bs Supporting Information

Similar documents
Effect of the organic functionalization of flexible MOFs on the. adsorption of CO 2

Structure-Property Relationships of Porous Materials for Carbon Dioxide Separation and Capture

Supporting Information. Co-adsorption and Separation of CO 2 -CH 4 Mixtures in. the Highly Flexible MIL-53(Cr) MOF.

Supporting Information

Molecular Insight into the Adsorption and Diffusion of Water in the Versatile Hydrophilic/Hydrophobic Flexible MIL-53(Cr) MOF

Schwarzites for Natural Gas Storage: A Grand- Canonical Monte Carlo Study

Impact of the flexible character of MIL-88 iron(iii) dicarboxylates on the

Adsorption Separations

Hydrophobic Metal-Organic Frameworks for Separation of Biofuel/Water Mixtures Introduction Methods

Opening the Gate: Framework Flexibility in ZIF-8 Explored by Experiments and Simulations

Porous Solids for Biogas Upgrading

Understanding Inflections and Steps in Carbon Dioxide Adsorption Isotherms in Metal-Organic Frameworks. Supporting Information

Supporting information

SUPPORTING INFORMATION Flexibility and Disorder in Modified-Linker MIL-53 Materials

Supplementary Information for

Supplementary Information

Screening Metal Organic Frameworks by Analysis of Transient Breakthrough of Gas Mixtures in a Fixed Bed Adsorber

Complex Adsorption of Short Linear Alkanes in the Flexible Metal-Organic-Framework MIL-53(Fe)

New Materials and Process Development for Energy-Efficient Carbon Capture in the Presence of Water Vapor

Competitive I 2 Sorption by Cu-BTC from Humid Gas Streams

Simultaneously High Gravimetric and Volumetric Gas Uptake Characteristics of the Metal Organic Framework NU-111

Supporting Information. High-throughput Computational Screening of the MOF Database for. CH 4 /H 2 Separations. Sariyer, 34450, Istanbul, Turkey

MgO-decorated carbon nanotubes for CO 2 adsorption: first principles calculations

A. Ghoufi,*, A. Subercaze, Q. Ma,, P.G. Yot, Y. Ke, I. Puente-Orench,, T. Devic, # V. Guillerm, # C. Zhong, C. Serre, # G. Feŕey, # and G.

Cleaner Car Exhausts Using Ion-Exchanged Zeolites: Insights From Atomistic Simulations

ELECTRONIC SUPPLEMENTARY INFORMATION

High Pressure Methane Adsorption on a Series of MOF-74: Molecular Simulation Study

A new tetrazolate zeolite-like framework for highly selective CO 2 /CH 4 and CO 2 /N 2 separation

Electronic Supplementary Information (ESI)

Supporting Information

RSC Advances Supporting Information

Storage of Hydrogen, Methane and Carbon Dioxide in Highly Porous Covalent Organic Frameworks for Clean Energy Applications

ADSORPTION OF XYLENE ISOMERS IN METAL ORGANIC FRAMEWORK UiO-66 BY MOLECULAR SIMULATIONS

High compressibility of a flexible Metal-Organic-Framework

Metal-Organic Frameworks and Porous Polymer Networks for Carbon Capture

Functionalized flexible MOF as filler in mixed matrix membranes for highly selective separation of CO 2 from CH 4 at elevated pressures

Theoretical comparative study on hydrogen storage of BC 3 and carbon nanotubes

On the application of consistency criteria to. calculate BET areas of micro- and mesoporous. metal-organic frameworks

Acid-functionalized UiO-66(Zr) MOFs and their evolution after intra-framework cross-linking: structural features and sorption properties.

Ethers in a Porous Metal-Organic Framework

Supporting information (SI)

Diffusion of propylene adsorbed in Na-Y and Na-ZSM5 zeolites: Neutron scattering and FTIR studies

Electronic Supplementary Information for. Non-interpenetrated IRMOF-8: synthesis, activation, and gas sorption

Modelling of Adsorption and Diffusion in Dual-Porosity Materials: Applications to Shale Gas

Cooperative Template-Directed Assembly of Mesoporous Metal-Organic Frameworks

Energetic Performances of the Metal-Organic Framework ZIF-8 by. High Pressure Water Intrusion-Extrusion Experiments

1. Materials All chemicals and solvents were purchased from Sigma Aldrich or SAMCHUN and used without further purification.

Thomas Roussel, Roland J.-M. Pellenq, Christophe Bichara. CRMC-N CNRS, Campus de Luminy, Marseille, cedex 09, France. Abstract.

Supporting Information

Supporting Information

Supporting Information

Molecular Simulation Study of CH 4 /H 2 Mixture Separations Using Metal Organic Framework Membranes and Composites

Adsorptive separation of methanol-acetone on isostructural series of. metal-organic frameworks M-BTC (M = Ti, Fe, Cu, Co, Ru, Mo): A

Supplementary Information (SI)

PORE SIZE DISTRIBUTION OF CARBON WITH DIFFERENT PROBE MOLECULES

Electronic Supporting Information (ESI) Porous Carbon Materials with Controllable Surface Area Synthsized from Metal-Organic Frameworks

Introduction. Monday, January 6, 14

Electronic Supporting information (ESI) for

FINAL PUBLISHABLE SUMMARY REPORT

Preparation of biomass derived porous carbon: Application for methane energy storage

CO 2 and CH 4 Separation by Adsorption Using Cu-BTC Metal-Organic Framework

Characterisation of Porous Hydrogen Storage Materials: Carbons, Zeolites, MOFs and PIMs

Hydrogen Adsorption and Storage on Porous Materials. School of Chemical Engineering and Advanced Materials. Newcastle University United Kingdom

Extracting organic molecules out of water using the metal-organic framework

Supplementary Information Room-Temperature Synthesis of UiO-66 and Thermal Modulation of Densities of Defect Sites

Electronic Supplementary Information

Metal organic Frameworks as Adsorbents for Hydrogen Purification and Pre-Combustion Carbon Dioxide Capture

Supporting Information

Pore size analysis of > hypothetical zeolitesw

Electronic Supplementary Information. Selective Sorption of Light Hydrocarbons on a Family of

