Metal–Organic Framework Composite for CO2 Capture

Ionic Liquid/Metal–Organic Framework Composite for CO2 Capture: A Computational Investigation. Yifei Chen ... E-mail: [email protected]. ... ACS Appl...
0 downloads 0 Views 963KB Size
ARTICLE pubs.acs.org/JPCC

Ionic Liquid/MetalOrganic Framework Composite for CO2 Capture: A Computational Investigation Yifei Chen, Zhongqiao Hu, Krishna M. Gupta, and Jianwen Jiang* Department of Chemical and Biomolecular Engineering, National University of Singapore, 117576 Singapore

bS Supporting Information ABSTRACT: A composite of ionic liquid [BMIM][PF6] supported on metalorganic framework IRMOF-1 is investigated for CO2 capture by molecular computation. Due to the confinement effect, IL in the composite exhibits an ordered structure as observed from radial distribution functions. The bulky [BMIM]+ cation resides in the open pore of IRMOF-1, whereas the small [PF6] anion prefers to locate in the metal cluster corner and possesses a strong interaction with the framework. [BMIM]+ exhibits a greater mobility than [PF6], which is also observed in simulation and experimental studies for imidazolium-based ILs in the bulk phase. With increasing IL ratio in the composite and thus enhancing confinement effect, the mobility of [BMIM]+ and [PF6] is reduced. Ions in the composite interact strongly with CO2; in particular, the [PF6] anion is the most favorable site for CO2 adsorption. The composite selectively adsorbs CO2 from the CO2/N2 mixture, with selectivity significantly higher than many other supported ILs. Furthermore, the selectivity increases with increasing IL ratio in the composite. This computational study, for the first time, demonstrates that IL/MOF composite might be potentially useful for CO2 capture.

1. INTRODUCTION Currently, more than 85% of energy sources in use are fossil fuels (coal, oil, and gas).1 Combustion of fossil fuels produces a huge amount of CO2 emissions into the atmosphere, which is approximately 30 gigatons per year. Since the industrial revolution, CO2 concentration in the atmosphere has increased from 280 to 385 ppm.2 It was estimated that the atmospheric CO2 concentration would increase up to 570 ppm by 2100, which could lead to a global temperature increase by 1.9 C and sea level increase by 38 cm.3 Carbon capture and sequestration (CCS) is now not only a scientific interest but a societal issue for environmental protection. As a key step in CCS, CO2 is required to be captured from CO2 emissions. Several techniques have been proposed for CO2 capture such as amine scrubbing, cryogenic distillation, sorbent adsorption, and membrane permeation. Along with this, there has been considerable interest in using ionic liquids (ILs) for CO2 capture with reduced energy cost.4 As a unique class of green solvents, ILs are nonvolatile and nonflammable with high thermal stability and thus might potentially substitute traditional solvents.5,6 Since the first observation of high CO2 dissolution in ILs,7 a large number of experimental and theoretical studies have examined the CO2 behavior in various ILs.811 For example, Shiflett and Yokozeki measured the solubilities and diffusivities of CO2 in 1-n-butyl-3methylimidazolium hexafluorophosphate [BMIM][PF6] and 1-n-butyl-3-methylimidazolium tetrafluoroborate [BMIM][BF4].12 Bara et al. synthesized imidazolium-based ILs with one, two, or three oligo(ethylene glycol) substituents and determined the solubilities of CO2, N2, and CH4.13 Compared to the r 2011 American Chemical Society

corresponding alkyl analogues, these ILs were observed to have 3075% higher ideal solubility selectivities for CO2/N2 and CO2/CH4 mixtures. By measuring CO2 solubility in a range of ILs, Muldoon et al. found that ILs containing more fluoroalkyl chains on either the cation or the anion improve CO2 solubility when compared to less fluorinated ILs.14 From simulations, Shi and Maginn calculated the sorption of CO2 and gas mixtures in 1-n-hexyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide [HMIM][TF2N].15,16 The quantum chemical COSMORS method was applied by Gonzalez-Miquel et al. to predict CO2/N2 selectivity in 224 ILs, and they found [SCN]-based ILs can enhance selective separation.17 On the other hand, taskspecific ILs have been investigated for CO2 capture, e.g., by functionalizing ILs18 or tuning the basicity of ILs.19 There are two major issues in the use of ILs for CO2 capture, i.e., the cost and high viscosity of ILs. As a new strategy, supported ILs (SILs) have been recently proposed in which ILs impregnate the pores of a solid porous support.20 In addition to enhanced mechanical strength, SILs reduce the amount needed and viscosity of ILs and thus increase separation efficiency. Scovazzo et al. first tested polyethersulfone-supported nonhexafluorophosphate ILs for gas separation and found the ideal selectivity versus permeability ratios in these SILs were above the Robeson upper bound for polymers.21 Brennecke and co-workers demonstrated the high-temperature separation of a CO2/H2 mixture using an Received: August 30, 2011 Revised: September 30, 2011 Published: October 18, 2011 21736

