Response of Metal Sites toward Water Effects on Postcombustion CO2

Feb 23, 2016 - *E-mail: [email protected] (J.Y.)., *E-mail: [email protected]. ... Coordinatively unsaturated metal (CUM) sites of metal–organic fr...
7 downloads 0 Views 2MB Size
Research Article pubs.acs.org/journal/ascecg

Response of Metal Sites toward Water Effects on Postcombustion CO2 Capture in Metal−Organic Frameworks Jiamei Yu,*,†,‡,§ Yufeng Wu,*,† and Perla B. Balbuena‡,§ †

Institute of Recycling Economy, Beijing University of Technology, Beijing, 100124, China Artie McFerrin Department of Chemical Engineering and §Department of Materials Science and Engineering, Texas A&M University, College Station, Texas 77842-3122, United States



S Supporting Information *

ABSTRACT: Coordinatively unsaturated metal (CUM) sites of metal−organic frameworks (MOFs) play an important role in water adsorption and its effects on CO2 capture. To explore the trends among various metal identities toward water, M-HKUST-1 analogous to HKUST-1 based on different metal sites M (M = Zn, Co, Ni, and Mg) are chosen as model frameworks and their water effects on CO2 capture are evaluated. Our results demonstrate that Mg-based framework shows diverse effects by water coordination from Zn-, Co-, and Ni-based frameworks. Water coordination lowers CO2 uptake dramatically in Mg-HKUST-1 whereas improves CO2 adsorption in Zn-, Co-, and Ni-HKUST-1 at most of the pressure ranges we studied. The detailed evaluation of interaction energy indicates that with the increased loading of CO2 molecules both the weaker interactions between CO2 and the framework and CO2−CO2 contribute to the lower CO2 uptakes of the hydrated Mg-HKUST-1. These findings could benefit CUMs screening in MOFs’ design for CO2 capture and other applications. KEYWORDS: Molecular simulations, Water effects, Metal−organic frameworks (MOFs), Coordinatively unsaturated metal sites (CUMs), Adsorption



the presence of water in the flue gas lowers the CO2 adsorption and selectivity in MOFs since water competes with CO2 for the adsorption sites.22−26 Barbarao et al.24 showed that even with 0.1% water in CO2/CH4 mixture, the CO2/CH4 selectivity in rho-ZMOF decreases by 1 order of magnitude. Lin et al.26 showed that the presence of H2O preferentially occupies the CO2 binding sites, resulting in a large reduction of the CO2 uptakes. However, this is not always the case. For example, Yazaydin et al.27 reported that the presence of 4 wt % water coordinated to the CUMs in the HKUST-1 framework can enhance CO2 capture capacity and selectivity over N2 and CH4. Water molecules bound to CUMs in HKUST-1 and the selectively of small gases on the metal centers have been further explored by a couple of groups.28−30 The authors have previously investigated the effects of coordinated water on CO2 adsorption and CO2/N2 separation properties in MOFs based on two typical MOFs with CUMs: HKUST-1 and MgMOF-74.31,32 It has been found that water coordination improves CO2 adsorption in HKUST-1 but lowers CO2 uptakes in Mg-MOF-74. Density functional theory (DFT) calculations indicate that water coordinated on Cu in HKUST1 increases CO2 binding affinity. In contrast, water coordinated on Mg in Mg-MOF-74 lowers CO2 binding affinity. Those works indicates the CUMs identity plays a central role in

INTRODUCTION Coal-fired power plants emit flue gas comprising primarily of CO2 and N2 with the remaining fraction a mixture of water, O2, and other trace impurities.1 Selective removal of CO2 from power plant emissions is required and would help to mitigate atmospheric CO2 levels. The adsorbent material, as the separation media, plays a very important role. Porous materials relying on physical adsorption of CO2 provide promising energy efficient alternatives to the current amine-based absorption systems. Many porous solids including activated carbons, zeolites and porous polymers have been of great interests for this purpose.2−5 Metal−organic frameworks (MOFs) are being intensively investigated in separation of CO2 from gas mixtures, such as natural gas and flue gas due to their flexible chemistry.6−10 Particularly, MOFs allow facile optimization of the pore structure, surface functionalization, and property tuning based on molecular design for specific applications.6,11 Of various MOFs, MOFs with coordinatively unsaturated metal sites (CUMs) are regarded as an ideal sorbent for CO2 capture, because the CUMs show high adsorption energy for CO2.12−19 However, there are still issues which need to be addressed for practical usage scenarios. One concern is the influence of impurities in flue gas mixtures on the CO2 capture in MOFs. Those impurities may influence the performance of MOFs on CO2 capture significantly.20,21 Among various impurities, water has drawn special attentions since it can take up about 10% molar concentration in flue gas mixtures.1 It is generally accepted that © XXXX American Chemical Society