New Journal of Chemistry Electronic Supplementary Information

Diffusion of Water and Diatomic Oxygen in Poly(3-hexylthiophene) Melt: A Molecular Dynamics Simulation Study

Microporous Carbon adsorbents with high CO 2 capacities for industrial applications

A flexible MMOF exhibiting high selectivity for CO 2 over N 2, CH 4 and other small gases. Supporting Information

DMOF-1 as a Representative MOF for SO 2 Adsorption in both Humid and Dry Conditions

Supporting information. Mechanical Properties of Microcrystalline Metal-Organic Frameworks. (MOFs) Measured by Bimodal Amplitude Modulated-Frequency

Supporting Information. From Metal-Organic Framework to Nanoporous Carbon: Toward a Very High Surface Area and Hydrogen Uptake

Chain Length Effects of Linear Alkanes in Zeolite Ferrierite. 2. Molecular Simulations 1

Molecular Dynamics Simulation on Permeation of Acetone/Nitrogen Mixed Gas

Iranian Journal of Oil & Gas Science and Technology, Vol. 2 (2013), No. 4, pp

Supplementary Information

Adsorption Processes. Ali Ahmadpour Chemical Eng. Dept. Ferdowsi University of Mashhad

Adsorption Isotherm Measurements of Gas Shales for Subsurface Temperature and Pressure Conditions

ADSORPTION IN MICROPOROUS MATERIALS: ANALYTICAL EQUATIONS FOR TYPE I ISOTHERMS AT HIGH PRESSURE

Diffusion of CO 2 in large crystals of Cu-BTC MOF

Supporting Online Material for

SUPPORTING INFORMATION

Hydrogen diffusion in potassium intercalated graphite studied by quasielastic neutron scattering

MD Thermodynamics. Lecture 12 3/26/18. Harvard SEAS AP 275 Atomistic Modeling of Materials Boris Kozinsky

Impact of phosphorylation over the encapsulation of nucleosides analogues within porous iron(iii) Metal Organic Frameworks MIL-100(Fe) nanoparticles.

Preparation of Nanofibrous Metal-Organic Framework Filters for. Efficient Air Pollution Control. Supporting Information

Investigation of Mixed Gas Sorption in Lab-Scale. Dr. Andreas Möller

Unusual Entropy of Adsorbed Methane on Zeolite Templated Carbon. Supporting Information. Part 2: Statistical Mechanical Model

Supplementary material: The origin of the measured chemical shift of 129 Xe in UiO-66 and UiO-67 revealed by DFT investigations

Edinburgh Research Explorer

Supporting Information

Permeation of Hexane Isomers across ZSM-5 Zeolite Membranes

Synthesis and catalytic properties of MIL-100(Fe), an iron(iii) carboxylate with large pores

Ion-Gated Gas Separation through Porous Graphene

Supplementary Information. Supplementary Figure 1 Synthetic routes to the organic linker H 2 ATBDC.

Transcription:

pubs.acs.org/jpcc Understanding the Thermodynamic and Kinetic Behavior of the CO 2 /CH 4 Gas Mixture within the Porous Zirconium Terephthalate UiO-66(Zr): A Joint Experimental and Modeling Approach Qingyuan Yang,,^ Andrew D. Wiersum, Herve Jobic, Vincent Guillerm, Christian Serre, Philip L. Llewellyn,, * and Guillaume Maurin, * Institut Charles Gerhardt Montpellier, UMR CNRS 5253, UM2, ENSCM, Place E. Bataillon, 34095 Montpellier cedex 05, France Laboratoire Chimie Provence, Universites Aix-Marseille I, II et III - CNRS, UMR 6264, Centre de Saint Jer^ome, 13397 Marseille, France Institut de Recherches sur la Catalyse et l Environnement de Lyon, CNRS, Universite de Lyon, 2. Av. A. Einstein, 69626 Villeurbanne, France Institut Lavoisier, UMR CNRS 8180-Universite de Versailles St Quentin en Yvelines, 45 avenue des Etats-Unis, 78035 Versailles, ^Department France of Chemical Engineering, Beijing University of Chemical Technology, Beijing, China, 100029 ) bs Supporting Information ABSTRACT: A combination of experimental (gravimetry, microcalorimetry, and quasi-elastic neutron scattering) measurements and molecular modeling was employed to understand the coadsorption of CO 2 and CH 4 in the zirconium terephthalate UiO-66(Zr) material from both the thermodynamic and kinetic points of view. It was shown that each type of molecules adsorb preferentially in two different porosities of the material, that is, while CO 2 occupy the tetrahedral cages, CH 4 are pushed to the octahedral cages. Further, a very unusual dynamic behavior was also pointed out with the slower molecule, that is, CO 2, enhancing the mobility of the fast one, that is, CH 4, that contrasts with those usually observed so far for the CO 2 /CH 4 mixture in narrow window zeolites where the molecules are most commonly diffusing independently or slowing-down the partner species. Such behavior was interpreted in light of molecular simulations that evidenced a jump type mechanism involving a tetrahedral cages octahedral cages tetrahedral cages sequence that occurs more frequently for CH 4 when in presence of CO 2. The consequences in terms of CO 2 /CH 4 selectivity and the possible use of this MOF-type material in a PSA process are then discussed. It is thus clearly emphasized that this MOF material combines several favorable features including a good selectivity, high working capacity, and potential easy regenerability that make it as a good alternative candidate of the conventional NaX Faujasite used in pressure swing adsorption. INTRODUCTION Carbon dioxide is often considered as an impurity in raw natural gas streams, where methane is the major component. Efficient removal of CO 2 for natural gas upgrading is of economic and technology importance, since the presence of CO 2 reduces its energy content and also can lead to pipeline corrosion. 1,2 It is also a great challenge to remove CO 2 from flue gas in order to reduce the greenhouse gas emissions and also to purify biogas. 3 Pressure swing adsorption (PSA) technology based on porous adsorbents is known to be one of the most efficient and affordable processes for CO 2 separation. 4 Much research effort has been carried out to identify and develop a suitable porous material with high CO 2 affinity and capacity. Although conventional zeolites such as the NaX type Faujasite are mostly adequate for CO 2 capture, 5 it is difficult to regenerate them without significant heating which leads to low productivity and great cost. Metal organic framework (MOF) hybrid porous solids have recently attracted great interest in many prospective applications including gas separation/storage due to their many fascinating features. 6 9 Stemming from the almost infinite possibility to vary their chemistry through metal center, organic linker as well as their geometric arrangement, up until now a wide variety of MOFs have been explored to examine their performance on the separation of CO 2 /CH 4 gas mixtures, both experimentally 10 15 and theoretically. 16 21 However, one of the areas in which MOFs Received: March 21, 2011 Revised: June 14, 2011 Published: June 14, 2011 r 2011 American Chemical Society 13768 dx.doi.org/10.1021/jp202633t J. Phys. Chem. C 2011, 115, 13768 13774

