Unravelling the Water Adsorption Mechanism in the

14 hours ago - In this context, a full understanding of the water adsorption mechanism in the most promising MOFs containing coordinated unsaturated s...
0 downloads 0 Views 4MB Size
Subscriber access provided by Nottingham Trent University

C: Surfaces, Interfaces, Porous Materials, and Catalysis

Unravelling the Water Adsorption Mechanism in the Mesoporous MIL-100(Fe) MOF Paulo G.M. Mileo, Kyung Ho Cho, Jaedeuk Park, Sabine Devautour-Vinot, Jong-San Chang, and Guillaume Maurin J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.9b06228 • Publication Date (Web): 30 Aug 2019 Downloaded from pubs.acs.org on August 30, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Unravelling the water adsorption mechanism in the mesoporous MIL-100(Fe) MOF Paulo G. M. Mileo,a Kyung Ho Cho,b Jaedeuk Park,b Sabine Devautour-Vinot,a Jong-San Changb,c * and G. Maurina * a

Institut Charles Gerhardt Montpellier, UMR-5253, Université de Montpellier, CNRS, ENSCM, Place E. Bataillon, 34095 Montpellier cedex 05, France.

b

Research Group for Nanocatalyst and Chemical Safety Research Center, Korea Research Institute of Chemical Technology (KRICT), Gajeong-ro 141, Yuseong, Daejeon 34114, South Korea. c

Department of Chemistry, Sungkyunkwan University, Suwon 440-476, South Korea

*Corresponding authors: [email protected] and [email protected] .

KEYWORDS. Metal-Organic Frameworks, Open metal sites, Water Adsorption, MIL100(Fe), Molecular simulations, Microscopic adsorption mechanism.

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 36

ABSTRACT. Adsorption-based heat transfer (AHT) devices are promising alternatives for green energy production and (re)usage, however they are still limited by the low performance of their benchmark adsorbent materials. Metal-organic frameworks (MOFs) have been ranked amongst the most promising water adsorbents for this application owing to their potential superior water uptake and moderate hydrophilicity. However, there is still a need to rationalize and understand at the microscopic scale the water adsorption performances of this family of materials to further guide the selection of the next generation water adsorbents. In this context, a full understanding of the water adsorption mechanism in the most promising MOFs containing coordinated unsaturated sites is still highly challenging. Here, we explore the water adsorption in the mesoporous MOF MIL-100(Fe) containing coordinated unsaturated Fe(III) sites by combining advanced modeling and experimental tools. As a first stage, Density Functional Theory calculations are performed to derive an accurate force field to describe the specific interactions between water and the coordinated unsaturated Fe(III) sites. This force field is further implemented in a Grand Canonical Monte Carlo scheme to simulate the water adsorption isotherm and enthalpy in the whole range of relative pressure. A validation of the microscopic models and force field parameters is gained from a very good agreement between the experimental and simulated water adsorption data. As a further step, we provide an unprecedented description of the water adsorption microscopic mechanism in this very promising AHT water-adsorbent by a careful analysis of the MIL-100(Fe)/H2O interactions at low and intermediate relative pressure as well as the hydrogen bond network and cluster formation at higher relative pressure.

ACS Paragon Plus Environment

2

Page 3 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

INTRODUCTION

Alternative and eco-friendly technologies for heating and cooling operations are currently considered as a response to the worldwide increase in energy demand. In comparison with the standard mechanical heat pump systems, adsorption-based heat pumps (AHP) show several advantages, e.g. the use of working fluids (water and alcohols) of lower toxicity versus the conventional halogenated refrigerants and the involvement of low-grade energy sources (e.g. solar, geothermal, industrial waste heat, etc.) for regeneration1,2. The selection of the optimal water adsorbents in such applications depends on a number of variables including the desired driving and cooling temperatures, the water uptake and the ability to release the heats of adsorption and condensation. In this context besides the silico-alumino phosphate SAPO-343,4, a series of MOFs including ISE-15, MIL-1006–8, MIL-1019,10, HKUST-111,12, CAU-1013,14, MIL16015,16, NH2-MIL-12517,18, Al-Fumarate19–21, CPO-2722, NH2-UiO-6618,23, MOF-80124, MOF80824,25, Co2Cl2(BTDD)26, Zn(NDI-X)27, Y-shp-MOF-528, MFU-4l29, MIL-53(Al)-TDC30, Crsoc-MOF-131, (Ni,Co)2Cl2(BBTA)32 and more recently UiO-66-NH3Cl33, MIP-20034, DUT67(Zr)35, CUK-136, Ni2(F,Cl,Br,OH)2BTDD37, CAU-2338 have been considered as promising water-adsorbents to be integrated in AHP systems using water as working fluid. Indeed, the unique chemical versatility in terms of metal centers and organic linkers, as well as pore size/shape richness of this family of materials offers a unique opportunity to design solids with a wide spectrum of water adsorption behaviours from hydrophobic to mildly or highly hydrophilic adsorbents. Such control of the hydrophilicity/hydrophobicity nature of the MOFs that governs the pressure range of the water adsorption and indeed the driving temperature of the AHP process was also achieved by the consideration of post-synthetic strategies, such as the ligand functionalization39–41 or the introduction of defects42–44 Moreover, some mesoporous

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 36

MOFs8,10,26,31 revealed high water uptake which favours a gain in terms of working capacity at each AHP charge/discharge cycle.

However, in order to optimize the performance of MOFs for AHP applications, besides the rationale analysis that has been very recently performed on the behaviour of a series of MOFs with alcohol as working fluid45 there is still a need to understand the key parameters that control the water adsorption in this family of materials at the atomistic scale. This is a challenging objective considering first the difficulty to model the water molecule itself as indicated by the large variety of force fields available in the literature46–51, that have been considered to describe its properties (e.g. polarization, boiling and freezing temperatures, etc). In addition, capturing accurately the behaviour of water confined in porous materials like MOFs represents an even higher degree of complexity. In this context, we have contributed to gain insight into the water adsorption mechanisms in a series of MOFs, including MIL-5352, UiO-66-(CO2H)253,54, Alfumarate19, CAU-10 derivatives55, MIL-16015, MIP-20034, KAUST-7’56, and more recently CUK-1 36.

As a follow up of this quest, here we aim to scrutinize the microscopic water adsorption behaviour of MIL-100(Fe) (MIL stands for Materials of Institut Lavoisier) which has been envisaged as one of the most promising MOFs for AHP applications owing to its stability to water, large water uptake and relatively high Coefficient of Performances (COP)6,7,57. This mesoporous MOF is built from the assembly of trimers of Fe(III) octahedra and trimesate carboxylates that leads to the formation of supertetrahedra of ca. 9.5 Å. These supertetrahedra further assemble themselves into a MTN zeolitic architecture delimiting two mesoporous cages of 25 Å and 29 Å free apertures accessible through pentagonal and hexagonal windows of 5.5

ACS Paragon Plus Environment

4

Page 5 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

and 8.6 Å respectively8 (Figure 1). In its pristine form, each inorganic trimer of MIL-100(Fe) contains 2 Fe(III) coordinated to H2O and 1 Fe(III) bounded to 1 terminal –F or -OH group. The activation of the material allows the evacuation of the coordinated water leading to 2 Fe(III) coordinated unsaturated sites (CUS)58. Modelling of the adsorption in this MOF is rather challenging because of its large unit cell volume (380 000 Å3) and more importantly the presence of Fe(III)-CUS that requires a special attention to accurately capture the strength of interactions with the guests via an adequate description of the corresponding interatomic potential.

Considerable computational effort has been devoted to explore the adsorption of guest molecules to diverse CUS present in MOFs 59–82. Quantum mechanical (QM)-based approach has been used to calculate the host/guest interaction energy curve that was further fitted using analytical expressions to be further implemented in force field-based simulations. Two QMstrategies have been mostly employed based on the consideration of either the periodic structure61,66,68,71,76–79,81,82 or representative clusters.