Received: January 13, 2016 Revised: February 19, 2016

A

DOI: 10.1021/acssuschemeng.6b00080 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

PBE functional.53 Spin polarization is considered in the calculation. The surface Brillouin zone is sampled with 4 × 4 × 4 k-points generated by the Monkhorst−Pack method.54 The convergence criteria for energy and force are 10−5 eV and 0.01 eV/Å respectively. Both lattice parameters and atoms were allowed to relax in the simulations. The initial structures for MHKUST-1 were obtained by replacing Cu sites in HKUST-1 by Mg, Zn, Co, and Ni (Figure S1). DFT calculations with GAUSSIAN 09 suite of programs55 using the B3LYP functional with the 6-31+G (d) basis set were performed for two purposes. First is to optimize the position of hydrogen atoms of the coordinated water molecules in MHKUST-1 frameworks. The second is to calculated the atomic partial charges for MOFs using the CHELPG methods.56

identifying the water effects on CO2 capture. Although the studies noted above provides us important information about water effects in some specific MOFs, none of an isostructural MOF series has been explored to determine trends among various metal ions. Such studies are valuable since they can eliminate the influence induced by all other variables such as pore size, pore shape, and surface area, therefore providing direct insight into the nature of the CO2−metal (water) interactions. For this purpose, the effects on CO2 adsorption and its selectivity over N2 by water coordination in a series of isostructural frameworks with different CUMs are systematically investigated in this work. The isostructural frameworks MHKUST-1 (where M = Mg, Zn, Co, and Ni) analogous to HKUST-1 (with CUMs Cu)33 are chosen as model materials since HKUST-1 has been the subject of intense study in CO2 capture for its large surface area, high pore volume, high chemical stability, and the Lewis acidity of CUMs in its framework.34−42 Variations in the electronic structure and the geometry of the structural building unit are examined and used to rationalize trends for water effects. This work could contribute to the CUMs screening in MOFs’ design for CO2 capture and other applications.



RESULTS AND DISCUSSION By replacing Cu with Mg, Zn, Co, and Ni, the crystal structures of M-HKUST-1 (where M = Mg, Zn, Co, and Ni) were first optimized to get the equilibrium lattice parameters (given in Table 1) and atomic coordinates for M-HKUST-1.



Table 1. Equilibrium Lattice Parameters of M-HKUST-1 and Distances between Water and M in Half and Fully Hydrated M-HKUST-1 (M = Mg, Zn, Co, and Ni)

COMPUTATIONAL METHODS GCMC simulations43 using the MUSIC program44 were employed to evaluate the adsorption of single components and their mixtures in M-HKUST-1. All of the MOFs were treated as rigid frameworks. A cutoff radius of 12.8 Å was applied to the Lennard-Jones (LJ) potential and Ewald summations were used to evaluate long-range effects of the electrostatic interactions.45 Each GCMC simulation consisted of 1 × 107 steps to guarantee equilibrium and 1 × 107 production steps. CO2 was modeled as a rigid linear triatomic molecule with one charged LJ interaction site located at each atom.46 The LJ potential parameters for O−O interactions are σOO = 0.305 nm and εOO/kB = 79.0 K, and for C−C interactions are σCC = 0.280 nm and εCC/kB = 27.0 K with a C−O bond length of 0.116 nm. Partial point charges are centered at each LJ site (qO = −0.35e and qC = 0.70e). The N2 molecule was also represented as a rigid three-site model with two sites located at two N atoms and the third one located at its center of mass (COM). The site at each N atom was mimicked by a LJ interaction potential σNN = 0.331 nm and εNN/kB = 36.0 K with a bond length of 0.110 nm between two N atoms. A negative charge was assigned on each N atom with qN = −0.482e and a positive charge at the COM site with qCOM = 0.964e in each N2 molecule.46 Partial charges for the water molecule were calculated from DFT and the LJ potential parameters were taken from the TIP3P47 water model. The LJ potential parameters for the framework atoms in M-HKUST-1 were taken from the DREIDING48 or UFF49 force field. Table S1 lists the LJ parameters for water and framework atoms. Lorentz−Berthelot mixing rules were employed to calculate the interactions between the framework sites and each gas species. The Vienna ab initio simulation package (VASP) code,50 a DFT program package51 with plane wave basis sets was used to optimize the crystal structures of M-HKUST-1 (where M = Mg, Zn, Co, and Ni). The interaction between ions and electrons was described using the projector augmented wave method.52 A plane wave basis set with a cutoff energy of 520 eV was used to expand the wave function of valence electrons. The nonlocal exchange-correlation energy was evaluated using the

material

a0 (Å)