Figure 1. Illustration of the UiO-66(Zr) structure. (Left) Octahedral cage; (right) tetrahedral cages. Hydrogen atoms on the organic linkers were omitted for clarity. The large yellow spheres represent the void regions inside the cages. have received some concerns is in terms of their stabilities. Indeed some of the most well-known materials have been claimed to be not sufficiently stable toward humidity, 22 24 while it is an indispensable prerequisite for the industrial operations to facilitate handling and thus to reduce the cost. Once this limitation step is overpassed, screening such kinds of MOFs will become more realistic for natural gas purification. Willis et al. have meanwhile undergone a systematic study of the hydrothermal stability of MOFs, which showed that for a given linker, increasing the charge of the metal is a way to increase the hydrothermal stability of the resulting MOF. 25 With these considerations in hand, it has been shown previously that the zirconium terephthalate, denoted UiO-66(Zr) solid (UiO for University of Oslo), 26 is thermally stable up to 540 C and we further ensured that it remains unaltered upon water adsorption/ desorption cycles (see complementary characterization in the Supporting Information). This material is built up from Zr 6 O 4 - (OH) 4 (CO 2 ) 12 clusters linked by terephthalate anions (see Figure 1), leading to a three-dimensional arrangement of micropores with each centric octahedral cage surrounded by eight corner tetrahedral cages (free diameters of ca. 11 and 8 Å for the two types of cages, respectively) and connected through narrow windows (ca. 6 Å). Both the micropore range and stability upon various conditions makes this material ideal to probe its gas separation ability prior to envisage industrial applications. The present work is thus based on a combination of appropriate experimental and modeling tools to first probe the CO 2 /CH 4 separation performance of this MOF material from both the thermodynamic and kinetic points of view when compared to conventional adsorbents, that is, NaX, and further to provide a detailed analysis of the microscopic mechanism in play. MODELS AND METHODS Synthesis and Activation. UiO-66(Zr) was prepared from a large scale mixture of zirconium tetrachloride ZrCl 4, terephthalic acid HO 2 CC 6 H 4 CO 2 H, hydrochloric acid and dimethylformamide in the 25 mmol/50 mmol/50 mmol/150 ml ratio. The slurry was then introduced in a 750 ml Teflon liner and further introduced in a metallic PAAR bomb. The system was heated overnight (16 h) at 220 C. The resulting white product was filtered off, washed with DMF to remove the excess of unreacted terephthalic acid, then washed again with acetone and dried at room temperature. The sample was finally calcined at 250 C under vacuum (5 mbar) to remove the DMF from the framework. X-ray powder diffractogram (Br uker D5000, λ Cu 1.5406 Å) of the sample is in a very good agreement with the one calculated from the published structure (see Supporting Information Figure S1). As with such microporous materials, the pressure range in which the BET equation works is well below the standard (0.05 0.3) p/p range. In our case the p/p range used was 0.003 0.05 bar. We used the strict criteria suggested by Rouquerol et al. 27 and promoted by Snurr and co-workers. 28 Then, nitrogen sorption measurments (BelSorp Max apparatus) gave a BET surface area of 1067(3) m 2 /g, very similar to the theoretical accessible surface area of 1021 m 2 /g (see the calculation details in Supporting Information). Moreover, the pore volume determined by Helium adsorption experiment is around 0.40 cm 3 /g, again in excellent agreement with the simulated value (0.42 cm 3 /g) obtained by the thermodynamic method. 29 Thus, the above agreements indicate that the sample used in this work is well activated. Gas Adsorption Experiments. Prior to adsorption experiments, the samples were placed under a secondary vacuum and heated to various temperatures for 16 h. The final outgassing temperature chosen here to fully dehydroxylate the sample was 250 C. The experiments were carried out at 303 K up to a maximum pressure of 50 bar using a laboratory made gas dosing system connected to a commercial gravimetric adsorption device (Rubotherm Pr azisionsmesstechnik GmbH). 30 A step by step gas introduction mode was used. Equilibrium was assumed when the variation of weight remained below 0.03% for 20 min. A typical adsorption experiment takes ca. 24 h using around 1 g of sample. The microcalorimetry experiments were performed by means of a manometric dosing apparatus linked to the sample cell housed in a Tian-Calvet type microcalorimetrer. 31 Around 0.3 g of sample was used for these experiments. The mixture adsorption experiments were carried out using a homemade set up consisting of a commercial gravimetric device coupled with a manometric dosing system and chromatographic gas dosing device. For these experiments, around 1 g of sample was attached to the balance while a second amount of sample was also placed in the cellule to enhance the accuracy of the measurements. In the present study, the equilibrium was considered to be satisfactory when the sample weight varied by less than 80 μg over a 15 min interval. The details of the methodology used in this work is fully described in our previous work. 31 In this series of measurements, the density was used as the results obtained were adequate in terms of errors and the composition of the adsorbed phase was 13769 dx.doi.org/10.1021/jp202633t J. Phys. Chem. C 2011, 115, 13768 13774

