × 1 supercell and annealed from 500 K to 70 K over a series of Monte Carlo translation/rotation trial moves to minimize the binding energy. The temperature was reduced by an exponential scale with a factor of 0.9995. The resulting binding site is shown in Figures 6 and Figure S23 (SI). CO2 becomes electrostatically trapped via interactions between oxygen on CO2 and hydrogen on the pyridyl moieties as shown in Figure 6. A total of eight such interactions (C1− H1···O(CO2), C2−H2···O(CO2), C4−H4···O(CO2), C5− H5···O(CO2), C37−H37···O(CO2), C38−H38···O(CO2), C40−H40···O(CO2), C41−H41···O(CO2)) with an average interaction distance of 3.91 Å are responsible for holding the CO2 molecule inside the channel. Binding energy was



CONCLUSIONS We have successfully synthesized a permanently microporous 2-fold interpenetrated 3D Co-MOF by employing mixed linkers approach possessing a 1D channel of ∼3.6 × 5.0 Å2 along the [101] direction. Single component gas sorption studies showed selective CO2 sorption over N2 and CH4. GCMC simulations on CO2 sorption revealed electrostatically trapped CO2 molecules via interactions between oxygen on CO2 and hydrogen on the pyridyl moieties of the spacers. High CO2/N2 and CO2/CH4 gas mixture selectivities under ambient conditions were achieved through IAST calculation and validated through dynamic breakthrough experiments with high CO2 loadings in mixed gas phases. Considering limited examples of microporous MOFs based on Co(II) ions compared to their Zn(II) and Cd(II) counterparts and a rich library of organic ligands/linkers, this study may be useful toward development of permanently microporous Co(II) MOFs for preferential CO2 sorption for various gas mixture separations with obvious high CO2 loadings in mixed phases.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.9b01402. Instrumental details, crystallographic parameters, PXRD data, TGA plot, Qst plot, IAST calculation details with related plots and data, dynamic breakthrough experiments, molecular simulation data (PDF) Accession Codes

CCDC 1898059 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Arun Pal: 0000-0002-0665-3446 David G. Madden: 0000-0003-3875-9146 Madhab C. Das: 0000-0002-6571-8705 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS A.P. acknowledges UGC for the research fellowship (SRF). This research work is financially supported by SERB, New Delhi as an Early Carrier Research Award (ECR/2015/ 000041) to M.C.D.



REFERENCES

(1) Yamasaki, A. An Overview of CO2 Mitigation Options for Global Warming-Emphasizing CO2 Sequestration Options. J. Chem. Eng. Jpn. 2003, 36, 361−375. F