dx.doi.org/10.1021/jp208361p | J. Phys. Chem. C 2011, 115, 21736–21742

The Journal of Physical Chemistry C amine-functionalized IL encapsulated on a cross-linked nylon support.22 Using both hydrophilic and hydrophobic polyvinylidene fluoride (PVDF) as supports, Neves et al. measured gas permeation in 1-n-alkyl-3-methylimidazolium-based ILs.23 Their results show that the hydrophobic PVDF-SILs are more stable and have a higher affinity for CO2. Most studies to date have used either organic polymers or inorganic zeolites as supports to prepare SILs.24,25 In the past decade, a new family of porous materials, namely, metalorganic frameworks (MOFs) have emerged.26 The vast choice of organic linkers and metal oxides allow an unlimited number of MOF structures to be tuned in a rational way. Consequently, MOFs possess a wide range of surface areas (up to 6500 m2/g) and pore sizes (up to 4.5 nm) and thus could act as versatile porous supports for SILs. Although numerous experimental and simulation studies have been reported separately for ILs811 and MOFs,2729 we are not aware of any study for MOF-supported ILs. In this study, we propose a MOF-supported IL (or IL/MOF composite) and demonstrate its application for CO2/N2 separation by a computational approach. The microscopic properties of IL confined in the MOF are also provided; this fundamental understating is indispensable for the synthesis of novel MOFs using ILs as templates30 or bridging ligands.31 Following this section, the models for IL, MOF, and IL/MOF composite are briefly described in section 2. The methods used including molecular dynamics (MD) and Monte Carlo (MC) are also introduced. In section 3, the structural and dynamic properties of IL in MOF are presented and then the adsorption behavior of a CO2/N2 mixture in the IL/MOF composite is examined. Finally, concluding remarks are summarized in section 4.

2. MODELS AND METHODS 2.1. [BMIM][PF6]. The IL considered here is 1-n-butyl-3methylimidazolium hexafluorophosphate [BMIM][PF6], which is one of the commonly studied imidazolium-based ILs. Figure S1 in the Supporting Information illustrates the molecular structure of [BMIM][PF6]. The interactions of IL molecules include bonded stretching, bending, torsional potentials, and nonbonded LennardJones (LJ) and electrostatic potentials. The potential parameters of anion [PF6] were adopted from Lopes et al.32 For cation [BMIM]+, the bonded interactions and nonbonded LJ interactions were represented by the AMBER force field.33 To estimate the atomic charges of [BMIM]+, density functional theory (DFT) calculations with the LeeYangParr correlation (B3LYP) functional were carried out using the GAUSSIAN 03 package.34 Initially [BMIM]+ was optimized at the 6-31G(d) basis set, and the electrostatic potentials were calculated at the 6-311+G(d,p) basis set. Thereafter, the atomic charges were estimated by fitting the electrostatic potentials. Table S1, Supporting Information, lists the atomic charges of [BMIM]+ and [PF6]. To validate the force field, the density of [BMIM][PF6] was estimated from NPT (constant temperature and pressure) MD simulation using GROMACS package v.4.5.3.35 The initial velocities in MD simulation were assigned according to the MaxwellBoltzmann distribution at 300 K. The equations of motion were integrated with a time step of 2 fs by the Leapfrog algorithm. Temperature was controlled by a velocity-rescaled Berendsen thermostat with a relaxation time of 0.1 ps, and pressure was maintained at 1 atm using the Berendsen algorithm. The LJ interactions were evaluated with a cutoff value of 14 Å, and the electrostatic interactions were calculated using the