determined from the mass balance. More details of the procedure can be found in the Supporting Information. Quasi-Elastic Neutron Scattering. The quasi-elastic neutron scattering experiments were performed on the time-of-flight spectrometer IN6, at the Institut Laue-Langevin, Grenoble, France. The incident neutron energy was taken as 3.12 mev, corresponding to a wavelength of 5.1 Å. After scattering by the sample, the neutrons are analyzed as a function of flight time and angle. The wave-vector transfer Q varies with the scattering angle and ranged from 0.25 to 1.2 Å 1. Spectra from different detectors were grouped to obtain reasonable counting statistics and to avoid the Bragg peaks of the UiO-66(Zr) framework. The timeof-flight spectra were then converted to energy spectra. The elastic energy resolution is given by a Gaussian function with a half-width at half-maximum that varied from 40 μev at small Q to 50 μev at large Q. The UiO-66(Zr) sample was contained in an aluminum container, which was connected to a gas inlet system. After measuring the scattering of the empty MOF, a loading of 4CH 4 /uc was investigated and then three concentrations of CO 2 were introduced in the cell, corresponding to 7, 12, and 14 molecules/uc In addition, to deeply understand the effect of the presence of CO 2 on the self-diffusivity of CH 4 the D s values of pure CH 4 in UiO-66(Zr) for other loadings (8.4, 10.5, and 16.5 CH 4 /uc) were also measured at 230 K. Grand Canonical Monte Carlo Simulations. Grand canonical Monte Carlo (GCMC) simulations were performed to investigate the adsorption of the single components CO 2 and CH 4 and their binary mixtures in the dehydroxylated UiO-66(Zr). These simulations were conducted at 303 K based on a model that includes electrostatic and Lennard-Jones interactions. The CO 2 molecule was treated as rigid using the widely used three point charge model 32 while CH 4 was represented by a neutral united atom model. 33 In our previous work, 34 it has been found that the framework flexibility of this material has a significant effect on the diffusion behavior of CH 4 and CO 2 gases. Thus, the framework of the dehydroxylated UiO-66(Zr) was also treated as fully flexible by implementing our newly derived fully bonded force field, allowing us to consistently study the adsorption and further the diffusion behaviors of CO 2 /CH 4 mixture. Details of the intermolecular and intramolecular parameters are provided in the Supporting Information. The LJ potential parameters for the adsorbate adsorbate and adsorbate adsorbent interactions and the DFT derived partial charges for all the atoms of the UiO- 66(Zr) framework are reported in the Supporting Information. For the simulations of pure components, molecules were involved in three types of trials: attempts (i) to displace a molecule (translation or rotation), (ii) to create a new molecule, and (iii) to delete an existing molecule. For the simulations of mixture, an attempt to exchange molecular identity was introduced as an additional type of trial to speed up the equilibrium and reduce the statistical errors. The simulation box consisted of 18 (3 3 2) unit cells for the porous material. The simulations with larger boxes showed that no finite-size effects existed using the above box. A cutoff radius of 14.0 Å was applied to the Lennard-Jones (LJ) interactions, while the long-range electrostatic interactions were handled by the Ewald summation technique. Periodic boundary conditions were applied in all three dimensions. Peng Robinson equation of state was used to convert the pressure to the corresponding fugacity used in the GCMC simulations. Furthermore, the differential adsorption enthalpy at zero coverage was directly calculated by the fluctuation theory 35 during the simulations. For each state point, GCMC simulations consisted of 2 10 7 steps to ensure the equilibration, followed by 2 10 7 steps to sample the desired thermodynamic properties. Molecular Dynamics Simulations. On the basis of the DL_POLY 2.20 simulation package, 36 molecular dynamics (MD) simulations were carried out to study the codiffusion behaviors of CO 2 /CH 4 mixture confined in the dehydroxylated UiO-66(Zr). All the MD runs were performed considering a full flexibility of the dehydroxylated UiO-66(Zr). The simulation box also consisted of 18 (3 3 2) unit cells. The calculations were performed in the canonical (NVT) ensemble at 230 K for 0, 12, 14 CO 2 molecules mixed with 4 CH 4 molecules per unit cell (uc), which are consistent with the range of loadings explored experimentally. In addition, MD simulations at higher loadings (4CH 4 +20CO 2, 4CH 4 +30CO 2, 4CH 4 +40CO 2 /uc) were also conducted to span a wider range of loading than one explored by QENS. The MD simulations were performed as follows: molecules were randomly inserted into the MOF lattice, and then relaxed using 2 10 5 NVT Monte Carlo cycles. Velocities from the Maxwell Boltzmann distribution at the required temperature were assigned to all the adsorbate molecules and the framework atoms. Further, prior to starting the production run of 2 10 7 MD steps (i.e., 20 ns), each MD system was equilibrated with 1 10 6 MD steps. The positions of each adsorbate molecule were stored every 5000 MD steps for subsequent analysis. Nose Hoover thermostat was used to maintain the constant temperature condition. It was checked that MD simulations conducted in microcanonical (NVE) ensemble lead to equivalent results. The velocity Verlet algorithm was used to integrate the Newton equations and the QUATERNION algorithm was applied for the rotational motion of the rigid linear CO 2 molecules. The long-range Coulombic interactions were evaluated by the Ewald summation method, while all the LJ interactions were calculated with a cutoff radius of 14.0 Å. The time step used in the MD simulations was taken as 1.0 fs, and periodic boundary conditions were applied in all three dimensions. The self-diffusion coefficients of the adsorbate molecules were calculated from the mean-square displacements (MSD) method using Einstein relations. In order to improve the statistics of the calculation, 5 MD independent trajectories generated from different initial configurations were sampled for each loading. RESULTS AND DISCUSSION As a first step, a combination of gravimetry measurements and grand canonical Monte Carlo (GCMC) simulations was employed to probe the coadsorption properties of the dehydroxylated UiO-66(Zr) form. The sample was synthesized on the multigrams scale and activated using slightly different conditions as those published previously. 26 Details of the different characterizations are reported in the Supporting Information. The static coadsorption experiments were carried out at 303 K using a laboratory made gas dosing system coupled with a commercial gravimetric adsorption apparatus and chromatographic gas dosing device. 31,37,38 Two mixtures were investigated with 50 and 75% mol. CO 2 in CH 4. The sample was pretreated at 523 K prior to the adsorption measurements in order to start with the dehydroxylated form of UiO-66(Zr). Figure 2 shows the soobtained gravimetric coadsorption isotherms at 303 K, which are compared with those obtained for the single components. Note that all the isotherms are expressed in absolute adsorbed amounts. For the binary mixtures, the total mass uptakes were decomposed into the contribution of each component. The concentrations of 13770 dx.doi.org/10.1021/jp202633t J. Phys. Chem. C 2011, 115, 13768 13774

Figure 2. Absolute adsorption isotherms for CO 2 (circle) and CH 4 (square) from their single gases and binary mixture at 303 K in the dehydroxylated UiO-66(Zr). Gravimetry measurements (filled symbols) and GCMC simulations (open symbols). Bulk compositions of CO 2 CH 4 : (a) 0 100 and 100 0, (b) 50 50 (error bars are indicated here), (c) 75 25. The error bars for the simulations are less than 2% for both CH 4 and CO 2 adsorbed amounts in the whole range of pressure. Regarding the experiments, the error is more important ranging from 8 to 12% for the adsorbed amounts of CO 2 at low and high pressure respectively while it exceeds 15% for CH 4 due to the small adsorbed amounts. CO 2 and CH 4 in the gas phase were deduced from the measurement of the gas phase density via an equation of state, and the composition of the adsorbed phase was then derived from the mass balance. The single gas component Langmuir isotherms (Figure 2a) present a high CO 2 uptake, ca. 212.6 cm 3 (STP)/cm 3, which is larger than those reported for NaX (185.3) and activated carbons (142.6) 2 under the same conditions. It is further shown that at low pressure the adsorbed amount of CO 2 is much higher than that of CH 4. This observation is consistent with a more energetic interaction between CO 2 and the UiO-66(Zr) surface as emphasized by the higher adsorption enthalpy at zero coverage estimated by microcalorimetry for CO 2 ( 26.2 kj/mol) compared to CH 4 ( 16.4 kj/mol). Figure 2b,c shows that the adsorbed amounts of CH 4 are very small for both investigated gas mixture compositions while the CO 2 uptake remains rather high. The resulting CO 2 /CH 4 selectivities for both mixtures are about 5.5 at low pressure. As can be seen from Supporting Information Table S5, some other MOFs, such as Cu-BTC, are interesting with higher selectivities than UiO-66(Zr), however, their structures are revealed unstable upon humidity. 23,25 The so-obtained selectivity (5 7) is comparable to or somewhat higher than those previously reported for other non modified MOFs under similar conditions, that is, MIL-53 (4 7) 11,13 and ZIF-8 (4 7), 1 which are known to be stable or sligthly altered in the presence of water, respectively. This selectivity remains obviously lower than those previously reported for the conventional NaX Faujasite type zeolite. 31,37,38 However, the significantly lower energetic interaction between CO 2 and the UiO-66(Zr) surface ( 26.2 kj/mol) compared to those in play in NaX zeolites ( 45.0 kj/ mol) 31 suggest that this MOF in addition to its high moisture stability is potentially regenerable under milder conditions. Indeed, complementary experiments (see Supporting Information) have evidenced that a full regeneration is obtained at about 120 C, which is significantly lower than the temperature required for NaX (160 C). Further, the working capacity, defined as the difference of the adsorbed amounts of CO 2 in the binary mixture between 10 and 1 bar, that is, pressures of the production and regeneration steps in the PSA process is about 3.3 and 3.6 mmol 3 g 1 (91.7 and 100.1 cm 3 (STP)/cm 3 ) for the 50 and 75% mol CO 2 in CH 4, respectively, which is almost three times higher than that observed for NaX (37.0) 31 and slightly lower than that for the water unstable Cu-BTC (123.2) 14 under similar experimental conditions. Thus, it results that the UiO-66(Zr) is promising as this material shows higher performance in terms of working capacity per volume with respect to the reference material NaX, and its selectivity is within the same range of value than other humidity resistant MOFs (see Supporting Information Table S5). Figure 2 also presents the calculated absolute isotherms for both single and mixture components using GCMC simulations. An atomistic representation of the dehydroxylated UiO-66(Zr) framework was built using a computational-assisted structure determination, as described in the Supporting Information, that was successfully employed in the past for other MOFs. 39,40 One observes in Figure 2a an excellent agreement between the experimental and calculated isotherms for both single gas components that validates the forcefields and models used. This remains also true from an energetic point of view as the simulated enthalpies for CO 2 and CH 4 of 25.3 and 18.8 kj/mol concur well with the experimental values. Further, although our predictions do not fit perfectly the experimental data for both binary mixtures, they confirm that the CO 2 adsorbed amounts for both mixtures are only slightly affected by the presence of CH 4 in the 13771 dx.doi.org/10.1021/jp202633t J. Phys. Chem. C 2011, 115, 13768 13774

gas phase whereas the CH 4 uptake drastically decreases (Figure 2b,c). The simulated selectivity of CO 2 over CH 4 as a function of the pressure is reported in Supporting Information Figure S7 (see the Supporting Information for the details of this calculation). It is found that this selectivity remains almost unchanged in the pressure range examined in this work with only a small dependence on the gas mixture composition, the so-obtained range of value comprised between 5 and 7 being consistent with the experimental ones. Beyond the fair agreement between experimental and simulated coadsorption isotherms, a further step consisted of exploring the microscopic coadsorption mechanism for the CO 2 /CH4 gas mixture in this UiO-66(Zr) system. At low pressure, both guest molecules are preferentially adsorbed in the tetrahedral cages. The resulting arrangements illustrated in Supporting Information Figure S8a do not differ from those observed for the single component adsorption. However, compared with CH 4 a larger concentration of CO 2 molecules is found to be adsorbed in the tetrahedral cages, which is consistent with the higher affinity of CO 2 in this material as discussed above. When the pressure increases, the CO 2 molecules still occupy the tetrahedral cages as they interact more strongly with the pore wall than CH 4, some molecules of this latter adsorbate being thus pushed to the octahedral cages (Supporting Information Figure S8b). Finally at high pressure (Supporting Information Figure S8c), CH 4 molecules are mainly adsorbed in the octahedral cages, while in addition to the occupation of the tetrahedral cages, CO 2 molecules also start to accumulate in these cavities due to the limited space in the tetrahedral cages. Figure 3. Comparison between experimental (crosses) and calculated (solid lines) QENS spectra obtained for CH 4 in UiO-66(Zr) perturbed by increasing loadings of CO 2 : (a) 0, (b) 7, (c) 12, (d) 14 CO 2 /uc (Q = 0.45 Å 1 ). Further, to obtain a full picture of the separation process in this MOF material, besides the investigations on the equilibrium performance of this solid discussed above, probing the dynamics of the binary mixture is of great importance. Quasi-elastic neutron scattering (QENS) measurements coupled with molecular dynamics (MD) have been revealed in the past few years as a valuable tool to follow the diffusivity of various single species in MOF type materials. 41,42 Using this dual strategy, a further step here aims at (i) determining the loading dependence of the selfdiffusivity D s for CH 4 and (ii) further addressing via a detailed analysis of both QENS spectra and MD runs, the microscopic codiffusion mechanism. Note that as far as we know, it is the first experimental investigation that probes the kinetics of binary CO 2 /CH 4 mixtures in MOFs, while only some predictions have been reported so far. 43 45 The in situ QENS measurements were performed at 230 K at the Institut Laue-Langevin, using the time-of-flight spectrometer IN6. Since the scattering from hydrogen is much larger than the one from other atoms, in CH 4 /CO 2 mixtures the scattering will be dominated by CH 4. The diffusion of the CH 4 molecules can be characterized from the shape of the spectra, the larger the broadening of the peak, the larger the diffusivity. It appears from Figure 3 that adding CO 2 to the sample increases the broadening and thus the self-diffusivity of CH 4. Figure 4a reports the D s values of methane obtained from QENS for a concentration of 4 CH 4 /uc as a function of the total mixture loading. One first notices that the D s value for the single CH 4 gas component, 2.0 10 9 m 2 /s, is comparable to those previously reported for NaY (5.3 10 9 m 2 /s at 200 K) 46 and remains within the same order of magnitude than those in NaX Faujasites (3.0 10.0 10 9 m 2 3 s 1 at 223 K). 47 under similar experimental conditions. A significant increase of D s for CH 4 is observed as the concentration of CO 2 in the mixture increases. The so-obtained trend for D s was then compared to those extracted for the pure CH 4 component to see whether this behavior is due to the increase in the total loading or to the presence of CO 2. From Figure 4a, it can be unambiguously stated that CO 2 unusually tends to enhance the diffusivity for CH 4 leading to D s values larger than those in the pure gas for the whole range of the investigated loading. This diffusion behavior deviates from previous findings reported so far for the CO 2 /CH 4 mixture in various zeolites with narrow windows such as LTA, CHA and DDR 48,49 and larger channel, that is, NaY Faujasite, 46 where the diffusivity of CH 4 remains only almost unchanged 48 or decreases 46,50 due to the presence of the CO 2. It should be noted that such observation has Figure 4. (a) QENS data for the D s of CH 4 in pure gas (red squares) and in CO 2 /CH 4 mixture (blue circles) as a function of the total loading at 230 K, (b) D s of CH 4 as function of the CO 2 concentration at 230 K; QENS (full circles), MD (empty circles). The error bars are indicated in the figures. 13772 dx.doi.org/10.1021/jp202633t J. Phys. Chem. C 2011, 115, 13768 13774

Figure 5. Typical illustrations of the global diffusion mechanisms in the dehydroxylated UiO-66(Zr) following one targeted CH 4 molecule in the case of (a) 4 CH 4 /uc in absence of CO 2, (b) 4 CH 4 /uc in presence of 12 CO 2 molecules/uc. The positions from 1 to 6 correspond to jump sequences of CH 4 observed during the MD trajectories. The average diffusion times for CH 4 to pass from tetrahedral to octahedral cages (for example, from 1 to 2) are (a) 230 ps and (b) 100 ps, respectively. already been reported for the codiffusion of para- and ortho-xylene mixtures in the CIT-1 zeolites with two different channels. 51 MD simulations were further performed to gain some molecular insight into this unusual diffusivity profile and the resulting microscopic codiffusion mechanism. Recent studies in MOFs have shown that the framework flexibility has a profound influence on the diffusion of guest molecules confined in MOFs. 41,52,53 Indeed, all the MD simulations in this work were also performed by considering a full flexibility of the dehydroxylated UiO-66(Zr). As shown in Figure 4b, within the range of CO 2 loadings examined experimentally, the simulations also lead to an increasing profile while they underestimate the QENS data. In addition, we have also simulated the D s values for CO 2 in both single gas and in the same range of gas mixture compositions (see Supporting Information Figure S10). It is found that CO 2 diffuses about 5 times slower than CH 4 when in single gas, which is two times lower than those previously reported for NaY ( 10) 44 consistent with a higher selectivity in the Faujasite type zeolite. However, here more interestingly this difference remains almost the same in the gas mixture. One can further mention a rather unusual dynamic behavior in this UiO-66(Zr) material where the slower molecule enhances the mobility of the fast one. This result differs from what has been predicted so far for the diffusion of CO 2 / CH 4 gas mixture in narrow window type zeolites which show that either (i) the molecules are diffusing independently leading to the absence of mutual speeding-up or slowing-down the partner species 48,49 or (ii) the faster CO 2 molecules retard the slowly diffusion CH 4 species. 50 To shed some light into the microscopic diffusion mechanism and further interpret the D s profile, the MD trajectories were carefully analyzed. This highlights that the dynamics of both species mainly involve jumps following the sequence: tetrahedral cages octahedral cages tetrahedral cages. Figure 5 provides an illustration of the global diffusion mechanism for CH 4 that remains the same whether CO 2 is present or not. It has been further shown that even at low CO 2 concentrations, the CH 4 molecules are more frequently pushed into the octahedral cages than in the single component diffusion. Such a situation explains the faster diffusivity for CH 4 as these molecules spend less time in the tetrahedral cages, where the stronger interactions with the pore wall occur. This observation can be rationalized by the plot of the residence time for CH 4 within the tetrahedral cages as a function of the CO 2 content which is reported in Supporting Information Figure S11. It is clearly stated that in the range of the experimentally explored CO 2 loading, this residence time is much shorter than in the absence of CO 2. This effect contributes to a faster diffusivity for CH 4 in the range of CO 2 concentration below 14 molecules/uc. This conclusion is illustrated in Figure 5 that reports the trajectories followed by CH 4 in the single component and binary mixture. It is shown that the time required for CH 4 to pass from a tetrahedral to octahedral cages is shorter when CO 2 is present. Further, the CO 2 molecules spend more time in the tetrahedral cages as they are more strongly interacting with the pore wall, thus explaining their slower diffusivity compared to CH 4. MD simulations were also carried out beyond the CO 2 loadings examined by QENS measurements. The results shown in Figure 4 suggest that D s for CH 4 passes through a maximum for a loading of 14 CO 2 /uc before decreasing when the CO 2 concentration increases due to the steric hindrance which further drastically limits the number of jumps above-mentioned. CONCLUSIONS Our joint experimental modeling strategy shows that the UiO-66(Zr) solid can be very promising for CO 2 /CH 4 gas mixture separation with a good selectivity, very high working capacity and low cost regenerability combined to its stability under various conditions. One would expect that grafting polar functional groups on the organic linkers will enhance the affinity of the CO 2 leading to an improvement of the selectivity that would make this material as a solid alternative solution to the conventional NaX adsorbent used in PSA applications. From a fundamental point of view, both coadsorption and codiffusion mechanisms have been elucidated, revealing a very unusual dynamic behavior with the slower molecule, that is, CO 2, enhancing the mobility of the fast one, that is, CH 4 that strongly differs with what has been previously reported for CO 2 /CH 4 mixture in various narrow windows zeolites such as LTA and DDR. This unusual behavior should be also expected for this mixture diffusion in zeolites with two types of microporous voids. ASSOCIATED CONTENT b S Supporting Information. Synthesis route, experimental coadsorption/codiffusion procedures, and details of the molecular simulations. This material is available free of charge via the Internet at http://pubs.acs.org. 13773 dx.doi.org/10.1021/jp202633t J. Phys. Chem. C 2011, 115, 13768 13774