62–64,69,70,72–75

While cluster calculations

enable the use of more sophisticated QM methods, they assume that the surrounding atoms and the topology of the framework are not expected to contribute much to the guest/CUS interaction energy. Alternatively, periodic calculations take into account the effect of the whole framework and typically avoid basis set superposition errors (BSSE) with the use of plane waves59. However, they are time consuming since they require a large number of initial guest configurations to get an accurate evaluation of the interaction energy. Here, QM calculations were employed using a representative fragment of the framework conjugated with a water molecule. The host/guest interaction energy was calculated from this system using plane waves, a methodology already reported in the literature for other CUS MOFs/guest systems69,70,72 .

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

A few attempts have been already reported in the literature

Page 6 of 36

83–86

to computationally

explore the thermodynamics and/or kinetics of water adsorption in MIL-100(Fe) via Monte Carlo and Molecular Dynamics simulations. However none of them were able to capture the adsorption behavior of water in the whole range of relative pressure and more especially in the low domain of pressure since these previous works considered generic force fields that are known to dramatically fail to represent the interactions between CUS sites and adsorbate molecules. As a leap-forward, here quantum-calculations were first considered to derive a specific force field to accurately describe the interactions between the Fe(III)-CUS and H2O. This new set of potential parameters was further implemented in a Monte Carlo scheme to provide an unprecedented full microscopic interpretation of the multi-steps present in the experimental water adsorption isotherm collected on a well-activated MIL-100(Fe) sample by gravimetry measurements.

ACS Paragon Plus Environment

6

Page 7 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

a)

b)

d)

c)

Figure 1 – (a) Illustration of the 3D-crystal structure of MIL-100(Fe) integrating (b) supertetrahedra formed by the association of a Fe3O cluster – containing 2 Fe(III)-CUS and typically 1 terminal OH group – with trimesate ligands, leading to the presence of two cages of 25 Å (c) and 29 Å (d) of free pore aperture accessible via hexagonal and pentagonal windows respectively.

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 36

METHODS 1. Molecular simulations Quantum calculations and derivation of force field parameters The MIL-100(Fe) framework was modelled using a crystal structure previously elucidated by X-ray diffraction8 integrating 1 counter-anion per Fe3O trimer. It has been previously demonstrated that the nature of the counter anions (-OH, -F, -SO4,…) has only a tiny impact on the overall adsorption behaviour of this MOF.7 Therefore, we selected arbitrarily the incorporation of –OH groups in MIL-100(Fe). A cluster containing the inorganic Fe3O node saturated by formate (Figure 2) was further cleaved from the crystal structure and further used to calculate at the DFT level the interaction energy for H2O as a function of the distance separating this molecule and the Fe(III) CUS. It was previously demonstrated that the consideration of such a cluster is good enough to accurately capture the interaction energy between a small molecule and a CUS present in MOFs.63,73 The geometry optimization of this cluster containing 1 H2O adsorbed molecule was first performed using the Quantum Espresso package87,88 with plane wave functions as basis set and the GGA PBE functional89. Dispersion corrections were considered using the DFT-D2 scheme90. The ion cores were described by Vanderbilt ultrasoft pseudopotentials91 with a kinetic energy cutoff of 60 Ry and a density cutoff of 600 Ry. A gamma centred k-sampling of the reciprocal space was performed with a Gaussian smearing of 0.005 Ry broadening. The ground spin states of the central metallic atoms in these clusters were checked, showing that all Fe(III) are high spin (sextet) with a lowest energy spin configuration of (+5,-5,+5), thus leading to a total spin value of the cluster S = 5 (sextet). The Fe(III) CUS-H2O interaction potential curve was then constructed from single-point energy calculations performed at different Fe-H2O distances keeping the same orientation of water with respect to the cluster.

ACS Paragon Plus Environment

8

Page 9 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The interaction energy was calculated using the following expression: EINT = EC+H2O – (EC + EH2O), where EC+H2O corresponds to the energy of the optimized water-loaded cluster while EC and EH2O are the energies of the cluster model and the molecule considered as single species respectively.

Figure 2 – Cluster used for the parameterization of the Fe(III) CUS – H2O interactions. In blue, red, grey and white are represented respectively Fe, O, C and H atoms. The interactions between H2O and all the atoms of the MIL-100(Fe) framework with the exception of the Fe(III) CUS were represented by 12-6 Lennard-Jones (LJ) and Coulombic potentials to respectively model the van der Waals and electrostatic interactions (Equation 1). The 12-6 LJ parameters were taken from the generic force fields DREIDING92 and UFF93 for the atoms present in the organic and inorganic nodes respectively while the partial charges for all atoms of the MOF framework were calculated using the ESP fitting approach94. To do so, two representative clusters (Figure 3) of the inorganic and organic nodes of the MOF were cleaved from the optimized framework and saturated with hydrogen atoms in order to calculate the partial atomic charges in MIL-100(Fe) at the DFT level employing the GGA/PBE89 functional

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 36

combined with a double numerical basis set containing polarization functions (DNP), as implemented in the Dmol3 module95. The corresponding partial charges are reported in Table 1.

a)

b) OOH H OH O3 Fe2

C1 C2 C3

O1

H1

Fe1 O2

Figure 3 – Selected cluster models representing the atoms of the inorganic node (Fe1, Fe2, O1, O2, O3, OOH, and HOH) (a) and the organic linker (C1, C2, C3, H1) (b) considered in the calculation of the ESP partial charges in the OH- version of the MIL-100(Fe). In blue, red, grey and white are represented respectively Fe, O, C and H atoms. Table 1 – Atomic partial charges considered for each atom type of MIL-100(Fe) defined in Figure 3. Atom type

q (e)

Fe1

1.3872

Fe2

1.2487

C1

0.5450

C2

0.1387

C3

-0.2152

O1

-0.7023

O2

-0.5912

O3

-0.5236

OOH

-0.6673

HOH

0.3767

H1

0.1638

ACS Paragon Plus Environment

10

Page 11 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Water was described by two four-site representations, i.e. the TIP4P/200549 and TIP4P-Ew48 models corresponding to one LJ site centred on the oxygen atom and three charged sites. The interactions between H2O and Fe(III) CUS were described by a Buckingham potential and a Coulombic term (Equation 2). The parameterization of the repulsion (A and B) and dispersive (C) coefficients of the Buckingham potential was carried out by fitting the DFT-calculated Fe(III) CUS– H2O interaction energy (EFe-H2O). This curve was obtained by subtracting to the quantum-derived total interaction energy (EINT), the 12-6 LJ (ELJ) and Coulombic (EQ) contributions between the H2O molecules and the rest of the atoms of the cluster, i.e. EFe–H2O = EINT – ELJ - EQ. The charges obtained for the cluster atoms were the same applied to the rest of the framework with a slight change for the charges of the terminal hydrogen atoms in order to maintain the charge neutrality of the cluster. The overall interatomic potential to describe the H2O/MIL-100(Fe) interaction corresponds to the following analytical formula:

𝐸 𝑟

⎧ ⎪ ⎨ ⎪ ⎩

,

,

1 𝑞𝑞 4𝜋𝜀 𝑟

4𝜀

1 𝑞𝑞 4𝜋𝜀 𝑟

𝐴𝑒

𝜎 𝑟

𝜎 𝑟 C 𝑟

,𝑖

,𝑖

𝐹𝑒 𝐼𝐼𝐼 𝐶𝑈𝑆

𝐹𝑒 𝐼𝐼𝐼 𝐶𝑈𝑆

1 2

where rij stands for the distance separating the atoms of MIL-100(Fe) and water (i and j respectively), εij and σij for the crossed 12-6 LJ parameters between these sites and (qi, qj) the partial charges of centres i and j. Monte Carlo simulations Grand Canonical Monte Carlo (GCMC) simulations were carried out at 298 K to predict the adsorption isotherms and to identify the most preferential sitting sites/interactions of H2O and the MOF pore wall. These calculations were performed using a simulation box made of a single unit