ARTICLE

particle-mesh Ewald method with a grid spacing of 1.2 Å and a fourth-order interpolation. The simulation trajectories were saved with a time interval of 2 ps. Table S2, Supporting Information, gives the simulated densities of [BMIM][PF6] at five different temperatures (25, 40, 50, 60, and 70 C). Good agreement is observed between the simulated and experimental data. Though the simulations slightly underestimate experimental data, the deviations are less than 0.5%. To a certain extent, this suggests that the force field used for [BMIM][PF6] is reliable. 2.2. IRMOF-1. The MOF support in this study for the IL/ MOF composite is IRMOF-1, which is a prototype MOF. As shown in Figure S2, Supporting Information, IRMOF-1 has a cubic structure and a lattice constant of 25.832 Å.36 The formula of IRMOF-1 is Zn4O(BDC)3, where BDC is 1,4-benzenedicarboxylate. Each oxide-centered Zn4O tetrahedron is edge bridged by six carboxylate linkers resulting in an octahedral Zn4O(O2C)6 building unit, which reticulates into a three-dimensional structure. Straight channels exist in the framework with alternating size of 15 and 12 Å. The atomic charges of IRMOF-1 framework atoms were calculated from DFT based on a fragmental cluster shown in Figure S3, Supporting Information. The dangling bonds of the cluster were terminated by methyl groups. The 6-31G(d) basis set was used for all atoms except metal atoms, for which the LANL2DZ basis set was used. After DFT calculation, the atomic charges were fitted to the electrostatic potentials, as listed in Figure S3, Supporting Information. The dispersion interactions of IRMOF-1 framework atoms were represented by the LJ potential with parameters from the universal force field (UFF).37 The LorentzBerthelot combining rules were used to evaluate the cross-interaction parameters. A number of simulation studies have demonstrated that UFF can accurately describe gas adsorption in various MOFs.3840 As shown in Figure S4, Supporting Information, for CO2 adsorption in IRMOF-1,41 the agreement between simulation and experiment is fairly good. 2.3. IL/IRMOF-1 Composite and CO2 Capture. To examine the properties of IL/IRMOF-1 composite, we consider four different weight ratios of IL over IRMOF-1 (WIL/IRMOF‑1 = 0.4, 0.86, 1.27, and 1.5). At each ratio, the desired number of [BMIM][PF6] were added into IRMOF-1 and subject to energy minimization using the steepest descent method. Then, MD simulation was conducted for [BMIM][PF6] in the IRMOF-1 framework at 300 K. The LJ and electrostatic interactions were evaluated using the same manner as described above. The time step was 2 fs, and the trajectory was saved with a time interval of 2 ps. The total MD simulation duration was 20 ns, with the final 10 ns used for analysis. Figure 1 schematically illustrates an equilibrium snapshot of the IL/IRMOF-1 composite at WIL/IRMOF‑1 = 0.4. The capability of IL/IRMOF-1 for CO2 capture was evaluated by simulating adsorptive separation of CO2/N2 mixture in the composite. The grand canonical Monte Carlo (GCMC) method was used in the simulation. CO2 was represented as a three-site molecule, and its intrinsic quadrupole moment was described by a partial charge model. The partial charges on C and O atoms were qC = 0.576e and qO = 0.288e (e = 1.6022  1019), respectively. The CO bond length used was 1.18 Å, and the bond angle — OCO was 180. CO2CO2 interactions were modeled as a combination of LJ and electrostatic potentials.42 N2 was mimicked as a two-site model with the LJ potential parameters fitted to the experimental bulk properties.43 The bulk 21737

dx.doi.org/10.1021/jp208361p |J. Phys. Chem. C 2011, 115, 21736–21742

The Journal of Physical Chemistry C composition of CO2/N2 mixture was assumed to be 0.15:0.85, representing a typical flue gas. The IL/IRMOF-1 composite was treated as rigid during adsorption simulation. To evaluate the LJ interactions, a spherical cutoff of 15 Å was used with long-range corrections added. The electrostatic interactions were calculated using the Ewald sum. The real/reciprocal space partition parameter and the cutoff for reciprocal lattice vectors were chosen to

ARTICLE

be 0.2 Å1 and 8, respectively, to ensure convergence of the Ewald sum. The number of trial moves in a typical GCMC simulation was 2  107, in which the first 107 moves were used for equilibration and the subsequent 107 moves for ensemble averages. Five types of trial moves were randomly attempted in GCMC simulation: displacement, rotation, and partial regrowth at a neighboring position; complete regrowth at a new position; swap with reservoir including creation and deletion with equal probability.

3. RESULTS AND DISCUSSION 3.1. Structure and Dynamics of IL in IL/IRMOF-1. To examine the structure of IL in IL/IRMOF-1 composite, radial distribution functions g(r) were calculated by

gij ðrÞ ¼

Figure 1. [BMIM][PF6]/IRMOF-1 composite at a weight ratio WIL/IRMOF‑1 = 0.4. N, blue; C in [BMIM]+, green; P, pink; F, cyan; Zn, orange; O, red; C in IRMOF-1, gray; H, white.