DOI: 10.1021/acs.inorgchem.9b01402 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (2) (a) Sayari, A.; Belmabkhout, Y. Stabilization of AmineContaining CO2 Adsorbents: Dramatic Effect of Water Vapor. J. Am. Chem. Soc. 2010, 132, 6312−6314. (b) Belmabkhout, Y.; Sayari, A. Effect of Pore Expansion and Amine Functionalization of Mesoporous Silica on CO2 Adsorption over a Wide Range of Conditions. Adsorption 2009, 15, 318−328. (3) (a) Special Issue on MOF: Zhou, H.-C.; Long, J. R.; Yaghi, O. M. Introduction to Metal−Organic Frameworks. Chem. Rev. 2012, 112, 673−674. (b) Klein, N.; Senkovska, I.; Gedrich, K.; Stoeck, U.; Henschel, A.; Mueller, U.; Kaskel, S. A Mesoporous Metal−Organic Framework. Angew. Chem., Int. Ed. 2009, 48, 9954−9957. (c) Banerjee, R.; Phan, A.; Wang, B.; Knobler, C.; Furukawa, H.; O’Keeffe, M.; Yaghi, O. M. High-Throughput Synthesis of Zeolitic Imidazolate Frameworks and Application to CO2 Capture. Science 2008, 319, 939−943. (d) 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. (e) Kitagawa, S.; Kitaura, R.; Noro, S.-I. Functional Porous Coordination Polymers. Angew. Chem., Int. Ed. 2004, 43, 2334− 2375. (f) Zhang, J. − P.; Zhang, Y. − B.; Lin, J. − B.; Chen, X. − M. Metal Azolate Frameworks: From Crystal Engineering to Functional Materials. Chem. Rev. 2012, 112, 1001−1033. (g) Zhang, M.; Feng, G.; Song, Z.; Zhou, Y.-P.; Chao, H.-Y.; Yuan, D.; Tan, T. T. Y.; Guo, Z.; Hu, Z.; Tang, B. Z.; Liu, B.; Zhao, D. Two-Dimensional Metal− Organic Framework with Wide Channels and Responsive Turn-On Fluorescence for the Chemical Sensing of Volatile Organic Compounds. J. Am. Chem. Soc. 2014, 136, 7241−7244. (h) Shen, K.; Zhang, L.; Chen, X.; Liu, L.; Zhang, D.; Han, Y.; Chen, J.; Long, J.; Luque, R.; Li, Y.; Chen, B. Ordered Macro-Microporous MetalOrganic Framework Single Crystals. Science 2018, 359, 206−210. (i) Lustig, W. P.; Mukherjee, S.; Rudd, N. D.; Desai, A. V.; Li, J.; Ghosh, S. K. Metal-Organic Frameworks: Functional Luminescent and Photonic Materials for Sensing Applications. Chem. Soc. Rev. 2017, 46, 3242−3285. (j) Mollick, S.; Mukherjee, S.; Kim, D.; Qiao, Z.; Desai, A. V.; Saha, R.; More, Y. D.; Jiang, J.; Lah, M. S.; Ghosh, S. K. Hydrophobic Shielding of Outer Surface: Enhancing the Chemical Stability of Metal−Organic Polyhedra. Angew. Chem., Int. Ed. 2019, 58, 1041−1045. (k) Williams, K.; Meng, L.; Lee, S.; Lux, L.; Gao, W.; Ma, S. Imparting Bronsted Acidity into a Zeolitic Imidazole Framework. Inorg. Chem. Front. 2016, 3, 393−396. (l) Miller, R. G.; Southon, P. D.; Kepert, C. J.; Brooker, S. Commensurate CO2 Capture, and Shape Selectivity for HCCH over H2CCH2, in Zigzag Channels of a Robust CuI(CN)(L) Metal−Organic Framework. Inorg. Chem. 2016, 55, 6195−6200. (4) (a) Themed collection on MOF: Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective Gas Adsorption and Separation in Metal−organic Frameworks. Chem. Soc. Rev. 2009, 38, 1477−1504. (b) Zhang, Z.; Zaworotko, M. J. Template-directed Synthesis of Metal−organic Materials. Chem. Soc. Rev. 2014, 43, 5444−5455. (c) Deria, P.; Mondloch, J. E.; Karagiaridi, O.; Bury, W.; Hupp, J. T.; Farha, O. K. Beyond Post-synthesis Modification: Evolution of Metal−organic Frameworks via Building Block Replacement. Chem. Soc. Rev. 