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 36

cell of MIL-100(Fe). The adsorption enthalpy at low coverage was computed using the revised Widom’s test particle insertion method.96 The host/guest and guest/guest interactions were treated using 12-6 LJ and coulombic potentials as defined above. The coulombic interactions were calculated using the Ewald summation technique with a 10-6 precision and a 12 Å cut-off radius was considered to evaluate the short-range dispersion interactions. For each state point of these simulations, 2x108 Monte Carlo steps following 2x108 equilibration steps have been used. The analysis of the preferential interactions/locations of the guest species was performed through a critical analysis of representative snapshots and from the plots of the radial distribution functions (RDFs) between MOF and guest averaged over all configurations generated by the GCMC simulations. The calculation of the number of hydrogen bonds was conducted using the following geometric criteria: a hydrogen bond is computed when the distance between a donor (D) and acceptor (A) atoms is shorter than 3.5 Å and the angle between the D–H vector and the D–A vector is lower than 37º. These criteria are the same we used previously to describe the Hbond network in other materials53,97–99. Molecular Dynamics simulations The self-diffusivity for water (Ds) in the fully saturated MIL-100(Fe) was explored by means of Molecular Dynamics (MD) simulations performed in the NVT ensemble at 298 K for 10 ns using a time step of 1 fs and the Berendsen thermostat as implemented in the DL_POLY package The GCMC simulated configuration was considered as a starting point while the same force field parameters and charges were the same as those used in the Monte Carlo calculations. The selfdiffusivity was determined using the Einsten relation applied to the Mean-Square displacements averaged over multiple time step origin100.

ACS Paragon Plus Environment

12

Page 13 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

2. Experimental characterization of the water adsorption in MIL-100(Fe) MIL-100(Fe) was prepared from the hydrothermal reaction of trimesic acid, metallic iron, HF, nitric acid and H2O at 423 K for 8 hours, as reported elsewhere.101 The composition of the reaction mixture was 1.0 Fe0:0.67BTC:2.0HF:0.6HNO3:277H2O. The as-synthesized MIL100(Fe) was carefully purified by solvent extraction treatments using hot water and ethanol without chemical treatment using NH4F. The chemical formula of the resulting MIL-100(Fe) is assigned to [Fe3O(H2O)2F0.81(OH)0.19){C6H3(CO2)3}2.nH2O (n≈14.5)], based on chemical analysis. The crystallinity of the sample was confirmed by powder X-ray diffraction analysis (see SI). Thermogravimetric (TG) analysis was carried out in a thermogravimetric analyzer. Analysis was performed in dry nitrogen flow of 100 ml/min. The temperature was increased from 32 to 700oC applying a heating rate of 10oC/min. The particle morphology and crystal size were analyzed using a scanning electron microscope. The evaluation of the BET area was carried out from N2 physisorption isotherms measured at liquid nitrogen temperature (77 K) after dehydration under vacuum at 423 K for 12 h using Micromeritics Tristar 3020 while the pore volume was estimated by a single point method at P/P0 = 0.99. The resulting BET area and pore volume of MIL-100(Fe) were estimated as 2050 m2/g and 1.08 cm3/g, respectively consistent with the previous data reported for this material6,7. The water adsorption isotherms were collected by an intelligent gravimetric analyzer (IGA, Hiden Analytical Ltd.). The IGA was automatically operated to precisely control the water vapor pressure (1-95% RH) and temperature (293-313 K). Prior to adsorption experiments, the sample was dehydrated at 423 K for 6 hours under high vacuum ( 0.2. This observation emphasizes that the choice of the water model is crucial to describe correctly the water adsorption in confined environment. The deviation observed between the two water models mostly comes from the slightly different considered atomic charges and LJ parameters for H2O. Furthermore, a good agreement was also obtained between the GCMC calculated adsorption enthalpy at very low coverage using the TIP4P-Ew model and the experimental heat of adsorption reported by Jeremias et al. 6 (76 kJ.mol-1 and 90 kJ.mol-1 respectively). One can also observe that the experimental heat of adsorption we obtained is well reproduced in the whole range of water loading (Figure 5). This excellent agreement on macroscopic data fully validated the microscopic models considered to describe the H2O/MIL-100(Fe) system with the use of the TIP4P-Ew model to represent H2O, a crucial requirement to further get reliable insight into the microscopic adsorption mechanism in play. Such a good accordance between the experimental and simulated adsorption isotherms is obtained via the consideration of the CUS-water interactions as emphasized by Figure S4 which compares the simulated adsorption isotherms with and without taking into account these specific interactions. Fe(III) CUS are expected to be the preferential adsorption sites for water molecules at the initial step of adsorption, the resulting high affinity leading to the sharp increase in the uptake

ACS Paragon Plus Environment

18

Page 19 of 36

observed at very low P/P0. This is clearly confirmed by the GCMC simulations which evidenced that for a relative pressure of 5x10-4, the amount of water adsorbed is equivalent to the concentration of CUS sites (136 atoms/unit cell or 3.25×10-3 mmol.g-1). In this domain of pressure, the RDF for the Fe(III) CUS-OW (OW corresponding to the oxygen atoms of water) pair (Figure 8a) shows a sharp peak at the equilibrium distance of 2.10 Å. The integration of this peak shows a value of N(r) ≈ 1, indicating that all Fe(III) CUS are coordinated by water.

60

a)

2.0

16

1.5

12

0

b)

P/P = 0.001 0 P/P = 0.01 0 P/P = 0.1 0 P/P = 0.2

4

3

8

2

0.5

4

1

0.0

0 2.0

N (r)

g (r)

1.0

N (r)

40

g (r) 20

0

1

2

3

4

2.5

3.0

r (Å)

3.5

0 4.0

r (Å) 16

0

c)

P/P = 0.001 0 P/P = 0.01 0 P/P = 0.1 0 P/P = 0.2

12

0.5 0.4 0.3

8 0.2 4

0 2.0

N (r)

0

g (r)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0.1

2.5

3.0

3.5

0.0 4.0

r (Å)

Figure 8 – RDFs (solid lines) and their integration (dashed lines) for the pair Fe(III) CUS – OW at P/P0 = 5x10-4 (a) and the pairs OW – OW (b) and OW – O-OH (c) at partial pressures of 0.001 (black lines), 0.01 (red line), 0.1 (blue line) and 0.2 (green line).

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 36

Once the CUS are fully coordinated by water, a second adsorption regime takes place associated with a gradual increase in the uptake up to P/P0 = 0.2. This region is assigned to the interactions between the adsorbed water molecules themselves and with the water molecules coordinated to Fe(III) CUS, which henceforth act as a hydrophilic group, and the terminal -OH groups bounded to Fe (III) CUS. This statement is supported by the RDFs plotted in Figure 8b and 8c, which unambiguously show a higher probability of these interactions between the water molecules and these groups when the partial pressure increases as attested by a corresponding enhancement of the intensity of the corresponding RDF peaks. An in-depth analysis of the snapshots over the whole relative pressure allowed us to identify the sequential water adsorption regimes. The first region corresponds to the water coordination of Fe(III) CUS at very low pressures as discussed above and identified in Figures 9a and 9b. At this stage, the water molecules interact mostly with Fe(III) CUS and only a few molecules interact with the rest of the MOF framework. As the pressure increases (P/P0 = 0.01), the extra H2O molecules form hydrogen bonds with the water coordinated to the Fe(III) CUS labelled as H2OC and the terminal OH groups more effectively in the pentagonal windows but also in the hexagonal ones (Figures 9c and 9d). When the relative pressure increases, the adsorbed water molecules start clustering in the pentagonal windows (P/P0 = 0.1), as seen in Figures 9e and 9f. This clustering continues when the pressure increases up to a point where all the pentagonal windows are occupied by the H2O molecules (P/P0 = 0.2). At this stage, it is remarkable that this clustering also takes place in the hexagonal windows (Figures 9g and 9h). From this point, the water molecules start to populate the cages of the MOF framework following a pore filling mechanism that is associated with the first step observed at P/P0 > 0.2. However, the H2O molecules do not fill both cages of the MOF uniformly. At P/P0 = 0.3, a comparison between