Figure 2. Radial distribution functions of [BMIM]+ and [PF6] in IL/ IRMOF-1 at WIL/IRMOF‑1 = 0.4 and in the bulk phase, respectively. The solid lines are in IL/IRMOF-1, and the dash lines are in the bulk phase.

ΔNij V 4πr 2 ΔrNi Nj

ð1Þ

where r is the distance between the geometric centers-of-mass of species i and j, ΔNij is the number of species j around i within a shell from r to r + Δr, V is the volume, and Ni and Nj are the numbers of species i and j. Essentially, g(r) gives the ratio of local density at position r to averaged density in the system. Figure 2 shows the g(r) of [BMIM]+ and [PF6] in IL/IRMOF-1 composite at a weight ratio WIL/IRMOF‑1 = 0.4. For comparison, the g(r) of IL in the bulk phase are also plotted. Due to the confinement effect of the IRMOF-1 framework, IL molecules are packed more ordered in the composite. Consequently, the peaks of all ion pairs in the composite are observed to be higher than in the bulk phase. In particular, the attractive [BMIM]+[PF6] pair exhibits a pronounced peak at r = 4.4 Å. Nevertheless, the [BMIM]+[BMIM]+ or [PF6][PF6] pair has a lower peak because of an unfavorable repulsive interaction. The peak position for the [BMIM]+[BMIM]+ or [PF6][PF6] pair in the composite is slightly shifted to a greater value compared to that in the bulk phase, indicating the distance of cationcation or anionanion increases. This is because the confinement effect in the composite leads to a sharper g(r) for the cationanion pair, and thus, each cation (or anion) is surrounded by a larger number of anions (or cations). Figure 3 further shows the g(r) of [BMIM]+ and [PF6] around IRMOF-1 framework atoms (O1, O2, Zn, and C3 atoms) at WIL/IRMOF‑1 = 0.4. Nonzero g(r) is observed at r < 5 Å for [BMIM]+C3 and [BMIM]+O2 but at r > 5 Å for [BMIM]+Zn and [BMIM]+O1. This indicates [BMIM]+ is

Figure 3. Radial distribution functions of (a) [BMIM]+ and (b) [PF6] in IL/IRMOF-1 around O1, O2, Zn, and C3 atoms of IRMOF-1 at WIL/IRMOF‑1 = 0.4. 21738

dx.doi.org/10.1021/jp208361p |J. Phys. Chem. C 2011, 115, 21736–21742

The Journal of Physical Chemistry C

ARTICLE

proximal to the benzene ring and carboxylate group in IRMOF-1 rather than the metal cluster. With a chain-like structure, [BMIM]+ prefers to reside in the open pore of IRMOF-1 due to the configuration entropy effect. The small corner near the metal cluster cannot accommodate [BMIM]+. However, pronounced peaks are observed for [PF6] around all four atoms at r < 5 Å. In particular, the peaks around O1 and Zn atoms are higher, suggesting [PF6] is preferentially located in the metal cluster corner. Compared to [BMIM]+, [PF6] exhibits higher peaks around the four atoms; therefore, [PF6] possesses a stronger interaction with the framework. Note that the peaks at r = 1011 Å are due to the presence of framework atoms on the other side of the structure. The dynamics of IL is evaluated by mean-squared displacement (MSD) from MSDðtÞ ¼

1 N jΔri ðtÞj2 N i¼1



ð2Þ

where N is the number of ions and Δri(t) is the displacement of ith ion at time t. Figure 4 shows the MSDs of [BMIM]+ and [PF6] in IL/IRMOF-1 at WIL/IRMOF‑1 = 0.4, 0.86, and 1.27. As expected, the mobility of IL in the bulk phase is greater than in the composite (data not shown). In addition, the mobility of both [BMIM]+ and [PF6] reduces with increasing WIL/IRMOF‑1 in the composite, as attributed to the enhanced confinement effect. At a given WIL/IRMOF‑1, the mobility of bulky [BMIM]+ is greater than [PF6]. Such an interesting phenomenon was also observed in simulations44,45 and experiments46,47 for imidazolium-based ILs in the bulk phase and interpreted as a result of the less hindered

Figure 4. Mean-squared displacements of [BMIM]+ and [PF6] in IL/IRMOF-1 at WIL/IRMOF‑1 = 0.4, 0.86, and 1.27.