AUTHOR INFORMATION Corresponding Author *E-mail: (G.M.) guillaume.maurin@univ-montp2.fr; (P.L.L.) philip.llewellyn@univ-provence.fr. ACKNOWLEDGMENT The research leading to these results has received funding from the European Community s Seventh Framework Programme (FP7/2007-2013) under Grant Agreement 228862. We thank Dr. M. M. Koza for his help during the measurements on the IN6 spectrometer at the Institut Laue Langevin, Grenoble, France. REFERENCES (1) Venna, S. R.; Carreon, M. J. Am. Chem. Soc. 2010, 132, 76 78. (2) Millward, A. R.; Yaghi, O. M. J. Am. Chem. Soc. 2005, 127, 17998 17999. (3) D Alessandro, D. M.; Smit, B.; Long, J. R. Angew. Chem., Int. Ed. 2010, 49, 6058 6082. (4) Hernandez-Maldonado, A. J.; Yang, R. T. AIChE J. 2004, 50, 791 801. (5) Caskey, S. R.; Wong-Foy, A. G.; Matzger, A. J. J. Am. Chem. Soc. 2008, 130, 10870 10871. (6) Deng, H.; Doonan, C. J.; Furukawa, H.; Ferreira, R. B.; Towne, J.; Knobler, C. B.; Wang, B.; Yaghi, O. M. Science 2010, 327, 846 850. (7) Sumida, K.; Hill, M. R.; Horike, S.; Dailly, A.; Long, J. R. J. Am. Chem. Soc. 2009, 131, 15120 15121. (8) Li, J. R.; Kuppler, R. J.; Zhou, H. C. Chem. Soc. Rev. 2009, 38, 1477 1504. (9) Ferey, G.; Serre, C.; Devic, T.; Maurin, G.; Jobic, H.; Llewellyn, P. L.; De Weireld, G.; Vimont, A.; Daturi, M.; Chang, J. S. Chem. Soc. Rev. 2011, 40, 550 562. (10) Couck, S.; Denayer, J. F. M.; Baron, G. V.; Remy, T.; Gascon, J.; Kapteijn, F. J. Am. Chem. Soc. 2009, 131, 6326 6327. (11) Finsy, V.; Ma, L.; Alaerts, L.; De Vos, D. E.; Baron, G. V.; Denayer, J. F. M. Microporous Mesoporous Mater. 2009, 120, 221 227. (12) Xue, M.; Zhang, Z. J.; Xiang, S. C.; Jin, Z.; Liang, C. D.; Zhu, G.; Qiu, S.; Chen, B. L. J. Mater. Chem. 2010, 20, 3984 3988. (13) Hamon, L.; Llewellyn, P. L.; Devic, T.; Ghoufi, A.; Clet, G.; Guillerm, V.; Pirngruber, G. D.; Maurin, G.; Serre, C.; Driver, G.; van Beek, W.; Jolima^itre, E.; Vimont, A.; Daturi, M.; Ferey, G. J. Am. Chem. Soc. 2009, 131, 17490 17499. (14) Hamon, L.; Jolima^itre, E.; Pirngruber, G. D. Ind. Eng. Chem. Res. 2010, 49, 7497 7503. (15) Banerjee, R.; Furukawa, H.; Britt, D.; Knobler, C.; O Keeffe, M.; Yaghi, O. M. J. Am. Chem. Soc. 2009, 131, 3875 3877. (16) Yazaydin, A. O.; Snurr, R. Q.; Park, T. H.; Koh, K.; Liu, J.; LeVan, M. D.; Benin, A. I.; Jakubczak, P.; Lanuza, M.; Galloway, D. B.; Low, J. J.; Willis, R. R. J. Am. Chem. Soc. 2009, 131, 18198 18199. (17) Keskin, S.; Sholl, D. S. Langmuir 2009, 25, 11786 11795. (18) Bae, Y.-S.; Spokoyny, A. M.; Farha, O. M.; Snurr, R. Q.; Hupp, J. T.; Mirkin, C. Chem. Commun 2010, 46, 3478 3480. (19) Yang, Q.; Zhong, C. J. Phys. Chem. B 2006, 110, 17776 17783. (20) Babarao, R.; Jiang, J. W.; Sandler, S. I. Langmuir 2009, 25, 5239 5247. (21) Martin-Calvo, A.; Garcia-Perez, E.; Castillo, J. M.; Calero, S. Phys. Chem. Chem. Phys. 2008, 10, 7085 7091. (22) Li, Y. W.; Yang, R. T. Langmuir 2007, 23, 12937 12944. (23) Liang, Z. J.; Marshall, M.; Chaffee, A. L. Energy Fuels 2009, 23, 2785 2789. (24) Greathouse, J. A.; Allendorf, M. D. J. Am. Chem. Soc. 2006, 128, 10678 10679. (25) Low, J. J.; Benin, A. I.; Jakubczak, P.; Abrahamian, J. F.; Faheem, S. A.; Willis, R. R. J. Am. Chem. Soc. 2009, 131, 15834 15842. (26) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, K. P. A. J. Am. Chem. Soc. 2008, 130, 13850. (27) Rouquerol, F.; Rouquerol, J.; Sing, K. Adsorption by Powders and Porous Solids: Principles, Methodology and Applications; Academic Press: New York, 1999. (28) D uren, T.; Millange, F.; Ferey, G.; Walton, K. S.; Snurr, R. Q. J. Phys. Chem. C 2007, 111, 15350 15356. (29) Myers, A. L.; Monson, P. A. Langmuir 2002, 18, 10261 10273. (30) De Weireld, G.; Frere, M.; Jadot, R. Meas. Sci. Technol. 1999, 10, 117 126. (31) Ghoufi, A.; Gaberova, L.; Rouquerol, J.; Vincent, D.; Llewllyn, P. L.; Maurin, G. Microporous Mesoporous Mater. 2009, 119, 117 128. (32) Harris, J. G.; Yung, K. J. Phys. Chem. 1995, 99, 12021 12024. (33) Martin, M. G.; Siepmann, J. I. J. Phys. Chem. B 1998, 102, 2569 2577. (34) Yang, Q.; Jobic, H.; Salles, F.; Kolokolov, D.; Guillerm, V.; Serre, C.; Maurin, G. Chem. Eur. J. 2011; DOI: 10.1002/chem.201003596. (35) Do, D. D.; Do, H. D. J. Phys. Chem. B 2006, 110, 17531 17538. (36) Smith, W.; Forester, T. R. J. Mol. Graph. 1996, 14, 136 141. (37) Llewellyn, P. L.; Maurin, G. Chimie 2005, 8, 283 302. (38) Maurin, G.; Llewellyn, P. L.; Bell, R. G. J. Phys. Chem. B 2005, 109, 16084 16091. (39) Salles, F.; Maurin, G.; Serre, C.; Llewellyn, P. L.; Kn ofel, C.; Choi, H. J.; Filinchuk, Y ; Oliviero, L.; Vimont, A.; Long, J. R.; Ferey, G. J. Am. Chem. Soc. 2010, 132, 13782 13788. (40) Devic, T.; Horcajada, P.; Serre, C.; Salles, F.; Maurin, G.; Moulin, B.; Heurtaux, D.; Clet, G.; Vimont, A.; Greneche, J.-M.; Ouay, B. L.; Moreau, F.; Magnier, E.; Filinchuk, Y.; Marrot, J.; Lavalley, J.-c.; Daturi, M.; Ferey, G. J. Am. Chem. Soc. 2010, 132, 1127 1136. (41) Salles, F.; Jobic, H.; Ghoufi, A.; Llewellyn, P. L.; Serre, C.; Bourrelly, S.; Ferey, G.; Maurin, G. Angew. Chem., Int. Ed. 2009, 121, 8485 8489. (42) Rosenbach, N., Jr.; Jobic, H.; Ghoufi, A.; Salles, F.; Maurin, G.; Bourrelly, S.; Llewellyn, P. L.; Devic, T.; Serre, C.; Ferey, G. Angew. Chem., Int. Ed. 2008, 47, 6611 6615. (43) Babarao, R.; Jiang, J. W. Langmuir 2008, 24, 5474 5484. (44) Krishna, R. J. Phys. Chem. C 2009, 113, 19756 1971. (45) Watanabe, T.; Keskin, S.; Nair, S.; Sholl, D. S. Phys. Chem. Chem. Phys. 2009, 11, 11389 11394. (46) Deroche, I.; Maurin, G.; Borah, B. J.; Yashonath, S.; Jobic, H. J. Phys. Chem. C 2010, 114, 5027 5034. (47) K arger, J.; Pfeifer, H.; Rauscher, M.; Walter, A. S. J. Chem. Soc., Faraday I 1980, 76, 717 737. (48) Krishna, R.; van Baten, J. M. Insights into Diffusion of Gases in Zeolites Gained from Molecular Dynamics Simulations. Microporous Mesoporous Mater. 2008, 109, 91 108. (49) Krishna, R.; van Baten, J. M.; Garcia-Perez, E.; Calero, S. Chem. Phys. Lett. 2006, 429, 219 224. (50) Krishna, R.; van Baten, J. M. Sep. Purif. Technol. 2008, 61, 414 423. (51) Sastre, G.; Corma, A.; Catlow, R. A. Top. Catal. 1999, 9, 215 224. (52) Greathouse, J. A.; Allendorf, M. D. J. Phys. Chem. C 2008, 112, 5795 5802. (53) Seehamart, K.; Nanok, T.; Krishna, R.; van Baten, J. M.; Remsungnen, T.; Fritzsche, S. Microporous Mesoporous Mater. 2009, 125, 97 100. 13774 dx.doi.org/10.1021/jp202633t J. Phys. Chem. C 2011, 115, 13768 13774