M−water distance (Å) in half hydrated frameworks

Mg-HKUST-1 Zn-HKUST-1 Co-HKUST-1 Ni-HKUST-1

26.724 26.869 26.270 26.412

2.108 2.148 2.138 2.093

M−water distance (Å) in fully hydrated frameworks 2.111 2.153 2.200 2.096

Based on the equilibrium M-HKUST-1 structures, the positions of water molecules were identified. Both the oxygen and hydrogen atoms of the water molecules were allowed to relax during optimizations. After optimization, the hydrogen atoms were observed to lean toward carboxylate oxygen atoms for all M-HKUST-1 frameworks. For half hydrated M-HKUST1 one coordinated water molecule and for fully hydrated MHKUST-1 two coordinated water molecules are included for each metal corner. The optimized orientations of water molecules in half and fully hydrated M-HKUST-1 are shown in Figure S2, and the optimized distances between water and metal sites are given in Table 1. Following the identification of positions of coordinated water molecules in the frameworks, the atomic partial charges calculations for dry, half, and fully hydrated M-HKUST-1 were performed. Figure 1 shows the fragmental clusters for charge calculations in dry, half, and fully hydrated M-HKUST-1 and Table 2 lists the corresponding derived charges. GCMC simulations were performed to simulate the pure CO2 adsorption isotherms. Figure 2 shows that water

Figure 1. Fragmental clusters for partial atomic charge calculations of dry (left), half (middle), and fully (right) hydrated M-HKUST-1. B

DOI: 10.1021/acssuschemeng.6b00080 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering Table 2. Partial Atomic Charges (e) for Dry, Half, and Fully Hydrated M-HKUST-1 M-HKUST-1

M(a/b)

O(a/b)

Ca

Cb

Cc

H

Ow

Hw

dry Mg-HKUST-1 dry Co-HKUST-1 dry Zn-HKUST-1 dry Ni-HKUST-1 half hydrated Mg-HKUST-1 half hydrated Co-HKUST-1 half hydrated Zn-HKUST-1 half hydrated Ni-HKUST-1 fully hydrated Mg-HKUST-1 fully hydrated Co-HKUST-1 fully hydrated Zn-HKUST-1 fully hydrated Ni-HKUST-1

1.665 1.231 1.351 1.098 1.683/1.658 1.490/1.141 0.985/1.388 1.308/0.924 1.667 1.420 1.118 1.175

−0.821 −0.659 −0.718 −0.632 −0.905/−0.857 −0.741/−0.697 −0.687/−0.694 −0.720/−0.634 −0.908/−0.832 −0.774/−0.685 −0.711/−0.639 −0.694/−0.618

0.802 0.653 0.641 0.682 0.882 0.710 0.723 0.707 0.866 0.648 0.707 0.654

−0.049 0.027 0.091 −0.007 −0.030 0.072 0.043 0.027 0.014 0.149 0.050 0.092

−0.060 −0.126 −0.191 −0.100 −0.097 −0.173 −0.145 −0.127 −0.143 −0.236 −0.151 −0.185

0.117 0.148 0.158 0.140 0.132 0.160 0.130 0.140 0.141 0.173 0.120 0.150

NA NA NA NA −0.758 −0.726 −0.702 −0.750 −0.757 −0.730 −0.730 −0.736

NA NA NA NA 0.409 0.388 0.430 0.386 0.407 0.380 0.430 0.382

Figure 2. Simulated adsorption isotherms for CO2 in dry and hydrated Zn-HKUST-1 (top left), Ni-HKUST-1 (top right), Co-HKUST-1 (bottom left), and Mg-HKUST-1 (bottom right) at 298 K.