displacement of the [BMIM]+ ring along the direction of the C1 atom (see Figure S1, Supporting Information).48 In the IL/ IRMOF-1 composite under study, [PF6] interacts more strongly with the framework as discussed above; this also contributes to the observed smaller mobility of [PF6]. To provide further insight into the dynamics of IL in the IL/ IRMOF-1 composite, velocity correlation functions were calculated by CðtÞ ¼ Ævð0Þ 3 vðtÞæ

ð3Þ

where v(t) is the velocity of the center-of-mass of the cation or anion at time t. To calculate C(t), a 100 ps trajectory was generated with a small time interval of 0.1 ps after 20 ns MD simulation. Figure 5 shows the reduced velocity correlation functions C(t)/C(0) of [BMIM]+ and [PF6] in the composite. The trend at different WIL/IRMOF‑1 is largely identical. For either cation or anion, the C(t)/C(0) approaches zero in less than 2 ps and then exhibits marginal oscillation. Therefore, the velocity of [BMIM]+ or [PF6] becomes uncorrelated rapidly within a picosecond scale. This is similar to the time scale observed for IL in the bulk phase48 and reveals that confinement has an insignificant effect on the velocity correlation of IL. 3.2. CO2 Capture in IL/IRMOF-1. Adsorption of CO2/N2 mixture in IL/IRMOF-1 composite was simulated to evaluate its performance for CO2 capture. The CO2/N2 mixture was assumed to possess a composition of 15:85 to mimic a typical flue gas. Figure 6 shows the simulation snapshot of CO2/N2 mixture with a total pressure of 1000 kPa in the composite at WIL/IRMOF‑1 = 0.4. [PF6] anions are observed to preferentially locate in the metal cluster corner of IRMOF-1, which was evidenced by the g(r) in Figure 3. Compared to N2, more CO2 molecules are adsorbed in the composite and proximal to the ions. This is attributed to the strong interaction between CO2 and ions. To identify the favorable sites in the composite for CO2 adsorption, Figure 7 shows the g(r) of CO2 around the framework (Zn) and IL (N1, N2, and P) atoms. A sharp peak is observed in the g(r) of CO2P at r ≈ 4.1 Å, indicating the [PF6] anion is the most favorable site. The g(r) of CO2N1 and CO2N2 are less pronounced with broad peaks at r ≈ 5.0 and 5.2 Å, respectively. As a comparison, the g(r) of CO2Zn has the lowest peak at r ≈ 4.5 Å. Structural analysis reveals that CO2 is preferentially adsorbed onto IL, particularly the [PF6] anion, in the IL/IRMOF-1 composite. Such behavior is remarkably different from the adsorption in IRMOF-1, in which the favorable

Figure 5. Velocity correlation functions of [BMIM]+ and [PF6] in IL/IRMOF-1 at WIL/IRMOF‑1 = 0.4, 0.86, 1.27, and 1.5. 21739

dx.doi.org/10.1021/jp208361p |J. Phys. Chem. C 2011, 115, 21736–21742

The Journal of Physical Chemistry C site was found to be near the metal cluster.41 In the IL/IRMOF-1 composite, the cations and anions (especially [PF6]) occupy the metal cluster corner and act as adsorption sites for CO2 adsorption. Figure 8a shows the g(r) of CO2 around the P atom of [PF6] at 10, 100, and 1000 kPa in the IL/IRMOF-1 composite at WIL/IRMOF‑1 = 0.4. The peak in g(r) of CO2P drops with increasing pressure, whereas the coordination number of CO2 molecules around P atoms increases (data not shown). This is because the

ARTICLE

Figure 6. Simulation snapshot of CO2/N2 mixture (Ptotal = 1000 kPa) in IL/IRMOF-1 at WIL/IRMOF‑1 = 0.4.