ACS Paragon Plus Environment

20

Page 21 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figures 9i and 9j shows that the smaller cages of the framework are filled first while the larger cages, the only ones containing hexagonal windows, remain practically unoccupied. This observation is consistent with the fact that the most confined environment corresponds to the preferential region for the water adsorption. Such a scenario continues up to P/P0 = 0.5 and besides, the larger cages start to be populated (Figures 9k and 9l) corresponding to the second step observed in the experimental adsorption isotherm at P/P0 = 0.55. Figures S5 and S6 provide a more detailed perspective of the formation of extensive clusters in the windows and cages of the framework respectively. Complementary MD simulations evidenced that the self-diffusivity for water at saturation of 4.8×10-10 m2/s is within the same order of magnitude than the value previously reported for water confined in diverse MOFs85,107–110.

a)

b)

c)

d)

e)

f)

g)

h)

i)

j)

k)

l)

ACS Paragon Plus Environment

21

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 36

Figure 9 – Representative snapshots showing the simulated preferential adsorption sites of water in the pentagonal (a, c, e, g, I and k) and hexagonal (b, d, f, h, j and l) windows of MIL-100(Fe) at P/P0 = 0.001 (a, b), 0.01 (c, d), 0.1 (e, f), 0.2 (g, h), 0.3 (i, j) and 0.5 (k, l). For clarity reasons, the water molecules and the MOF are represented differently. Framework atoms are represented by sticks where C, O, H and Fe atoms are in grey, red, white and blue respectively. Water molecules are represented in balls and sticks with O and H atoms in green and white respectively. Hydrogen bonds are represented by dashed blue lines. Furthermore, the desorption isotherm reported in Figure S7 evidences that the water release occurs in a single step at about P/P0 = 0.3. This relative pressure corresponds to the complete clustering of water in the pentagonal cages of the MOF (Figure 9i). Considering that the hexagonal cages are surrounded by pentagonal cages, we may then assume that the water molecules strongly adsorbed at the pentagonal cages retard the emptying of both cages. As a further step, a careful analysis of the hydrogen bonding of the water molecules was also considered. Figure 10 reports the evolution of the number of hydrogen bonds between each hydrophilic group of the MOF framework (H2Oc and –OH terminal) and the adsorbed water molecules as a function of the water loading. This plot clearly shows a larger fraction of terminal -OH groups involved in the hydrogen bonding vs H2OC in the range P/P0 < 0.2, for which the filling of the smaller pores is achieved. This observation is consistent with the higher intensity of the RDF corresponding to the terminal-OH / H2O pair (Figure 8c) compared to that of the H2OC /H2O pair (Figure 8b). When P/P0 = 0.2, the fraction of both hydrogen bonds towards terminalOH and H2OC becomes similar, each species establishing 1 hydrogen bond in average. Once this stage is achieved, we observe a sudden increase of the fraction of hydrogen bonds which follows the same trend than the adsorption isotherm, with a plateau of 2 hydrogen bonds per group attained at P/P0 > 0.5.

ACS Paragon Plus Environment

22

Page 23 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 10 – Evolution of the number of terminal -OH – H2O hydrogen bonds per -OH group (black squares) and the number of H2OC – H2O hydrogen bonds per H2OC molecule (red circles) as a function of the relative pressure. During the population of the cages by water, which occurs in the range 0.2 < P/P0 < 0.5, it is clearly stated that the number of hydrogen bonds per –OH and H2OC increases significantly up to reach a value of 2.5. In Figures 11a-d, the evolution of the number of hydrogen bonds involving OH groups and H2OC molecules is illustrated by representative snapshots considered at the beginning (P/P0 = 0.24, Figures 11a and 11b) and in the middle of the filling process (P/P0 = 0.44, Figures 11c and 11d). The participation of the -OH and H2OC groups in the filling process is reinforced by their involvement in the cluster formed by H2O within the cages of the MOF (cf. Figure 11c and 11d).

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

a)

Page 24 of 36

b)

0

0

P/P  = 0.24

c)

P/P  = 0.24

d)

0

P/P  = 0.44

0

P/P  = 0.44

Figure 11 – Snapshots showing the hydrogen bonds (blue dashed lines) formed between the H2O molecules and the -OH groups (a, c), H2OC and other adsorbed molecules (b, d) at the partial pressures of 0.24 and 0.44 respectively. At the beginning of the filling process, each hydrophilic group is in average hydrogen bonded to one H2O molecule. With the progress of the pore filling, the number of H2O molecules connected to these sites is doubled. The interaction between H2Oc and the Fe(III) CUS is illustrated by the dashed black lines.

ACS Paragon Plus Environment

24

Page 25 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

CONCLUSIONS Molecular simulations were coupled with gravimetric measurements to explore the water adsorption behaviour in the CUS containing mesoporous MOF MIL-100(Fe). The microscopic models used for both water and MIL-100(Fe) and the water/MIL-100(Fe) force field parameters derived by preliminary DFT calculations were validated by a good agreement between the experimental and simulated water adsorption isotherm and heat of adsorption. In this study, the TIP4P-Ew model was demonstrated to be the most adequate to represent the water molecule. The strong uncoordinated metal sites - water molecules interactions revealed at the initial stage of adsorption were correlated to the high heat of adsorption at zero coverage observed experimentally. Water molecules coordinated to these sites were observed to act as hydrophilic centres anchoring new water clusters via hydrogen bonds, favouring the hydrophilicity of the material. Moreover, the multi-step profile of the adsorption isotherm was explained by a sequential filling of the two distinct cages present in MIL-100(Fe). Although a direct design of water MOF adsorbents from the rationalization of the water adsorption performances of existing MOFs is still far to be achieved, a systematic exploration of the adsorption mechanism at the microscopic scale is expected to guide the development of new generation MOFs for water adsorption-based heat reallocation applications.

ASSOCIATED CONTENT

ACS Paragon Plus Environment

25

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 36

Supporting Information. The following file is available free of charge. Sample characterization data (PDF)

AUTHOR INFORMATION Corresponding Authors * E-mail: [email protected]. [email protected] Author Contributions K. H. C., J. P. and J.-S. C. synthesized, characterized and performed the adsorption measurements on the MOF sample; P.G.M.M., S.D.-V., and G.M. performed the computational simulations and the related analysis and they wrote the manuscript. Notes The authors declare no competing financial interest. ACKNOWLEDGMENTS P.G.M.M. thanks the National Counsel of Technological and Scientific Development (CNPq) for his PhD scholarship. The Korean authors are grateful to the Global Frontier Center for Hybrid Interface Materials (GFHIM, Grant No. NRF-2013M3A6B1078879) and the Center for Convergent Chemical Process (CCP, Grant No. SKC1810-4) for financial support. J. W. Yoon, Y. K. Hwang and U. -H Lee (KRICT) are also acknowledged for their helpful assistance and discussions of the water sorption experiments as well as the synthesis of the MIL-100(Fe).

ACS Paragon Plus Environment

26

Page 27 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

REFERENCES (1)

Demir, H.; Mobedi, M.; Ülkü, S. A Review on Adsorption Heat Pump: Problems and Solutions. Renew. Sustain. Energy Rev. 2008, 12, 2381–2403.

(2)

Hastürk, E.; Ernst, S. J.; Janiak, C. Recent Advances in Adsorption Heat Transformation Focusing on the Development of Adsorbent Materials. Curr. Opin. Chem. Eng. 2019, 24, 26–36.