energies (∼10 kJ/mol) between CO2 and dry Mg-HKUST-1 are much higher than those in Zn, Co, or Ni-based frameworks (2−3 kJ/mol), indicating the stronger interaction between the quadrupole moment of CO2 and the electric field created by Mg. Therefore, Mg-HKUST-1 shows higher CO2 adsorption compared with Zn, Co, or Ni-based frameworks (Figure 2). Moreover, Figure 3 shows that both Coulombic and nonCoulombic energies between CO2 and frameworks increase with the introduction of water coordination in M-HKUST-1. In Zn, Co, and Ni-based frameworks, the increase of Coulombic energies is much higher than that of the non-Coulombic energies. Therefore, Coulombic interaction between CO2 and the electric field created by coordinated water molecules is responsible for the enhanced CO2 uptakes. This is consistent with the results of water effects on CO2 capture in HKUST-1

coordination enhances CO2 uptakes predominantly in Zn-, Co-, and Ni-HKUST-1. In contrast, CO2 uptakes decrease dramatically by water coordination in Mg-HKUST-1 at most of the pressures we evaluated, except at pressure lower than ∼0.5 bar, where a slight enhancement of CO2 adsorption was observed. To gain some insight into why the Mg-based framework shows different water effects from Zn, Co, and Ni-based MOFs, the interaction energies of CO2-frameworks and CO2−CO2 in dry and hydrated M-HKUST-1 from the GCMC simulations were examined. The interaction energies were broken into Coulombic and non-Coulombic (Lennard-Jones) components. As shown in Figure 3, the non-Coulombic energies between CO2 and dry frameworks in Mg-HKUST-1 are comparable with those in Zn, Co, or Ni-HKUST-1. However, the Coulombic C

DOI: 10.1021/acssuschemeng.6b00080 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Figure 3. Interaction energies of CO2−frameworks and CO2−CO2 in dry and hydrated Zn-HKUST-1 (top left), Ni-HKUST-1 (top right), CoHKUST-1 (bottom left), and Mg-HKUST-1 (bottom right) at 298 K.

Figure 4. Simulated selectivity for CO2/N2 mixtures with bulk compositions of 50:50 (left) and 15:85 (right) in dry and hydrated Zn-HKUST-1 at 298 K.

(with CUM Cu) reported by Yazaydin et al.27 In contrast, the Coulombic and non-Coulombic energies increase comparably in Mg-based framework. Therefore, both the electrostatic interaction and van der Waals interaction between CO2 and framework are responsible for the improvement of CO2 adsorption in Mg-HKUST-1 at low pressures. However, the increase of Coulombic energies induced by coordinated water declines with more CO2 molecules adsorbed. Considering the heavier molecular mass of hydrated frameworks, the smaller Coulombic energy gap between dry and hydrated frameworks may lead to the lower CO2 uptakes in the hydrated M-HKUST-1 at higher pressures (Figure 2). For the Mg-HKUST-1, not only the increase of Coulombic energy but

also the increase of non-Coulombic energies drops at higher CO2 loading. The smaller gaps of both Coulombic and nonCoulombic energies between CO2 and hydrated frameworks may lead to the lower CO2 uptakes at higher pressures. In terms of CO2−CO2 interactions, different from Zn, Co, and Ni-based frameworks, where the interaction energies for CO2 and CO2 in the hydrated framework are almost identical or even higher than that in the corresponding dry framework, CO2−CO2 interactions in hydrated Mg-HKUST-1 is lower than the corresponding dry framework, particularly with the increased loading of CO2 molecules. The weaker interactions between CO2 molecules in hydrated Mg-HKUST-1 may D

DOI: 10.1021/acssuschemeng.6b00080 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Figure 5. Simulated selectivity for CO2/N2 mixtures with bulk compositions of 50:50 (left) and 15:85 (right) in dry and hydrated Ni-HKUST-1 at 298 K.

Figure 6. Simulated selectivity for CO2/N2 mixtures with bulk compositions of 50:50 (left) and 15:85 (right) in dry and hydrated Co-HKUST-1 at 298 K.