local density of CO2 increases less rapidly than the average density. A similar behavior was also found in our previous studies for CO2 capture in rho zeolite-like MOF.49 As shown in Figure 8b, the peak in g(r) of CO2P also drops with increasing WIL/IRMOF‑1 in the IL/ IRMOF-1 composite. At a low WIL/IRMOF‑1, only a few IL molecules exist and reside closely to the IRMOF-1 framework; consequently, CO2 molecules are largely localized because of favorable ionic adsorption sites. With increasing WIL/IRMOF‑1, however, the ions start to occupy the open pores and distribute homogenously in the IRMOF-1 framework; therefore, CO2 molecules are adsorbed uniformly and the local density relative to the average density drops. Separation of the CO2/N2 mixture is quantified by selectivity Si/j = (xi/xj)(yj/yi), where xi and yi are the mole fractions of component i in the adsorbed and bulk phase, respectively. Figure 9 shows the selectivity SCO2/N2 in IL/IRMOF-1 composite at different WIL/IRMOF‑1. In the absence of IL, SCO2/N2 remains nearly 5 over the entire pressure range; however, it increases with increasing WIL/IRMOF‑1. This is because IL acts as favorable sites for CO2 adsorption and more such sites are available when WIL/IRMOF‑1 increases. Nevertheless, SCO2/N2 decreases with increasing pressure particularly at high WIL/IRMOF‑1. The reason is that most adsorption sites are occupied at high pressures and thus their number reduces. At the highest WIL/IRMOF‑1 = 1.5 under this study, SCO2/N2 reaches approximately 90 at infinite dilution and 70 at 1 bar. This value is higher than the selectivities reported for the CO2/N2 mixture in pure [BMIM][PF6] (22.6),17 in PVDF-supported [BMIM][PF6] (20),23 and in many other supported ILs (1056).24

Figure 7. Radial distribution functions of CO2 (Ptotal = 10 kPa) around Zn, N1, N2, and P atoms in IL/IRMOF-1 at WIL/IRMOF‑1 = 0.4.

Figure 9. Selectivity of CO2/N2 mixture in IL/IRMOF-1 at WIL/IRMOF‑1 = 0, 0.4, 0.86, 1.27, and 1.5.

Figure 8. Radial distribution functions of CO2 around the P atom in IL/IRMOF-1: (a) Ptotal = 10, 100, 1000 kPa and WIL/IRMOF‑1 = 0.4; (b) WIL/IRMOF‑1 = 0.4, 0.86, 1.27, 1.5 and Ptotal = 100 kPa. 21740

dx.doi.org/10.1021/jp208361p |J. Phys. Chem. C 2011, 115, 21736–21742

The Journal of Physical Chemistry C It is intriguing to examine how CO2 sorption capacity in IL/ IRMOF-1 composite changes with WIL/IRMOF-1. Figure S5 in the Supporting Information shows the isotherms of CO2 in IL/IRMOF-1 at different W IL/IRMOF-1 . Compared to pure IRMOF-1, the composite has a greater CO2 capacity at low pressures (e.g., within 0100 kPa practically for post-combustion capture), but a smaller one at high pressures. This is because the addition of IL into IRMOF-1 leads to narrow constrained pores, which enhances potential overlap and sorption at low pressures; however, the free volume in IRMOF-1 reduces upon adding IL, as reflected by the smaller capacity at high pressures. On the other hand, compared to pure IL, the composite has a substantially greater capacity. At WIL/IRMOF-1 = 1.5, CO2 capacity in the composite is comparable to that in pure IL, whereas CO2/N2 selectivity in the composite is 34 times higher.

4. CONCLUSIONS We have examined the microscopic properties of [BMIM][PF6] in IL/IRMOF-1 composite and its capability for CO2 capture. IL molecules in the composite are packed more ordered than in the bulk phase, particularly for the attractive [BMIM]+ [PF6] pair. This is attributed to the confinement effect of the nanoporous IRMOF-1 framework. The bulky chain-like [BMIM]+ cation tends to reside in the open pore of IRMOF-1. In contrast, the small [PF6] has a strong interaction with the framework and stays preferentially in the metal cluster corner. As also observed in the bulk phase, the mobility of [BMIM]+ in the composite is greater than [PF6]. Due to the enhanced confinement effect when the ratio of IL in the composite increases, the mobility of both [BMIM]+ and [PF6] reduces. Nevertheless, confinement has no significant effect on velocity correlation. It is found that the velocity of [BMIM]+ or [PF6] is essentially uncorrelated after 2 ps, which is close to the time scale for bulk ILs. In the IL/IRMOF-1 composite, CO2 is more strongly adsorbed than N2 from the CO2/N2 mixture. From structural analysis, the [PF6] anion is identified to be the most favorable site for CO2 adsorption in the composite. In IRMOF-1, however, the favorable site is in the metal cluster corner. Upon adding IL into IRMOF-1, ions, especially [PF6], occupy the corner and act as strong ionic adsorption sites for CO2. With increasing IL ratio in the composite, ions start to occupy the open pores and distribute more homogenously; thus, CO2 molecules tend to locate more uniformly. Compared to IRMOF-1, the IL/IRMOF1 composite has a substantially higher selectivity for CO2/N2 separation. With increasing IL ratio, the selectivity in the composite increases as the number of ionic adsorption sites becomes larger. At a weight ratio IL/IRMOF-1 = 1.5, the selectivity is about 70 at ambient conditions (300 K and 1 bar), which is higher than in many other supported ILs. The simulation results suggest that the IL/IRMOF-1 composite is potentially a good material for CO2 capture. In this proof-of-concept computational study, the common [BMIM][PF6] and prototype IRMOF-1 are considered. Nevertheless, thousands of ILs and MOFs have been synthesized with their readily tunable structures; therefore, new candidates will be further explored. ’ ASSOCIATED CONTENT

bS

Supporting Information. Atomic types and charges in [BMIM][PF6], densities of [BMIM][PF6], atomic types and charges in IRMOF-1, and sorption isotherm of CO2 in IRMOF-1.