(3)

Aristov, Y. I. Novel Materials for Adsorptive Heat Pumping and Storage: Screening and Nanotailoring of Sorption Properties. J. Chem. Eng. Japan 2007, 40, 1242–1251.

(4)

Freni, A.; Bonaccorsi, L.; Calabrese, L.; Caprì, A.; Frazzica, A.; Sapienza, A. SAPO-34 Coated Adsorbent Heat Exchanger for Adsorption Chillers. Appl. Therm. Eng. 2015, 82, 1–7.

(5)

Henninger, S. K.; Habib, H. A.; Janiak, C. MOFs as Adsorbents for Low Temperature Heating and Cooling Applications. J. Am. Chem. Soc. 2009, 131, 2776–2777.

(6)

Jeremias, F.; Khutia, A.; Henninger, S. K.; Janiak, C. MIL-100(Al, Fe) as Water Adsorbents for Heat Transformation Purposes - a Promising Application. J. Mater. Chem. 2012, 22, 10148–10151.

(7)

Akiyama, G.; Matsuda, R.; Kitagawa, S. Highly Porous and Stable Coordination Polymers as Water Sorption Materials. Chem. Lett. 2010, 39, 360–361.

(8)

Horcajada, P.; Surblé, S.; Serre, C.; Hong, D.-Y.; Seo, Y.-K.; Chang, J.-S.; Grenèche, J.M.; Margiolaki, I.; Férey, G. Synthesis and Catalytic Properties of MIL-100(Fe), an Iron(III) Carboxylate with Large Pores. Chem. Commun. 2007, 100, 2820–2822.

(9)

Ehrenmann, J.; Henninger, S. K.; Janiak, C. Water Adsorption Characteristics of MIL-101 for Heat-Transformation Applications of MOFs. Eur. J. Inorg. Chem. 2011, 4, 471–474.

(10)

Férey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surblé, S.; Margiolaki, I. A Chromium Terephthalate-Based Solid with Unusually Large Pore Volumes and Surface Area. Science 2005, 309, 2040–2042.

(11)

Henninger, S. K.; Ernst, S. J.; Gordeeva, L.; Bendix, P.; Fröhlich, D.; Grekova, A. D.; Bonaccorsi, L.; Aristov, Y.; Jaenchen, J. New Materials for Adsorption Heat Transformation and Storage. Renew. Energy 2017, 110, 59–68.

ACS Paragon Plus Environment

27

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(12)

Page 28 of 36

Henninger, S. K.; Schmidt, F. P.; Henning, H.-M. Water Adsorption Characteristics of Novel Materials for Heat Transformation Applications. Appl. Therm. Eng. 2010, 30, 1692–1702.

(13) De Lange, M. F.; Zeng, T.; Vlugt, T. J. H.; Gascon, J.; Kapteijn, F. Manufacture of Dense CAU-10-H Coatings for Application in Adsorption Driven Heat Pumps: Optimization and Characterization. Cryst. Eng. Comm. 2015, 17, 5911–5920. (14)

Lenzen, D.; Bendix, P.; Reinsch, H.; Fröhlich, D.; Kummer, H.; Möllers, M.; Hügenell, P. P. C.; Gläser, R.; Henninger, S.; Stock, N. Scalable Green Synthesis and Full-Scale Test of the Metal–Organic Framework CAU-10-H for Use in Adsorption-Driven Chillers. Adv. Mater. 2018, 30, 1–6.

(15)

Cadiau, A.; Lee, J. S.; Damasceno Borges, D.; Fabry, P.; Devic, T.; Wharmby, M. T.; Martineau, C.; Foucher, D.; Taulelle, F.; Jun, C.-H.; et al. Design of Hydrophilic Metal Organic Framework Water Adsorbents for Heat Reallocation. Adv. Mater. 2015, 27, 4775–4780.

(16) Permyakova, A.; Skrylnyk, O.; Courbon, E.; Affram, M.; Wang, S.; Lee, U. H.; Valekar, A. H.; Nouar, F.; Mouchaham, G.; Devic, T.; et al. Synthesis Optimization, Shaping, and Heat Reallocation Evaluation of the Hydrophilic Metal–Organic Framework MIL160(Al). Chem. Sus. Chem. 2017, 10, 1419–1426. (17)

Gordeeva, L. G.; Solovyeva, M. V.; Aristov, Y. I. NH2 -MIL-125 as a Promising Material for Adsorptive Heat Transformation and Storage. Energy 2016, 100, 18–24.

(18)

Jeremias, F.; Lozan, V.; Henninger, S. K.; Janiak, C. Programming MOFs for Water Sorption: Amino-Functionalized MIL-125 and UiO-66 for Heat Transformation and Heat Storage Applications. Dalt. Trans. 2013, 42, 15967.

(19)

Jeremias, F.; Fröhlich, D.; Janiak, C.; Henninger, S. K. Advancement of Sorption-Based Heat Transformation by a Metal Coating of Highly-Stable, Hydrophilic Aluminium Fumarate MOF. RSC Adv. 2014, 4, 24073–24082.

(20) Alvarez, E.; Guillou, N.; Martineau, C.; Bueken, B.; Van de Voorde, B.; Le Guillouzer, C.; Fabry, P.; Nouar, F.; Taulelle, F.; de Vos, D.; et al. The Structure of the Aluminum Fumarate Metal-Organic Framework A520. Angew. Chemie Int. Ed. 2015, 54, 3664–3668. (21)

Henninger, S. K.; Jeremias, F.; Kummer, H.; Janiak, C. MOFs for Use in Adsorption Heat Pump Processes. Eur. J. Inorg. Chem. 2012, 2012, 2625–2634.

(22) Elsayed, E.; AL-Dadah, R.; Mahmoud, S.; Elsayed, A.; Anderson, P. A. Aluminium Fumarate and CPO-27(Ni) MOFs: Characterization and Thermodynamic Analysis for Adsorption Heat Pump Applications. Appl. Therm. Eng. 2016, 99, 802–812. (23) Lillerud, K. P.; Cavka, J. H.; Lamberti, C.; Guillou, N.; Bordiga, S.; Jakobsen, S.; Olsbye, U. A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks with Exceptional Stability. J. Am. Chem. Soc. 2008, 130, 13850–13851. (24)

Furukawa, H.; Gándara, F.; Zhang, Y.-B.; Jiang, J.; Queen, W. L.; Hudson, M. R.; Yaghi,

ACS Paragon Plus Environment

28

Page 29 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

O. M. Water Adsorption in Porous Metal–Organic Frameworks and Related Materials. J. Am. Chem. Soc. 2014, 136, 4369–4381. (25)

Cui, S.; Qin, M.; Marandi, A.; Steggles, V.; Wang, S.; Feng, X.; Nouar, F.; Serre, C. Metal-Organic Frameworks as Advanced Moisture Sorbents for Energy-Efficient High Temperature Cooling. Sci. Rep. 2018, 8, 2–10.

(26)

Rieth, A. J.; Yang, S.; Wang, E. N.; Dincǎ, M. Record Atmospheric Fresh Water Capture and Heat Transfer with a Material Operating at the Water Uptake Reversibility Limit. ACS Cent. Sci. 2017, 3, 668–672.

(27) Wade, C. R.; Corrales-Sanchez, T.; Narayan, T. C.; Dincă, M. Postsynthetic Tuning of Hydrophilicity in Pyrazolate MOFs to Modulate Water Adsorption Properties. Energy Environ. Sci. 2013, 6, 2172. (28)

Abdulhalim, R. G.; Bhatt, P. M.; Belmabkhout, Y.; Shkurenko, A.; Adil, K.; Barbour, L. J.; Eddaoudi, M. A Fine-Tuned Metal-Organic Framework for Autonomous Indoor Moisture Control. J. Am. Chem. Soc. 2017, 139, 10715–10722.