2014, 43, 5896−5912. (d) Nandi, S.; Collins, S.; Chakraborty, D.; Banerjee, D.; Thallapally, P. K.; Woo, T. K.; Vaidhyanathan, R. Ultralow Parasitic Energy for Postcombustion CO2 Capture Realized in a Nickel Isonicotinate Metal−Organic Framework with Excellent Moisture Stability. J. Am. Chem. Soc. 2017, 139, 1734−1737. (e) Shalini, S.; Nandi, S.; Justin, A.; Maity, R.; Vaidhyanathan, R. Potential of Ultramicroporous Metal−organic Frameworks in CO2 Clean-up. Chem. Commun. 2018, 54, 13472−13490. (f) Murray, L. J.; Dincă, M.; Long, J. R. Hydrogen Storage in Metal−organic Frameworks. Chem. Soc. Rev. 2009, 38, 1294−1314. (5) (a) Lin, R.-B.; Xiang, S.; Xing, H.; Zhou, W.; Chen, B. Exploration of Porous Metal−organic Frameworks for Gas Separation and Purification. Coord. Chem. Rev. 2019, 378, 87−103. (b) He, Y.; Zhou, W.; Qian, G.; Chen, B. Methane Storage in Metal−organic Frameworks. Chem. Soc. Rev. 2014, 43, 5657−5678. (c) Das, M. C.; Guo, Q.; He, Y.; Kim, J.; Zhao, C.-G.; Hong, K.; Xiang, S.; Zhang, Z.; Thomas, K. M.; Krishna, R.; Chen, B. Interplay of Metalloligand and

Organic Ligand to Tune Micropores within Isostructural Mixed-Metal Organic Frameworks (M′MOFs) for Their Highly Selective Separation of Chiral and Achiral Small Molecules. J. Am. Chem. Soc. 2012, 134, 8703−8710. (6) (a) Sharma, V.; De, D.; Saha, R.; Das, R.; Chattaraj, P. K.; Bharadwaj, P. K. A Cu(II)-MOF Capable of Fixing CO2 from Air and Showing High Capacity H2 and CO2 Adsorption. Chem. Commun. 2017, 53, 13371−13374. (b) Das, M. C.; Bharadwaj, P. K. A Porous Coordination Polymer Exhibiting Reversible Single-Crystal to SingleCrystal Substitution Reactions at Mn(II) Centers by Nitrile Guest Molecules. J. Am. Chem. Soc. 2009, 131, 10942−10949. (c) Tan, Y. − X.; Si, Y.; Wang, W.; Yuan, D. Tetrahedral Crosslinking of dia-type Nets into a Zeolitic GIS-type Framework for Optimizing Stability and Gas Sorption. J. Mater. Chem. A 2017, 5, 23276−23282. (d) Sikdar, N.; Bhogra, M.; Waghmare, U. V.; Maji, T. K. Oriented Attachment Growth of Anisotropic Meso/Nanoscale MOFs: Tunable Surface Area and CO2 Separation. J. Mater. Chem. A 2017, 5, 20959−20968. (e) Ye, Y.; Chen, S.; Chen, L.; Huang, J.; Ma, Z.; Li, Z.; Yao, Z.; Zhang, J.; Zhang, Z.; Xiang, S. Additive-Induced Supramolecular Isomerism and Enhancement of Robustness in Co(II)-Based MOFs for Efficiently Trapping Acetylene from Acetylene-Containing Mixtures. ACS Appl. Mater. Interfaces 2018, 10, 30912−30918. (f) Tomar, K.; Rajak, R.; Sanda, S.; Konar, S. Synthesis and Characterization of Polyhedral-Based Metal−Organic Frameworks Using a Flexible Bipyrazole Ligand: Topological Analysis and Sorption Property Studies. Cryst. Growth Des. 2015, 15, 2732− 2741. (g) Guo, Z.-J.; Yu, J.; Zhang, Y.-Z.; Zhang, J.; Chen, Y.; Wu, Y.; Xie, L.-H.; Li, J.