Figure 7. Simulated selectivity for CO2/N2 mixtures with bulk compositions of 50:50 (left) and 15:85 (right) in dry and hydrated Mg-HKUST-1 at 298 K.

also enhanced by water coordination in Mg-HKUST-1 (Figure 7). Figure 8 shows the adsorption isotherms for a 15:85 CO2/N2 mixture in dry and half hydrated M-HKUST-1 (where M = Zn, Co, Ni, and Mg). As observed in dry and half hydrated HKUST-1, N2 adsorption isotherms in both dry and hydrated M-HKUST-1 (where M = Zn, Ni, and Co) are similar. The selectivity improvement in hydrated frameworks is mainly attributed to the increased CO2 adsorption by water coordination in Zn-HKUST-1, Ni-HKUST-1, and CoHKUST-1. For Mg-HKUST-1, the improvement of selectivity

additionally contribute to the lower CO2 uptakes at higher pressures. Figures 4−7 show the selectivity for CO2/N2 mixtures with bulk composition 50:50 and 15:85 in dry and hydrated MHKUST-1 (where M = Zn, Ni, Co and Mg). As noted above, since the coordinated water increases the CO2 adsorption capacity in Zn-HKUST, Ni-HKUST-1, and Co-HKUST-1 frameworks, we expected the selectivity of CO2 over N2 in these frameworks to be improved by water coordination. The results shown in Figure 4−6 verify our prediction, where the CO2/N2 selectivity increases with hydration levels for both bulk compositions. It should be noted that the CO2/N2 selectivity is E

DOI: 10.1021/acssuschemeng.6b00080 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Figure 8. CO2 and N2 isotherms for a 15:85 CO2/N2 mixture in dry and half hydrated Zn-HKUST-1 (top left), Ni-HKUST-1 (top right), CoHKUST-1 (bottom left), and Mg- HKUST-1 (bottom right) at 298 K.



in hydrated frameworks also can be ascribed to the increased CO2 adsorption at low pressures.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.6b00080. Schematic diagram of M-HKUST-1 structure (Figure S1). Orientation of hydrogen atoms of water molecules in half and fully hydrated M-HKUST-1 (Figure S2). Force field parameters for water and framework atoms (Table S1) (PDF)

CONCLUSIONS

To explore the response of various metals toward water, water effects on CO2 capture in an isostructural series M-HKUST-1 (M = Zn, Co, Ni, and Mg) are investigated. We have found depending on CUMs identity, water coordination shows different effects on CO2 adsorption and selectivity. In Zn-, Co-, and Ni-based frameworks, water coordination on CUMs mainly improves CO2 uptakes and its selectivity over N2. However, water coordination lowers CO2 adsorption in MgHKUST-1 remarkably at most of pressure ranges we evaluated. A detailed analysis of simulation data revealed that the Coulombic interaction between the quadrupole moment of CO2 and the electric field created by coordinated water molecules is responsible for the enhanced CO2 uptakes in Zn-, Co-, and Ni-based frameworks. In contrast, both the electrostatic interaction and van der Waals interaction are responsible for the CO2 uptakes improvement in Mg-HKUST-1 at low pressures. With the increased loading of CO2 molecules in the frameworks, the gap of the interaction energies between dry and hydrated frameworks with CO2 narrows down. Considering the heavier molecular mass of hydrated frameworks, water coordination lowers CO2 uptakes at higher pressures, particularly in Mg-based framework. Moreover, the weaker interaction between CO2 and CO2 molecules by water coordination at higher pressures may additionally lead to the lower CO2 uptakes in Mg-HKUST-1.



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (J.Y.). *E-mail: [email protected]. (Y.W.). Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS

We are thankful for the financial support from the National Natural Science Foundation of China (No. 21506003) and ARPA-E of Department of Energy in USA through the IMPACCT program (AR0000073). Computational resources from the Texas A&M Supercomputing Facilities and Texas A&M University Brazos HPC cluster are gratefully acknowledged. F

DOI: 10.1021/acssuschemeng.6b00080 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering



(20) Yu, K.; Kiesling, K.; Schmidt, J. R. Trace Flue Gas Contaminants Poison Coordinatively Unsaturated Metal-Organic Frameworks: Implications for CO2 Adsorption and Separation. J. Phys. Chem. C 2012, 116, 20480−20488. (21) Sun, W. Z.; Lin, L. C.; Peng, X.; Smit, B. Computational screening of porous metal-organic frameworks and zeolites for the removal of SO2 and NOx from flue gases. AIChE J. 2014, 60, 2314− 2323. (22) Tan, K.; Zuluaga, S.; Gong, Q. H.; Gao, Y. Z.; Nijem, N.; Li, J.; Thonhauser, T.; Chabal, Y. J. Competitive Coadsorption of CO2 with H2O, NH3, SO2, NO, NO2, N2, O2, and CH4 in M-MOF-74 (M = Mg, Co, Ni): The Role of Hydrogen Bonding. Chem. Mater. 2015, 27, 2203−2217. (23) Liu, J.; Wang, Y.; Benin, A. I.; Jakubczak, P.; Willis, R. R.; LeVan, M. D. CO2/H2O Adsorption Equilibrium and Rates on Metal-Organic Frameworks: HKUST-1 and Ni/DOBDC. Langmuir 2010, 26, 14301−14307. (24) Babarao, R.; Jiang, J. W. Upgrade of natural gas in rho zeolitelike metal-organic framework and effect of water: a computational study. Energy Environ. Sci. 2009, 2, 1088−1093. (25) Huang, H.; Zhang, W.; Liu, D.; Zhong, C. Understanding the Effect of Trace Amount of Water on CO2 Capture in Natural Gas Upgrading in Metal−Organic Frameworks: A Molecular Simulation Study. Ind. Eng. Chem. Res. 2012, 51, 10031−10038. (26) Lin, L. C.; Lee, K.; Gagliardi, L.; Neaton, J. B.; Smit, B. ForceField Development from Electronic Structure Calculations with Periodic Boundary Conditions: Applications to Gaseous Adsorption and Transport in Metal-Organic Frameworks. J. Chem. Theory Comput. 2014, 10, 1477−1488. (27) Yazaydin, A. O.; Benin, A. I.; Faheem, S. A.; Jakubczak, P.; Low, J. J.; Willis, R. R.; Snurr, R. Q. Enhanced CO2 Adsorption in MetalOrganic Frameworks via Occupation of Open-Metal Sites by Coordinated Water Molecules. Chem. Mater. 2009, 21, 1425−1430. (28) Cockayne, E.; Nelson, E. B. Density functional theory metaGGA plus U study of water incorporation in the metal-organic framework material Cu-BTC. J. Chem. Phys. 2015, 143, 024701− 024707. (29) Toda, J.; Fischer, M.; Jorge, M.; Gomes, J. R. B. Water adsorption on a copper formate paddlewheel model of CuBTC: A comparative MP2 and DFT study. Chem. Phys. Lett. 2013, 587, 7−13. (30) Supronowicz, B.; Mavrandonakis, A.; Heine, T. Interaction of Small Gases with the Unsaturated Metal Centers of the HKUST-1 Metal Organic Framework. J. Phys. Chem. C 2013, 117, 14570−14578. (31) Yu, J. M.; Ma, Y. G.; Balbuena, P. B. Evaluation of the Impact of H2O, O2, and SO2 on Postcombustion CO2 Capture in Metal-Organic Frameworks. Langmuir 2012, 28, 8064−8071. (32) Yu, J. M.; Balbuena, P. B. Water Effects on Postcombustion CO2 Capture in Mg-MOF-74. J. Phys. Chem. C 2013, 117, 3383−3388. (33) Chui, S. S. Y.; Lo, S. M. F.; Charmant, J. P. H.; Orpen, A. G.; Williams, I. D. A chemically functionalizable nanoporous material [Cu3(TMA)(2)(H2O)(3)](n). Science 1999, 283, 1148−1150. (34) Wu, H.; Simmons, J. M.; Srinivas, G.; Zhou, W.; Yildirim, T. Adsorption Sites and Binding Nature of CO2 in Prototypical MetalOrganic Frameworks: A Combined Neutron Diffraction and FirstPrinciples Study. J. Phys. Chem. Lett. 2010, 1, 1946−1951. (35) Wade, C. R.; Dinca, M. Investigation of the synthesis, activation, and isosteric heats of CO2 adsorption of the isostructural series of metal-organic frameworks M3(BTC)2(M = Cr, Fe, Ni, Cu, Mo, Ru). Dalton Trans. 2012, 41, 7931−7938. (36) Montoro, C.; Garcia, E.; Calero, S.; Perez-Fernandez, M. A.; Lopez, A. L.; Barea, E.; Navarro, J. A. R. Functionalisation of MOF open metal sites with pendant amines for CO2 capture. J. Mater. Chem. 2012, 22, 10155−10158. (37) Pirngruber, G. D.; Hamon, L.; Bourrelly, S.; Llewellyn, P. L.; Lenoir, E.; Guillerm, V.; Serre, C.; Devic, T. A Method for Screening the Potential of MOFs as CO2 Adsorbents in Pressure Swing Adsorption Processes. ChemSusChem 2012, 5, 762−776. (38) Zhao, Y. X.; Seredych, M.; Zhong, Q.; Bandosz, T. J. Superior Performance of Copper Based MOF and Aminated Graphite Oxide