(29)

Wright, A. M.; Rieth, A. J.; Yang, S.; Wang, E. N.; Dincă, M. Precise Control of Pore Hydrophilicity Enabled by Post-Synthetic Cation Exchange in Metal–Organic Frameworks. Chem. Sci. 2018, 9, 3856–3859.

(30) Tannert, N.; Ernst, S. J.; Jansen, C.; Bart, H. J.; Henninger, S. K.; Janiak, C. Evaluation of the Highly Stable Metal-Organic Framework MIL-53(Al)-TDC (TDC = 2,5Thiophenedicarboxylate) as a New and Promising Adsorbent for Heat Transformation Applications. J. Mater. Chem. A 2018, 6, 17706–17712. (31) Towsif Abtab, S. M.; Alezi, D.; Bhatt, P. M.; Shkurenko, A.; Belmabkhout, Y.; Aggarwal, H.; Weseliński, Ł. J.; Alsadun, N.; Samin, U.; Hedhili, M. N.; et al. Reticular Chemistry in Action: A Hydrolytically Stable MOF Capturing Twice Its Weight in Adsorbed Water. Chem. 2018, 4, 94–105. (32) Rieth, A. J.; Wright, A. M.; Rao, S.; Kim, H.; Lapotin, A. D.; Wang, E. N.; Dinca, M. Tunable Metal-Organic Frameworks Enable High-Efficiency Cascaded Adsorption Heat Pumps. J. Am. Chem. Soc. 2018, 140, 17591–17596. (33)

An, H. J.; Sarker, M.; Yoo, D. K.; Jhung, S. H. Water Adsorption/Desorption over MetalOrganic Frameworks with Ammonium Group for Possible Application in Adsorption Heat Transformation. Chem. Eng. J. 2019, 373, 1064–1071.

(34)

Wang, S.; Lee, J. S.; Wahiduzzaman, M.; Park, J.; Muschi, M.; Martineau-Corcos, C.; Tissot, A.; Cho, K. H.; Marrot, J.; Shepard, W.; et al. A Robust Large-Pore Zirconium Carboxylate Metal–Organic Framework for Energy-Efficient Water-Sorption-Driven Refrigeration. Nat. Energy 2018, 3, 985–993.

(35)

Bon, V.; Senkovska, I.; Evans, J. D.; Wöllner, M.; Hölzel, M.; Kaskel, S. Insights into the Water Adsorption Mechanism in the Chemically Stable Zirconium-Based MOF DUT-67-a Prospective Material for Adsorption-Driven Heat Transformations. J. Mater. Chem. A

ACS Paragon Plus Environment

29

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 36

2019, 7, 12681–12690. (36)

Lee, J. S.; Yoon, J. W.; Mileo, P. G. M.; Cho, K. H.; Park, J.; Kim, K.; Kim, H.; de Lange, M. F.; Kapteijn, F.; Mqurin, G. Humphrey, S. M.; Chang, J.-S. The Porous Metal-Organic Framework CUK-1 for Heat Allocation toward Green Applications of Natural Refrigerant Water. ACS Appl. Mater. Interfaces 2019, 11, 25778-25789.

(37) Rieth, A. J.; Wright, A. M.; Skorupskii, G.; Mancuso, J. L.; Hendon, C. H.; Dinca, M. Record-Setting Sorbents for Reversible Water Uptake by Systematic Anion-Exchanges in Metal-Organic Frameworks. J. Am. Chem. Soc. 2019, doi.org/10.1021/jacs.9b06246.. (38)

Lenzen, D.; Zhao, J.; Ernst, S.-J.; Wahiduzzaman, M.; Ken Inge, A.; Fröhlich, D.; Xu, H.; Bart, H.-J.; Janiak, C.; Henninger, S.; et al. A Metal–Organic Framework for Efficient Water-Based Ultra-Low-Temperature-Driven Cooling. Nat. Commun. 2019, 10, 30253034.

(39)

Fei, H.; Pullen, S.; Wagner, A.; Ott, S.; Cohen, S. M. Functionalization of Robust Zr(IV)Based Metal–Organic Framework Films via a Postsynthetic Ligand Exchange. Chem. Commun. 2015, 51, 66–69.

(40) Cohen, S. M. Postsynthetic Methods for the Functionalization of Metal–Organic Frameworks. Chem. Rev. 2012, 112, 970–1000. (41) Cohen, S. M. The Postsynthetic Renaissance in Porous Solids. J. Am. Chem. Soc. 2017, 139, 2855–2863. (42)

Yuan, L.; Tian, M.; Lan, J.; Cao, X.; Wang, X.; Chai, Z.; Gibson, J. K.; Shi, W. Defect Engineering in Metal–Organic Frameworks: A New Strategy to Develop Applicable Actinide Sorbents. Chem. Commun. 2018, 54, 370–373.

(43) Cheetham, A. K.; Bennett, T. D.; Coudert, F.-X.; Goodwin, A. L. Defects and Disorder in Metal Organic Frameworks. Dalt. Trans. 2016, 45, 4113–4126. (44)

Ren, J.; Ledwaba, M.; Musyoka, N. M.; Langmi, H. W.; Mathe, M.; Liao, S.; Pang, W. Structural Defects in Metal–Organic Frameworks (MOFs): Formation, Detection and Control towards Practices of Interests. Coord. Chem. Rev. 2017, 349, 169–197.

(45)

Li, W.; Xia, X.; Cao, M.; Li, S. Structure–Property Relationship of Metal–Organic Frameworks for Alcohol-Based Adsorption-Driven Heat Pumps via High-Throughput Computational Screening. J. Mater. Chem. A 2019, 7, 7470–7479.

(46)

Berendsen, H. J. C.; Grigera, J. R.; Straatsma, T. P. The Missing Term in Effective Pair Potentials. J. Phys. Chem. 1987, 91, 6269–6271.

(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)

Horn, H. W.; Swope, W. C.; Pitera, J. W.; Madura, J. D.; Dick, T. J.; Hura, G. L.; HeadGordon, T. Development of an Improved Four-Site Water Model for Biomolecular

ACS Paragon Plus Environment

30

Page 31 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Simulations: TIP4P-Ew. J. Chem. Phys. 2004, 120, 9665–9678. (49)

Abascal, J. L.; Vega, C. A General Purpose Model for the Condensed Phases of Water: TIP4P/2005. J. Chem. Phys. 2005, 123, 234505.

(50)

Wang, L.-P.; Head-Gordon, T.; Ponder, J. W.; Ren, P.; Chodera, J. D.; Eastman, P. K.; Martinez, T. J.; Pande, V. S. Systematic Improvement of a Classical Molecular Model of Water. J. Phys. Chem. B 2013, 117, 9956–9972.

(51)

Khalak, Y.; Baumeier, B.; Karttunen, M. Improved General-Purpose Five-Point Model for Water: TIP5P/2018. J. Chem. Phys. 2018, 149, 224507.

(52)

Bourrelly, S.; Moulin, B.; Rivera, A.; Maurin, G.; Devautour-Vinot, S.; Serre, C.; Devic, T.; Horcajada, P.; Vimont, A.; Clet, G.; et al. Explanation of the Adsorption of Polar Vapors in the Highly Flexible Metal Organic Framework MIL-53(Cr). J. Am. Chem. Soc. 2010, 132, 9488–9498.

(53) Borges, D. D.; Devautour-Vinot, S.; Jobic, H.; Ollivier, J.; Nouar, F.; Semino, R.; Devic, T.; Serre, C.; Paesani, F.; Maurin, G. Proton Transport in a Highly Conductive Porous Zirconium-Based Metal-Organic Framework: Molecular Insight. Angew. Chemie - Int. Ed. 2016, 55, 3919–3924. (54)

Borges, D. D.; Semino, R.; Devautour-Vinot, S.; Jobic, H.; Paesani, F.; Maurin, G. Computational Exploration of the Water Concentration Dependence of the Proton Transport in the Porous UiO–66(Zr)–(CO 2 H) 2 Metal–Organic Framework. Chem. Mater. 2017, 29, 1569–1576.