-R. Water-Stable In(III)-Based Metal−Organic Frameworks with Rod-Shaped Secondary Building Units: SingleCrystal to Single-Crystal Transformation and Selective Sorption of C2H2 over CO2 and CH4. Inorg. Chem. 2017, 56, 2188−2197. (7) (a) Chand, S.; Elahi, S. M.; Pal, A.; Das, M. C. Metal Organic Frameworks and Other Crystalline Materials for Ultrahigh Superprotonic Conductivities of 10−2 S cm−1 or Higher. Chem. - Eur. J. 2019, 25, 6259−6269. (b) Yamada, T.; Otsubo, K.; Makiura, R.; Kitagawa, H. Designer Co-ordination Polymers: Dimensional Crossover Architectures and Proton Conduction. Chem. Soc. Rev. 2013, 42, 6655−6669. (c) Ramaswamy, P.; Wong, N. E.; Shimizu, G. K. H. MOFs as Proton Conductors Challenges and Opportunities. Chem. Soc. Rev. 2014, 43, 5913−5932. (d) Elahi, S. M.; Chand, S.; Deng, W.H.; Pal, A.; Das, M. C. Polycarboxylates Templated Coordination Polymers: Role of Templates for Superprotonic Conductivities up to 10−1 S cm−1. Angew. Chem., Int. Ed. 2018, 57, 6662−6666. (e) Chand, S.; Pal, S. C.; Pal, A.; Ye, Y.; Lin, Q.; Zhang, Z.; Xiang, S.; Das, M. C. Metalo Hydrogen-Bonded Organic Frameworks (MHOFs) as New Class of Crystalline Materials for Protonic Conduction. Chem. - Eur. J. 2019, 25, 1691−1695. (8) (a) Zhao, X.; Bu, X.; Zhai, Q.-G.; Tran, H.; Feng, P. Pore Space Partition by Symmetry-Matching Regulated Ligand Insertion and Dramatic Tuning on Carbon Dioxide Uptake. J. Am. Chem. Soc. 2015, 137, 1396−1399. (b) Elsaidi, S. K.; Mohamed, M. H.; Wojtas, L.; Chanthapally, A.; Pham, T.; Space, B.; Vittal, J. J.; Zaworotko, M. J. Putting the Squeeze on CH4 and CO2 through Control over Interpenetration in Diamondoid Nets. J. Am. Chem. Soc. 2014, 136, 5072−5077. (9) (a) 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. (b) Vaidhyanathan, R.; Iremonger, S. S.; Shimizu, G. K. H.; Boyd, P. G.; Alavi, S.; Woo, T. K. Direct Observation and Quantification of CO2 Binding Within an Amine-Functionalized Nanoporous Solid. Science 2010, 330, 650− 653. (c) Luo, J.; Wang, J.; Li, G.; Huo, Q.; Liu, Y. Assembly of a Unique Octa-nuclear Copper Cluster-based Metal−organic Framework with Highly Selective CO2 Adsorption over N2 and CH4. Chem. Commun. 2013, 49, 11433−11435. (d) Chen, K. − J.; Scott, H. S.; Madden, D. G.; Pham, T.; Kumar, A.; Bajpai, A.; Lusi, M.; Forrest, K. A.; Space, B.; Perry IV, J. J.; Zaworotko, M. J. Benchmark C2H2/CO2 G

DOI: 10.1021/acs.inorgchem.9b01402 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Hysteretic Sorption of Carbon Dioxide. Nat. Mater. 2012, 11, 710− 716. (b) Liang, Z.; Marshall, M.; Chaffee, A. L. CO2 AdsorptionBased Separation by Metal Organic Framework (Cu-BTC) versus Zeolite (13X). Energy Fuels 2009, 23, 2785−2789. (c) Nugent, P.; Rhodus, V.; Pham, T.; Tudor, B.; Forrest, K.; Wojtas, L.; Space, B.; Zaworotko, M. Enhancement of CO2 Selectivity in a Pillared pcu MOM Platform through Pillar Substitution. Chem. Commun. 2013, 49, 1606−1608. (d) Alawisi, H.; Li, B.; He, Y.; Arman, H. D.; Asiri, A. M.; Wang, H.; Chen, B. A Microporous Metal−Organic Framework Constructed from a New Tetracarboxylic Acid for Selective Gas Separation. Cryst. Growth Des. 2014, 14, 2522−2526. (e) Li, J.-R.; Yu, J.; Lu, W.; Sun, L.-B.; Sculley, J.; Balbuena, P. B.; Zhou, H.-C. Porous Materials with Pre-designed Single-molecule Traps for CO2 Selective Adsorption. Nat. Commun. 2013, 4, 204−212. (f) Phan, A.; Doonan, C. J.; Uribe-Romo, F. J.; Knobler, C. B.; O’Keeffe, M.; Yaghi, O. M. Synthesis, Structure, and Carbon Dioxide Capture Properties of Zeolitic Imidazolate Frameworks. Acc. Chem. Res. 2010, 43, 58−67. (g) Pal, A.; Mitra, A.; Chand, S.; Lin, J.-B.; Das, M. C. Two 2D microporous MOFs based on Bent Carboxylates and a Linear Spacer for Selective CO2 Adsorption. CrystEngComm 2019, 21, 535−543. (h) Li, Y.-Z.; Wang, H.-H.; Yang, H.-Y.; Hou, L.; Wang, Y.-Y.; Zhu, Z. An Uncommon Carboxyl-Decorated MOF with Selective Gas Adsorption and Catalytic Conversion of CO2. Chem. - Eur. J. 2018, 24, 865−871. (i) Yang, H.-Y.; Li, Y.-Z.; Shi, W.-J.; Hou, L.; Wang, Y.Y.; Zhu, Z. A New Layer-stacked Porous Framework Showing Sorption Selectivity for CO2 and Luminescence. Dalton Trans. 2017, 46, 11722−11727. (21) Chand, S.; Pal, A.; Das, M. C. A Moisture-Stable 3D Microporous CoII-Metal−Organic Framework with Potential for Highly Selective CO2 Separation under Ambient Conditions. Chem. - Eur. J. 2018, 24, 5982−5986. (22) (a) García, E. J.; Pérez-Pellitero, J.; Pirngruber, G. D.; Jallut, C.; Palomino, M.; Rey, F.; Valencia, S. Tuning the Adsorption Properties of Zeolites as Adsorbents for CO2 Separation: Best Compromise between the Working Capacity and Selectivity. Ind. Eng. Chem. Res. 2014, 53, 9860−9874. (b) Hamon, L.; Jolimaître, E.; Pirngruber, G. D. CO2 and CH4 Separation by Adsorption Using Cu-BTC MetalOrganic Framework. Ind. Eng. Chem. Res. 2010, 49, 7497−7503. (23) Wang, B.; Cote, A. P.; Furukawa, H.; O’Keeffe, M.; Yaghi, O. M. Colossal Cages in Zeolitic Imidazolate Frameworks as Selective Carbon Dioxide Reservoirs. Nature 2008, 453, 207−211.

and CO2/C2H2 Separation by Two Closely Related Hybrid Ultramicroporous Materials. Chem 2016, 1, 753−765. (10) (a) 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. (b) Yu, J.; Xie, L. − H.; Li, J. − R.; Ma, Y.; Seminario, J. M.; Balbuena, P. B. CO2 Capture and Separations Using MOFs: Computational and Experimental Studies. Chem. Rev. 2017, 117, 9674−9754. (11) Xiang, S.; He, Y.; Zhang, Z.; Wu, H.; Zhou, W.; Krishna, R.; Chen, B. Microporous Metal-organic Framework with Potential for Carbon Dioxide Capture at Ambient Conditions. Nat. Commun. 2012, 3, 954−963. (12) Lu, Z.; Godfrey, H. G. W.; da Silva, I.; Cheng, Y.; Savage, M.; Tuna, F.; McInnes, E. J. L.; Teat, S. J.; Gagnon, K. J.; Frogley, M. D.; Manuel, P.; Rudic, S.; Ramirez-Cuesta, A. J.; Easun, T. L.; Yang, S.; Schroder, M. Modulating Supramolecular Binding of Carbon Dioxide in a Redox-active Porous Metal-organic Framework. Nat. Commun. 2017, 8, 14212. (13) Myers, A. L.; Prausnitz, J. M. Thermodynamics of Mixed-gas Adsorption. AIChE J. 1965, 11, 121−127. (14) Pal, A.; Chand, S.; Das, M. C. A Water-Stable Twofold Interpenetrating Microporous MOF for Selective CO2 Adsorption and Separation. Inorg. Chem. 2017, 56, 13991−13997. (15) Blatov, V. A. IUCr CompComm Newsletter 2006, 7, 4. http:// www.topos.ssu.samara.ru. (16) Spek, A. L. Single-crystal Structure Validation with the Program PLATON. J. Appl. Crystallogr. 2003, 36, 7−13. (17) (a) Zheng, B.; Bai, J.; Duan, J.; Wojtas, L.; Zaworotko, M. J. Enhanced CO2 Binding Affinity of a High-Uptake rht-Type Metal− Organic Framework Decorated with Acylamide Groups. J. Am. Chem. Soc. 2011, 133, 748−751. (b) Lu, Z.; Meng, F.; Du, L.; Jiang, W.; Cao, H.; Duan, J.; Huang, H.; He, H. A Free Tetrazolyl Decorated Metal−Organic Framework Exhibiting High and Selective CO2 Adsorption. Inorg. Chem. 2018, 57, 14018−14022. (c) Tan, C.; Yang, S.; Champness, N. R.; Lin, X.; Blake, A. J.; Lewis, W.; Schröder, M. High Capacity Gas Storage by a 4,8-connected Metal−organic Polyhedral Framework. Chem. Commun. 2011, 47, 4487−4489. (d) Pal, A.; Chand, S.; Elahi, S. M.; Das, M. C. A Microporous MOF with a Polar Pore Surface Exhibiting Excellent Selective Adsorption of CO2 from CO2-N2 and CO2-CH4 Gas Mixtures with High CO2 Loading. Dalton Trans. 2017, 46, 15280−15286. (e) Nugent, P.; Belmabkhout, Y.; Burd, S. D.; Cairns, A. J.; Luebke, R.; Forrest, K.; Pham, T.; Ma, S.; Space, B.; Wojtas, L.; Eddaoudi, M.; Zaworotko, M. J. Porous Materials with Optimal Adsorption Thermodynamics and Kinetics for CO2 Separation. Nature 2013, 495, 80−84. (f) Demessence, A.; D’Alessandro, D. M.; Foo, M. L.; Long, J. R. Strong CO2 Binding in a Water-Stable, Triazolate Bridged Metal−Organic Framework Functionalized with Ethylenediamine. J. Am. Chem. Soc. 2009, 131, 8784−8786. (g) Wen, H.-M.; Chang, G.; Li, B.; Lin, R.-B.; Hu, T.-L.; Zhou, W.; Chen, B. Highly Enhanced Gas Uptake and Selectivity via Incorporating Methoxy Groups into a Microporous Metal−Organic Framework. Cryst. Growth Des. 2017, 17, 2172−2177. (18) Krishna, R.; van Baten, J. M. In Silico Screening of Metal− Organic Frameworks in Separation Applications. Phys. Chem. Chem. Phys. 2011, 13, 10593−10616. (19) (a) Alduhaish, O.; Wang, H.; Li, B.; Arman, H. D.; Nesterov, V.; Alfooty, K.; Chen, B. A Threefold Interpenetrated Pillared-Layer Metal−Organic Framework for Selective Separation of C2H2/CH4 and CO2/CH4. ChemPlusChem 2016, 81, 764−769. (b) Pal, A.; Lin, J. − B.; Chand, S.; Das, M. C. A 3D Microporous MOF with mab Topology for Selective CO2 Adsorption and Separation. ChemistrySelect 2018, 3, 917−921. (20) (a) Yang, S.; Lin, X.; Lewis, W.; Suyetin, M.; Bichoutskaia, E.; Parker, J. E.; Tang, C. C.; Allan, D. R.; Rizkallah, P. J.; Hubberstey, P.; Champness, N. R.; Thomas, K. M.; Blake, A. J.; Schröder, M. A Partially Interpenetrated Metal−organic Framework for Selective H

DOI: 10.1021/acs.inorgchem.9b01402 Inorg. Chem. XXXX, XXX, XXX−XXX