ARTICLE

This material is available free of charge via the Internet at http://pubs.acs.org.

’ AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected].

’ ACKNOWLEDGMENT The authors gratefully acknowledge the National Research Foundation of Singapore (R-279-000-261-281) and the National University of Singapore (R-279-000-297-112) for financial support. ’ REFERENCES (1) Facing the Hard Truths about Energy: A Comprehensive View to 2030 of Global Oil and Natural Gas; U.S. Department of Energy: Washington, D.C., 2007. (2) Doney, S. C.; Fabry, V. J.; Feely, R. A.; Kleypas, J. A. Annu. Rev. Mar. Sci. 2009, 1, 169. (3) Stewart, C.; Hessami, M.-A. Energy Convers. Manage. 2005, 46, 403. (4) Shiflett, M. B.; Drew, D. W.; Cantini, R. A.; Yokozeki, A. Energy Fuels 2010, 24, 5781. (5) Zhu, S. D.; Wu, Y. X.; Chen, Q. M.; Yu, Z. N.; Wang, C. W.; Jin, S. W.; Ding, Y. G.; Wu, G. Green Chem. 2006, 8, 325. (6) Plechkova, N. V.; Seddon, K. R. Chem. Soc. Rev. 2008, 37, 123. (7) Blanchard, L. A.; Hancu, D.; Beckman, E. J.; Brennecke, J. F. Nature 1999, 399, 28. (8) Bara, J. E.; Camper, D. E.; Gin, D. L.; Noble, R. D. Acc. Chem. Res. 2010, 43, 152. (9) Karadas, F.; Atilhan, M.; Aparicio, S. Energy Fuels 2010, 24, 5817. (10) Hu, Y. F.; Liu, Z. C.; Xu, C. M.; Zhang, X. M. Chem. Soc. Rev. 2011, 40, 3802. (11) Hasib-ur-Rahman, M.; Siaj, M.; Larachi, F. Chem. Eng. Proc. 2010, 49, 313. (12) Shiflett, M. B.; Yokozeki, A. Ind. Eng. Chem. Res. 2005, 44, 4453. (13) Bara, J. E.; Gabriel, C. J.; Lessmann, S.; Carlisle, T. K.; FInotello, A.; Gin, D. L.; Noble, R. D. Ind. Eng. Chem. Res. 2007, 46, 5380. (14) Muldoon, M. J.; Aki, S. N. V. K.; Anderson, J. L.; Dixon, J. K.; Brennecke, J. F. J. Phys. Chem. B 2007, 111, 9001. (15) Shi, W.; Maginn, E. J. J. Phys. Chem. B 2008, 112, 2045. (16) Shi, W.; Maginn, E. J. J. Phys. Chem. B 2008, 112, 16710. (17) Gonzalez-Miquel, M.; Palomar, J.; Omar, S.; Rodriguez, F. Ind. Eng. Chem. Res. 2011, 50, 5739. (18) Bates, E. D.; Mayton, R. D.; Ntai, I.; Davis, J. H., Jr J. Am. Chem. Soc. 2002, 124, 926. (19) Wang, C.; Luo, X.; Luo, H.; Jiang, D. E.; Li, H. R.; Dai, S. Angew. Chem., Int. Ed. 2011, 50, 4918. (20) Lozano, L. J.; Godinez, C.; de los Rios, A. P.; HernandezFernandez, F. J.; Sanchez-Segado, S.; Alguacil, F. J. J. Membr. Sci. 2011, 376, 1. (21) Scovazzo, P.; Kieft, J.; Finan, D. A.; Koval, C.; DuBois, D.; Noble, R. D. J. Membr. Sci. 2004, 238, 57. (22) Myers, C.; Pennline, H.; Luebke, D.; Ilconich, J.; Dixon, J. K.; Maginn, E. J.; Brennecke, J. F. J. Membr. Sci. 2008, 322, 28. (23) Neves, L. A.; Crespo, J. G.; Coelhoso, I. M. J. Membr. Sci. 2010, 357, 160. (24) Scovazzo, P. J. Membr. Sci. 2009, 343, 199. (25) Le Bideau, J.; Viau, L.; Vioux, A. Chem. Soc. Rev. 2011, 40, 907. (26) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’Keefe, M.; Yaghi, O. M. Science 2002, 295, 469. (27) Li, J. R.; Kuppler, R. J.; Zhou, H. C. Chem. Soc. Rev. 2009, 38, 1477. 21741

dx.doi.org/10.1021/jp208361p |J. Phys. Chem. C 2011, 115, 21736–21742

The Journal of Physical Chemistry C

ARTICLE

(28) D€uren, T.; Bae, Y. S.; Snurr, R. Q. Chem. Soc. Rev. 2009, 38, 1203. (29) Jiang, J. W.; Babarao, R.; Hu, Z. Q. Chem. Soc. Rev. 2011, 40, 3599. (30) Parnham, E. R.; Morris, R. E. Acc. Chem. Res. 2007, 40, 1005. (31) Han, L.; Zhang, S.; Wang, Y.; Yan, X.; Lu, X. Inorg. Chem. 2008, 48, 786. (32) Lopes, J. N. C.; Deschamps, J.; Padua, A. A. H. J. Phys. Chem. B 2004, 108, 2038. (33) Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Gould, I. R.; Merz, K. M.; Ferguson, D. M.; Spellmeyer, D. C.; Fox, T.; Caldwell, J. W.; Kollman, P. A. J. Am. Chem. Soc. 1995, 117, 5179. (34) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Zakrzewski, V. G.; Montgomery, J. A.; Stratmann, R. E.; Burant, J. C.; Dapprich, S.; Millam, J. M.; Daniels, A. D.; Kudin, K. N.; Strain, M. C.; Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.; Mennucci, B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G. A.; Ayala, P. Y.; Cui, Q.; Morokuma, K.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Cioslowski, J.; Ortiz, J. V.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Gomperts, R.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Gonzalez, C.; Challacombe, M.; Gill, P. M. W.; Johnson, B. G.; Chen, W.; Wong, M. W.; Andres, J. L.; Head-Gordon, M.; Replogle, E. S.; Pople, J. A. GAUSSIAN 03, Revision D.01 ed.; Gaussian, Inc.: Wallingford CT, 2004. (35) Van Der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A. E.; Berendsen, H. J. C. J. Comput. Chem. 2005, 26, 1701. (36) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’Keefe, M.; Yaghi, O. M. Science 2002, 295, 469. (37) Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard, W. A.; Skiff, W. M. J. Am. Chem. Soc. 1992, 114, 10024. (38) Skoulidas, A. I.; Sholl, D. S. J. Phys. Chem. B 2005, 109, 15760. (39) Garberoglio, G.; Skoulidas, A. I.; Johnson, J. K. J. Phys. Chem. B 2005, 109, 13094. (40) Babarao, R.; Hu, Z. Q.; Jiang, J. W.; Chempath, S.; Sandler, S. I. Langmuir 2007, 23, 659. (41) Babarao, R.; Jiang, J. W. Langmuir 2008, 24, 6270. (42) Hirotani, A.; Mizukami, K.; Miura, R.; Takaba, H.; Miya, T.; Fahmi, A.; Stirling, A.; Kubo, M.; Miyamoto, A. Appl. Surf. Sci. 1997, 120, 81. (43) Murthy, C. S.; Singer, K.; Klein, M. L.; McDonald, I. R. Mol. Phys. 1980, 41, 1387. (44) Morrow, T.; Maginn, E. J. Phys. Chem. B 2002, 106, 12807. (45) Lee, S. U.; Jung, J.; Han, Y. K. Chem. Phys. Lett. 2005, 406, 332. (46) Umecky, T.; Kanakubo, M.; Ikushima, Y. Fluid Phase Equilib. 2005, 228, 329. (47) Kanakubo, M.; Harri, s. K. R.; Tsuchihashi, N.; Ibuki, K.; Ueno, M. J. Phys. Chem. B 2007, 111, 2062. (48) Urahata, S. M.; Riberiro, M. C. C. J. Chem. Phys. 2005, 122, 024511. (49) Babarao, R.; Jiang, J. W. J. Am. Chem. Soc. 2009, 131, 11417.

21742

dx.doi.org/10.1021/jp208361p |J. Phys. Chem. C 2011, 115, 21736–21742