REFERENCES

(1) Lin, L. C.; Berger, A. H.; Martin, R. L.; Kim, J.; Swisher, J. A.; Jariwala, K.; Rycroft, C. H.; Bhown, A. S.; Deem, M. W.; Haranczyk, M.; Smit, B. In silico screening of carbon-capture materials. Nat. Mater. 2012, 11, 633−641. (2) Lee, J. S.; Kim, J. H.; Kim, J. T.; Suh, J. K.; Lee, J. M.; Lee, C. H. Adsorption equilibria of CO2 on zeolite 13X and zeolite X/Activated carbon composite. J. Chem. Eng. Data 2002, 47, 1237−1242. (3) Wang, Y.; LeVan, M. D. Adsorption Equilibrium of Carbon Dioxide and Water Vapor on Zeolites 5A and 13X and Silica Gel: Pure Components. J. Chem. Eng. Data 2009, 54, 2839−2844. (4) Drage, T. C.; Blackman, J. M.; Pevida, C.; Snape, C. E. Evaluation of Activated Carbon Adsorbents for CO2 Capture in Gasification. Energy Fuels 2009, 23, 2790−2796. (5) Du, N.; Park, H. B.; Robertson, G. P.; Dal-Cin, M. M.; Visser, T.; Scoles, L.; Guiver, M. D. Polymer nanosieve membranes for CO2capture applications. Nat. Mater. 2011, 10, 372−375. (6) Li, J. R.; Ma, Y. G.; McCarthy, M. C.; Sculley, J.; Yu, J. M.; Jeong, H. K.; Balbuena, P. B.; Zhou, H. C. Carbon dioxide capture-related gas adsorption and separation in metal-organic frameworks. Coord. Chem. Rev. 2011, 255, 1791−1823. (7) Bae, Y. S.; Snurr, R. Q. Development and Evaluation of Porous Materials for Carbon Dioxide Separation and Capture. Angew. Chem., Int. Ed. 2011, 50, 11586−11596. (8) Yang, Q. Y.; Zhong, C. L. Molecular simulation of carbon dioxide/methane/hydrogen mixture adsorption in metal-organic frameworks. J. Phys. Chem. B 2006, 110, 17776−17783. (9) 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. (10) Keskin, S.; Liu, J.; Rankin, R. B.; Johnson, J. K.; Sholl, D. S. Progress, Opportunities, and Challenges for Applying Atomically Detailed Modeling to Molecular Adsorption and Transport in MetalOrganic Framework Materials. Ind. Eng. Chem. Res. 2009, 48, 2355− 2371. (11) D’Alessandro, D. M.; Smit, B.; Long, J. R. Carbon Dioxide Capture: Prospects for New Materials. Angew. Chem., Int. Ed. 2010, 49, 6058−6082. (12) Bordiga, S.; Regli, L.; Bonino, F.; Groppo, E.; Lamberti, C.; Xiao, B.; Wheatley, P. S.; Morris, R. E.; Zecchina, A. Adsorption properties of HKUST-1 toward hydrogen and other small molecules monitored by IR. Phys. Chem. Chem. Phys. 2007, 9, 2676−2685. (13) Sumida, K.; Horike, S.; Kaye, S. S.; Herm, Z. R.; Queen, W. L.; Brown, C. M.; Grandjean, F.; Long, G. J.; Dailly, A.; Long, J. R. Hydrogen storage and carbon dioxide capture in an iron-based sodalite-type metal-organic framework (Fe-BTT) discovered via highthroughput methods. Chem. Sci. 2010, 1, 184−191. (14) Simmons, J. M.; Wu, H.; Zhou, W.; Yildirim, T. Carbon capture in metal-organic frameworks-a comparative study. Energy Environ. Sci. 2011, 4, 2177−2185. (15) Garcia, E. J.; Mowat, J. P. S.; Wright, P. A.; Perez-Pellitero, J.; Jallut, C.; Pirngruber, G. D. Role of Structure and Chemistry in Controlling Separations of CO2/CH4 and CO2/CH4/CO Mixtures over Honeycomb MOFs with Coordinatively Unsaturated Metal Sites. J. Phys. Chem. C 2012, 116, 26636−26648. (16) Dzubak, A. L.; Lin, L. C.; Kim, J.; Swisher, J. A.; Poloni, R.; Maximoff, S. N.; Smit, B.; Gagliardi, L. Ab initio carbon capture in open-site metal-organic frameworks. Nat. Chem. 2012, 4, 810−816. (17) Rana, M. K.; Koh, H. S.; Hwang, J.; Siegel, D. J. Comparing van der Waals Density Functionals for CO2 Adsorption in Metal Organic Frameworks. J. Phys. Chem. C 2012, 116, 16957−16968. (18) Liu, S.; Sun, L. X.; Xu, F.; Zhang, J.; Jiao, C. L.; Li, F.; Li, Z. B.; Wang, S.; Wang, Z. Q.; Jiang, X.; Zhou, H. Y.; Yang, L. N.; Schick, C. Nanosized Cu-MOFs induced by graphene oxide and enhanced gas storage capacity. Energy Environ. Sci. 2013, 6, 818−823. (19) McCormick, L. J.; Duyker, S. G.; Thornton, A. W.; Hawes, C. S.; Hill, M. R.; Peterson, V. K.; Batten, S. R.; Turner, D. R. Ultramicroporous MOF with High Concentration of Vacant Cu-II Sites. Chem. Mater. 2014, 26, 4640−4646. G