(55) Damasceno Borges, D.; Maurin, G.; Galvão, D. S. Design of Porous Metal-Organic Frameworks for Adsorption Driven Thermal Batteries. MRS Adv. 2017, 2, 519–524. (56) Mileo, P. G. M.; Adil, K.; Davis, L.; Cadiau, A.; Belmabkhout, Y.; Aggarwal, H.; Maurin, G.; Eddaoudi, M.; Devautour-Vinot, S. Achieving Superprotonic Conduction with a 2D Fluorinated Metal-Organic Framework. J. Am. Chem. Soc. 2018, 140, 13156-13160. (57)

Küsgens, P.; Rose, M.; Senkovska, I.; Fröde, H.; Henschel, A.; Siegle, S.; Kaskel, S. Characterization of Metal-Organic Frameworks by Water Adsorption. Microporous Mesoporous Mater. 2009, 120, 325–330.

(58)

Vimont, A.; Goupil, J.-M.; Lavalley, J.-C.; Daturi, M.; Surblé, S.; Serre, C.; Millange, F.; Férey, G.; Audebrand, N. Investigation of Acid Sites in a Zeotypic Giant Pores Chromium(III) Carboxylate. J. Am. Chem. Soc. 2006, 128, 3218–3227.

(59) Fang, H.; Demir, H.; Kamakoti, P.; Sholl, D. S. Recent Developments in First-Principles Force Fields for Molecules in Nanoporous Materials. J. Mater. Chem. A 2014, 2, 274–291. (60) Fischer, M.; Gomes, J. R. B.; Jorge, M. Computational Approaches to Study Adsorption in MOFs with Unsaturated Metal Sites. Mol. Simul. 2014, 40, 537–556. (61)

Rudenko, A. N.; Bendt, S.; Keil, F. J. Multiscale Modeling of Water in Mg-MOF-74: From Electronic Structure Calculations to Adsorption Isotherms. J. Phys. Chem. C 2014,

ACS Paragon Plus Environment

31

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 36

118, 16218–16227. (62)

Chen, L.; Morrison, C. A.; Düren, T. Improving Predictions of Gas Adsorption in Metal– Organic Frameworks with Coordinatively Unsaturated Metal Sites: Model Potentials, Ab Initio Parameterization, and GCMC Simulations. J. Phys. Chem. C 2012, 116, 18899– 18909.

(63)

Fischer, M.; Gomes, J. R. B.; Fröba, M.; Jorge, M. Modeling Adsorption in Metal-Organic Frameworks with Open Metal Sites: Propane/Propylene Separations. Langmuir 2012, 28, 8537–8549.

(64) 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. (65) Forrest, K. A.; Pham, T.; McLaughlin, K.; Belof, J. L.; Stern, A. C.; Zaworotko, M. J.; Space, B. Simulation of the Mechanism of Gas Sorption in a Metal-Organic Framework with Open Metal Sites: Molecular Hydrogen in PCN-61. J. Phys. Chem. C 2012, 116, 15538–15549. (66)

Becker, T. M.; Lin, L. C.; Dubbeldam, D.; Vlugt, T. J. H. Polarizable Force Field for CO2 in M-MOF-74 Derived from Quantum Mechanics. J. Phys. Chem. C 2018, 122, 24488– 24498.

(67)

Chen, L.; Grajciar, L.; Nachtigall, P.; Düren, T. Accurate Prediction of Methane Adsorption in a Metal-Organic Framework with Unsaturated Metal Sites by Direct Implementation of an Ab Initio Derived Potential Energy Surface in GCMC Simulation. J. Phys. Chem. C 2011, 115, 23074–23080.

(68) Lee, K.; Howe, J. D.; Lin, L. C.; Smit, B.; Neaton, J. B. Small-Molecule Adsorption in Open-Site Metal-Organic Frameworks: A Systematic Density Functional Theory Study for Rational Design. Chem. Mater. 2015, 27, 668–678. (69)

Bak, J. H.; Le, V. D.; Kang, J.; Wei, S. H.; Kim, Y. H. First-Principles Study of Electronic Structure and Hydrogen Adsorption of 3d Transition Metal Exposed Paddle Wheel Frameworks. J. Phys. Chem. C 2012, 116, 7386–7392.

(70)

Sun, Y. Y.; Kim, Y. H.; Zhang, S. B. Effect of Spin State on the Dihydrogen Binding Strength to Transition Metal Centers in Metal-Organic Frameworks. J. Am. Chem. Soc. 2007, 129, 12606–12607.

(71)

Haldoupis, E.; Borycz, J.; Shi, H.; Vogiatzis, K. D.; Bai, P.; Queen, W. L.; Gagliardi, L.; Siepmann, J. I. Ab Initio Derived Force Fields for Predicting CO 2 Adsorption and Accessibility of Metal Sites in the Metal–Organic Frameworks M-MOF-74 (M = Mn, Co, Ni, Cu). J. Phys. Chem. C 2015, 119, 16058–16071.

(72)

Kim, Y. H.; Kang, J.; Wei, S. H. Origin of Enhanced Dihydrogen-Metal Interaction in Carboxylate Bridged Cu2-Paddle-Wheel Frameworks. Phys. Rev. Lett. 2010, 105, 1–4.

(73) Grajciar, L.; Bludský, O.; Nachtigall, P. Water Adsorption on Coordinatively Unsaturated

ACS Paragon Plus Environment

32

Page 33 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Sites in CuBTC MOF. J. Phys. Chem. Lett. 2010, 1, 3354–3359. (74)

Fischer, M.; Hoffmann, F.; Fröba, M. Molecular Simulation of Hydrogen Adsorption in Metal-Organic Frameworks. Colloids Surfaces A Physicochem. Eng. Asp. 2010, 357, 35– 42.

(75)

Verma, P.; Xu, X.; Truhlar, D. G. Adsorption on Fe-MOF-74 for C1-C3 Hydrocarbon Separation. J. Phys. Chem. C 2013, 117, 12648–12660.

(76)

Kulkarni, A. R.; Sholl, D. S. DFT-Derived Force Fields for Modeling Hydrocarbon Adsorption in MIL-47(V). Langmuir 2015, 31, 8453–8468.

(77)

Kulkarni, A. R.; Sholl, D. S. Screening of Copper Open Metal Site MOFs for Olefin/Paraffin Separations Using DFT-Derived Force Fields. J. Phys. Chem. C 2016, 120, 23044–23054.

(78)

Campbell, C.; Ferreiro-Rangel, C. A.; Fischer, M.; Gomes, J. R. B.; Jorge, M. A Transferable Model for Adsorption in Mofs with Unsaturated Metal Sites. J. Phys. Chem. C 2017, 121, 441–458.

(79)

Campbell, C.; Gomes, J. R. B.; Fischer, M.; Jorge, M. New Model for Predicting Adsorption of Polar Molecules in Metal-Organic Frameworks with Unsaturated Metal Sites. J. Phys. Chem. Lett. 2018, 9, 3544–3553.

(80) Fetisov, E. O.; Shah, M. S.; Long, J. R.; Tsapatsis, M.; Ilja Siepmann, J. First Principles Monte Carlo Simulations of Unary and Binary Adsorption: CO 2 , N 2 , and H 2 O in MgMOF-74. Chem. Commun. 2018, 54, 10816–10819. (81)

Lin, L. C.; Lee, K.; Gagliardi, L.; Neaton, J. B.; Smit, B. Force-Field 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.

(82)

Zang, J.; Nair, S.; Sholl, D. S. Prediction of Water Adsorption in Copper-Based MetalOrganic Frameworks Using Force Fields Derived from Dispersion-Corrected DFT Calculations. J. Phys. Chem. C 2013, 117, 7519–7525.

(83) De Lange, M. F.; Gutiérrez-Sevillano, J. J.; Hamad, S.; Vlugt, T. J. H.; Calero, S.; Gascon, J.; Kapteijn, F. Understanding Adsorption of Highly Polar Vapors on Mesoporous MIL100 (Cr) and MIL-101 (Cr): Experiments and Molecular Simulations. J. Phys. Chem. C 2013, 117, 7613–7622. (84)

Splith, T.; Pantatosaki, E.; Kolokathis, P. D.; Fröhlich, D.; Zhang, K.; Füldner, G.; Chmelik, C.; Jiang, J.; Henninger, S. K.; Stallmach, F.; et al. Molecular Dynamics Phenomena of Water in the Metalorganic Framework MIL-100(Al), as Revealed by Pulsed Field Gradient NMR and Atomistic Simulation. J. Phys. Chem. C 2017, 121, 18065–18074.

(85)

Kolokathis, P. D.; Pantatosaki, E.; Papadopoulos, G. K. Atomistic Modeling of Water Thermodynamics and Kinetics within MIL-100(Fe). J. Phys. Chem. C 2015, 119, 20074–

ACS Paragon Plus Environment

33

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 36

20084. (86) Chen, Y.-R.; Liou, K.-H.; Kang, D.-Y.; Chen, J.-J.; Lin, L.-C. Investigation of the Water Adsorption Properties and Structural Stability of MIL-100(Fe) with Different Anions. Langmuir 2018, 34, 4180–4187. (87)

Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G. L.; Cococcioni, M.; Dabo, I.; et al. QUANTUM ESPRESSO: A Modular and Open-Source Software Project for Quantum Simulations of Materials. J. Phys. Condens. Matter 2009, 21, 395502.

(88) Giannozzi, P.; Andreussi, O.; Brumme, T.; Bunau, O.; Buongiorno Nardelli, M.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Cococcioni, M.; et al. Advanced Capabilities for Materials Modelling with Quantum ESPRESSO. J. Phys. Condens. Matter 2017, 29, 465901-465931. (89)

Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868.

(90) Grimme, S. Semiempirical GGA-Type Density Functional Constructed with a LongRange Dispersion Correction. J. Comput. Chem. 2006, 27, 1787–1799. (91)

Vanderbilt, D. Soft Self-Consistent Pseudopotentials in a Generalized Eigenvalue Formalism. Phys. Rev. B 1990, 41, 7892–7895.

(92)

Mayo, S. L.; Olafson, B. D.; Iii, W. a G.; Eb, E.; El, E. a E. T. DREIDING: A Generic Force Field for Molecular Simulations. J. Phys. Chem. 1990, 101, 8897–8909.

(93)

Rappé, A. K. K.; Casewit, C. J. J.; Colwell, K. S. S.; Goddard III, 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.

(94)

Heinz, H.; Suter, U. W. Atomic Charges for Classical Simulations of Polar Systems. J. Phys. Chem. B 2004, 108, 18341–18352.

(95) Delley, B. From Molecules to Solids with the DMol3 Approach. J. Chem. Phys. 2000, 113, 7756–7764. (96)

Vlugt, T. J. H.; García-Pérez, E.; Dubbeldam, D.; Ban, S.; Calero, S. Computing the Heat of Adsorption Using Molecular Simulations: The Effect of Strong Coulombic Interactions. J. Chem. Theory Comput. 2008, 4, 1107–1118.

(97)

Mileo, P. G. M.; Devautour-Vinot, S.; Mouchaham, G.; Faucher, F.; Guillou, N.; Vimont, A.; Serre, C.; Maurin, G. Proton-Conducting Phenolate-Based Zr Metal–Organic Framework: A Joint Experimental–Modeling Investigation. J. Phys. Chem. C 2016, 120, 24503–24510.

(98)

Mileo, P. G. M.; Kundu, T.; Semino, R.; Benoit, V.; Steunou, N.; Llewellyn, P. L.; Serre, C.; Maurin, G.; Devautour-Vinot, S. Highly Efficient Proton Conduction in a ThreeDimensional Titanium Hydrogen Phosphate. Chem. Mater. 2017, 29.

ACS Paragon Plus Environment

34

Page 35 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(99) Mileo, P. G. M.; Adil, K.; Davis, L.; Cadiau, A.; Belmabkhout, Y.; Aggarwal, H.; Maurin, G.; Eddaoudi, M.; Devautour-Vinot, S. Achieving Superprotonic Conduction with a 2D Fluorinated Metal–Organic Framework. J. Am. Chem. Soc. 2018, 140, 13156–13160. (100) Frenkel, D.; Smit, B.; Ratner, M. A. Understanding Molecular Simulation: From Algorithms to Applications. Phys. Today 1997, 50, 66–66. (101) Seo, Y. K.; Yoon, J. W.; Lee, J. S.; Lee, U. H.; Hwang, Y. K.; Jun, C. H.; Horcajada, P.; Serre, C.; Chang, J. S. Large Scale Fluorine-Free Synthesis of Hierarchically Porous Iron(III) Trimesate MIL-100(Fe) with a Zeolite MTN Topology. Microporous Mesoporous Mater. 2012, 157, 137–145. (102) Pan, H.; Ritter, J. A.; Balbuena, P. B. Examination of the Approximations Used in Determining the Isosteric Heat of Adsorption from the Clausius-Clapeyron Equation. Langmuir 1998, 14, 6323–6327. (103) de Lange, M. F.; Verouden, K. J. F. M.; Vlugt, T. J. H.; Gascon, J.; Kapteijn, F. Adsorption-Driven Heat Pumps: The Potential of Metal–Organic Frameworks. Chem. Rev. 2015, 115, 12205–12250. (104) Jarzcki, A. A.; Anbar, A. D.; Spiro, T. G.; Jarzecki, A.; Anbar, A. D.; Spiro, T. G. DFT Analysis of Fe(H 2 O) 6 3+ and Fe(H 2 O) 6 2+ Structure and Vibrations; Implications for Isotope Fractionation. J. Phys. Chem. A 2004, 108, 2726–2732. (105) Schultz, R. H.; Armentrout, P. B. Gas-Phase Metal Ion Ligation: Collision-Induced Dissociation of Aquairon (Fe(H2O)X+) and Iron-Methane (Fe(CH4)X+) (X=1-4). J. Phys. Chem. 1993, 97, 596–603. (106) Ricca, A.; Bauschlicher, C. W. Successive H2O Binding Energies for Fe(H2O)N+. J. Phys. Chem. 1995, 99, 9003–9007. (107) Splith, T.; Fröhlich, D.; Henninger, S. K.; Stallmach, F. Development and Application of an Exchange Model for Anisotropic Water Diffusion in the Microporous MOF Aluminum Fumarate. J. Magn. Reson. 2018, 291, 40–46. (108) Begum, S.; Wang, Z.; Donnadio, A.; Costantino, F.; Casciola, M.; Valiullin, R.; Chmelik, C.; Bertmer, M.; Kärger, J.; Haase, J.; et al. Water-Mediated Proton Conduction in a Robust Triazolyl Phosphonate Metal-Organic Framework with Hydrophilic Nanochannels. Chem. - A Eur. J. 2014, 20, 8862–8866. (109) Salles, F.; Bourrelly, S.; Jobic, H.; Devic, T.; Guillerm, V.; Llewellyn, P.; Serre, C.; Ferey, G.; Maurin, G. Molecular Insight into the Adsorption and Diffusion of Water in the Versatile Hydrophilic/Hydrophobic Flexible MIL-53(Cr) MOF. J. Phys. Chem. C 2011, 115, 10764–10776. (110) Zhang, K.; Lively, R. P.; Zhang, C.; Koros, W. J.; Chance, R. R. Investigating the Intrinsic Ethanol/Water Separation Capability of ZIF-8: An Adsorption and Diffusion Study. J. Phys. Chem. C 2013, 117, 7214–7225.

ACS Paragon Plus Environment

35

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 36

TOC Graphic

ACS Paragon Plus Environment

36