DOI: 10.1021/acssuschemeng.6b00080 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering Composites as CO2 Adsorbents at Room Temperature. ACS Appl. Mater. Interfaces 2013, 5, 4951−4959. (39) Koh, H. S.; Rana, M. K.; Hwang, J.; Siegel, D. J. Thermodynamic screening of metal-substituted MOFs for carbon capture. Phys. Chem. Chem. Phys. 2013, 15, 4573−4581. (40) Ye, S.; Jiang, X.; Ruan, L. W.; Liu, B.; Wang, Y. M.; Zhu, J. F.; Qiu, L. G. Post-combustion CO2 capture with the HKUST-1 and MIL-101(Cr) metal-organic frameworks: Adsorption, separation and regeneration investigations. Microporous Mesoporous Mater. 2013, 179, 191−197. (41) Raganati, F.; Gargiulo, V.; Ammendola, P.; Alfe, M.; Chirone, R. CO2 capture performance of HKUST-1 in a sound assisted fluidized bed. Chem. Eng. J. (Amsterdam, Neth.) 2014, 239, 75−86. (42) Yan, X. L.; Komarneni, S.; Zhang, Z. Q.; Yan, Z. F. Extremely enhanced CO2 uptake by HKUST-1 metal-organic framework via a simple chemical treatment. Microporous Mesoporous Mater. 2014, 183, 69−73. (43) Allen, M. P.; Tildesley, D. J. Computer Simulation of Liquides; Oxford University Press; Oxford, U. K., 1987. (44) Gupta, A.; Chempath, S.; Sanborn, M. J.; Clark, L. A.; Snurr, R. Q. Object-orientated programming paradigms for molecular modeling. Mol. Simul. 2003, 29, 29−46. (45) Deleeuw, S. W.; Perram, J. W.; Smith, E. R. Simulation of Electrostatic Systems in Periodic Boundary-Conditions 0.1. Lattice Sums and Dielectric-Constants. Proc. R. Soc. London, Ser. A 1980, 373, 27−56. (46) Potoff, J. J.; Siepmann, J. I. Vapor-liquid equilibria of mixtures containing alkanes, carbon dioxide, and nitrogen. AIChE J. 2001, 47, 1676−1682. (47) Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. Comparison of Simple Potential Functions for Simulating Liquid Water. J. Chem. Phys. 1983, 79, 926−935. (48) Mayo, S. L.; Olafson, B. D.; Goddard, W. A. Dreiding - a Generic Force-Field for Molecular Simulations. J. Phys. Chem. 1990, 94, 8897−8909. (49) Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard, W. A.; Skiff, W. M. Uff, a Full Periodic-Table Force-Field for Molecular Mechanics and Molecular-Dynamics Simulations. J. Am. Chem. Soc. 1992, 114, 10024−10035. (50) Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 11169−11186. (51) Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, 1133−1138. (52) Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 59, 1758−1775. (53) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865−3868. (54) Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188−5192. (55) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, P. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision A.08; Gaussian, Inc.: Wallingford, CT, 2009. (56) Breneman, C. M.; Wiberg, K. B. Determining Atom-Centered Monopoles from Molecular Electrostatic Potentials - the Need for

High Sampling Density in Formamide Conformational-Analysis. J. Comput. Chem. 1990, 11, 361−373.

H

DOI: 10.1021/acssuschemeng.6b00080 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX