Metal–Organic Frameworks - ACS Publications - American Chemical

Jul 23, 2015 - Zachary L. Terranova,. ‡. Matthew M. Agee,. ‡ and Francesco Paesani*. Department of Chemistry and Biochemistry, University of Calif...
0 downloads 0 Views 1MB Size
Page 1 of 28

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 Structure and Dynamics in Homochiral [Zn(l-L)(X)] Metal-Organic Frameworks Zachary L. Terranova,‡ Matthew M. Agee,‡ Francesco Paesani* Department of Chemistry and Biochemistry, University of California, San Diego 9500 Gilman Drive, La Jolla, CA 92093 KEYWORDS: Metal-organic frameworks, confined water, proton conduction, interfaces

The structural, thermodynamic, and dynamical properties of water adsorbed in two homochiral metal-organic frameworks (MOFs) with general formula [Zn(l-L)(X)], X = Cl and Br, and L = 3methyl-2-(pyridin-4-ylmethylamino)-butanoic acid, are investigated through molecular dynamics simulations. Water molecules establish distinct hydrogen-bonding patterns within the pores of the two MOFs, which directly correlate with the strength of the underlying framework-water interactions. In particular, at low loading, the Zn-Cl groups of [Zn(l-L)(Cl)] effectively provide a templating scaffold for the formation of one-dimensional hydrogen-bonded water chains that propagate along the MOF channels following the helicity of the framework. In contrast, the relatively weaker framework-water interactions in [Zn(l-L)(Br)] lead to less ordered water distributions inside the pores. The simulation results are in agreement with the available experimental data and provide molecular-level insights into specific hydrogen-bonding motifs and spatial arrangements of the water molecules inside the pores, which can be related to the different proton conductivities measured for the two MOFs.

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 28

1. INTRODUCTION The intrinsic chemical diversity and porosity make metal-organic frameworks (MOFs) promising materials for several technological applications, including gas storage (e.g., methane and hydrogen storage),1-11 carbon capture,12-17 hydrocarbon separation,18-20 catalysis,21-27 electrical28-29 and proton30-42 conductivity, magnetism43-48 and luminescence.49 However, despite their versatility, several frameworks display low stability when exposed to moisture, which has limited the use of MOFs in industrial applications thus far.50-51 On the other hand, MOF solubility has been successfully exploited for in vivo medical applications and water-unstable frameworks have been specifically synthesized for drug delivery and imaging applications under physiological conditions.52-55 A detailed understanding of the behavior of water within the MOF pores, including the identification of fundamental water-framework interactions and the characterization of possible degradation mechanisms is crucial to the rational design of new stable structures tailored for large-scale applications. For example, it is known that small amounts of water in the pores can enhance the CO2 adsorption capacity of certain MOFs, including MOF-100,56 HKUST-1,57 and MIL-10158 while, in other cases, water was found to be detrimental to carbon capture in more hydrophilic MOFs, since both H2O and CO2 molecules compete for the same binding sites within the pores.51 The structural properties of MOFs can also be affected by the presence of water. As mentioned above, some frameworks degrade irreversibly under mild conditions while other structures remain highly stable even when completely immersed in boiling water.51 In general, the stability of MOFs exposed to steam with different levels of saturation can be correlated with the estimated dissociation energy of the metal-ligand bonds. It has been shown that some MOFs,

ACS Paragon Plus Environment

2

Page 3 of 28

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

such as those belonging to the MIL-53 family, can undergo reversible transitions from large- to narrow-pore structures as a function of the amount of water adsorbed in the pores.59-62 More recently it has been demonstrated that water molecules can mediate proton conduction through the MOF pores.37 For example, a significant variation in proton conduction was observed upon water adsorption in a series of isostructural frameworks belonging to the MIL-53 family, with the differences in the measured conductivities being correlated with the acidity of the functional groups of the framework.63 A family of four homochiral MOFs with general formulas [Zn(γ-L)(X)], with γ = l or d, X = Cl and Br, and L = 3-methyl-2-(pyridin-4ylmethylamino)-butanoic acid, was also investigated for proton conduction.64 Although all four MOFs were found to adsorb water, proton conductivity (4.4 x 10-5 S cm-1) was only observed in the two [Zn(γ-L)(Cl)] enantiomeric structures. Proton conduction in MOFs with general formula (NH4)2(adp)[Zn2(ox)3] was also shown to be remarkably dependent on the amount of water adsorbed in the pores.33 A molecular-level understanding of the relationship between water behavior inside the pores and overall MOF properties has only recently started to emerge. Molecular dynamics (MD) simulations with classical force fields suggested that the displacement of the benzene dicarboxylate (BDC) linkers coordinated to the Zn centers in MOF-5 (also known as IRMOF-1) was likely involved in the degradation process upon exposure to moisture.65 However, subsequent simulations performed with an empirically parameterized reactive force field indicated that water hydrolysis promoted by direct interactions of H2O molecules with the Zn centers of the framework was actually responsible for the collapse of the MOF-5 structure.66 Grand canonical Monte Carlo (GCMC) simulations were employed to study the mechanisms of water adsorption in MIL-100(Cr) and MIL-101(Cr).67 The breathing effect of MIL-53(Cr)

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 28

induced by water adsorption was characterized using both MD and GCMC simulations.60, 68 The impact of nuclear quantum effects62 and electronic polarization61, 68 on the properties of water adsorbed in the MIL-53(Cr) pores was investigated through molecular simulations. These studies predicted relatively stronger interactions between water molecules and the µ2-OH groups of the framework as well as a significantly slower water dynamics in the MOF pores compared to liquid water as a result of confinement.68 These predictions were confirmed by subsequent ab initio MD simulations based on density functional theory (DFT).69-70 More recently, MD simulations were used to reveal the microscopic mechanisms associated with water-mediated proton transport in MIL-53 as a function of temperature, water loading, and pore size.34 The structure of the hydrated proton was found to resemble that of a distorted Zundel complex when the MIL-53 framework was in the narrow-pore configuration. A transition to Eigen-like structures was then observed at higher water loading when the pores opened as a result of the breathing effect.34 In this study, MD simulations are used to investigate the structure and dynamics of water in the [Zn(γ-L)(X)] MOFs (with X = Cl and Br) of Ref. 64. As mentioned above, relatively high proton conductivity was observed upon water adsorption in the Cl-substituted framework but not in the Br-substituted isomer.64 The main focus of this study is on characterizing the relationship between framework properties and water behavior as a function of loading as well as on identifying specific hydrogen-bonding motifs and spatial arrangements of the water molecules in the pores, which are directly related to the different proton conductivities measured for [Zn(γL)(Cl)] and [Zn(γ-L)(Br)]. The paper is organized as follows: The computational methodology is presented in Section 2, the structural and dynamical properties of water inside the pores as a

ACS Paragon Plus Environment

4

Page 5 of 28

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

function of loading are discussed in Section 3, and a brief summary and outlook are given in Section 4. 2. COMPUTATIONAL METHODOLOGY Since the proton conductivities measured in Ref. 64 were found to be dependent on the specific halogen present on the framework and not related to the specific enantiomeric form of the MOF structure, molecular models were developed only for the [Zn(γ-L)(X)] with γ = l. Following our previous studies,47, 71 fully flexible force fields were developed for both [Zn(l-L)(Cl)] and [Zn(lL)(Br)] MOFs, and used in MD simulations aimed at characterizing the behavior of water in the MOF pores as a function of loading. The General Amber Force Field (GAFF) was used to model the intramolecular interactions of the L ligand.72 Specific parameterizations of all relevant Zn-L and Zn-X interactions were derived from fits to ab initio data obtained for the reduced MOF

Figure 1. The molecular model used in the parameterization of the flexible force field for [Zn(lL)(Cl)]. H atoms are in white, C atoms are in grey, N atoms are in blue, O atoms are in red, the Cl atom is in green, and the Zn atom is light violet (center). An analogous model with Br replacing Cl was used for [Zn(l-L)(Br)].

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

Page 6 of 28

model shown in Figure 1. All ab initio calculations were carried out with Gaussian 0973 at the DFT level using the M062X functional74 in combination with the cc-pVDZ basis set for H, C, O, and Cl,75 and the ccpVDZ-PP basis set for Zn.76 After performing energy minimizations on both Cl- and Brsubstituted reduced models, potential energy scans along the corresponding normal modes were carried out to map the underlying energy landscape. Fits to the potential energy curves corresponding to normal modes with frequencies below 500 cm-1 were then performed with a genetic algorithm77 to determine the force field parameters associated with the description of all bonds, angles, and dihedrals containing the Zn centers. The threshold of 500 cm-1 (719 K) was specifically chosen to guarantee an accurate description of the low-frequency modes of the framework, which determine the overall flexibility of the MOF structures at ambient conditions. The atomic partial charges were obtained from fits to the electrostatic potential of an extended model (see Supporting Information) calculated using the CHELPG method.78 The water interactions were described by the aSPC/Fw model79 and the force field parameters associated with the nonbonded framework-water interactions were derived from the Lorentz-Berthelot mixing rule.80 Although highly accurate ab initio water potentials have recently become available (e.g., see Refs. 81-88), their integration in MD simulations of MOFs is still prohibitive due to the existing incompatibility with the force fields currently used to describe the frameworks. Research in this area, involving the development of ab initio potentials for both inorganic and organic MOF subunits, is ongoing in our group. On the other hand, besides being computationally expensive give the system size and simulation lengths required for an accurate modeling of MOF properties, popular DFT models commonly used in ab initio molecular dynamics simulations have been shown not to be particularly accurate in describing the

ACS Paragon Plus Environment

6

Page 7 of 28

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

properties of water (e.g., see Ref. 89 for recent benchmark calculations). Among classical water models, the aSPC/Fw model was specifically chosen because it is the underlying water force field employed in the anharmonic multistate empirical valence bond (aMS-EVB3) model of protonated water,79 which, following Ref. 34, we are currently using to investigate proton conduction in both [Zn(l-L)(Cl)] and [Zn(l-L)(Br)]. Specific details about the fitting procedure along with the complete list of the force field parameters are reported in the Supporting Information. All MD simulations were performed with DL_POLY Classic90 on [Zn(l-L)(X)] structures consisting of 2 x 2 x 2 primitive cells in periodic boundary conditions. The short-range interactions were truncated at an atom-atom distance of 9.0 Å, while the electrostatic interactions were treated using the Ewald method.80 The MD simulations were carried out with a variable number of water molecules per primitive cell. The water molecules were initially distributed uniformly in the MOF pores, and each system was equilibrated for 5 ns through MD simulations carried out in the constant stress and constant temperature (NσT) ensemble. The production runs were then performed in the canonical (NVT) ensemble for 5 ns, which were used to calculate all structural and thermodynamic properties, as well as in the microcanonical (NVE) ensemble for an additional 1 ns, which was used to calculate all dynamical properties.91 The temperature and pressure were maintained via Nosé-Hoover thermostat and barostat with relaxation times of 1 and 5 ps, respectively. The equations of motion were propagated according to the velocity Verlet algorithm with a time step of 0.5 fs.80 The self-diffusion coefficient, D, of the water molecules in the pores of the two [Zn(l-L)(X)] MOFs was obtained from the Einstein relation

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

D = lim

rcom ( t ) − rcom ( 0 )

t→∞

where rcom ( t ) − rcom ( 0 )

2

Page 8 of 28

2

6t

(1)

is the center-of-mass (com) mean square displacement of each water

molecule and the angle bracket indicates an ensemble average over all water molecules. The time scales associated with the orientational dynamics of the water molecules inside the pores were extracted from exponential fits to the corresponding time decays of the orientational time autocorrelation function C 2 ( t ) = P2 e ( 0 ) ⋅ e ( t ) , where P2 e ( 0) ⋅ e ( t ) is the second Legendre polynomial of the angle spanned in time by the unit vector e(t) lying on one OH bond of each H2O molecule, and the brackets indicate an ensemble average over both OH bonds and water molecules. 3. RESULTS AND DISCUSSION 3.1 Thermodynamic and structural properties Typically Zn-based MOFs are susceptible to hydrolysis when exposed to water and become unstable even when exposed to low relative humidity.65-66, 92 In the [Zn(l-L)(X)] MOFs studied here, however, the halogen atoms of the framework have a protecting effect on the Zn centers thus providing H2O molecules with more favorable binding sites, which, in turn, enables the framework to remain intact upon hydration. Analogous to bare ions in solution, the metal-halide groups on the framework are known to be good hydrogen bond acceptors and have the ability to perturb the structure and dynamics of liquid water.93-98 The degree of interaction between the water molecules adsorbed in the MOF pores and the framework itself can be quantified by the heat of adsorption, ∆H, shown in Figure 2. In this study, ∆H = H(0) - H(N), where H(N) is the enthalpy of the MOF at a given water loading, N,

ACS Paragon Plus Environment

8

Page 9 of 28

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 2. Differential heat of adsorption for water in [Zn(l-L)(X)] with X = Cl (green) and Br (blue) calculated for N = 2, 4, 6, 8, 10, 12, 18, and 24 H2O molecules per pore. See main text for details.

and H(0) is the enthalpy of the completely dehydrated MOF. Both H(N) and H(0) were calculated from NσT simulations and N corresponds to the number of H2O molecules per pore. For water in [Zn(l-L)(Cl)], ∆H is ~10 kcal/mol at N = 2, increases up to ~12 kcal/mol at N = 6 before dropping significantly at N = 8 and then stabilizing at the approximately constant value of ~9 kcal/mol for N ≥ 10. By contrast the heat of adsorption of water in [Zn(l-L)(Br)] displays a larger variation, starting at ~5 kcal/mol at N = 2, reaching a maximum of ~16 kcal/mol at N = 8, before decreasing to ~1 kcal/mol at N = 24. The variation of ∆H with N can be rationalized by considering the specific Zn—X ⋯ H—OH interactions that exist inside the pores of the two isomeric MOFs, with the Zn—Cl ⋯ H—OH interaction being relatively stronger.96-98 This correlates with the increase of ∆H in [Zn(l-L)(Cl)] up to N = 6, corresponding to the number of Zn—Cl groups within a pore. At higher loading, when all Zn—Cl sites in the pores are saturated,

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 28

the additional water molecules must interact with less favorable binding sites, thus explaining the large drop of ∆H at N = 8. The subsequent increase of ∆H for water in [Zn(l-L)(Cl)] suggests the formation of a stable hydrogen-bond network between the H2O molecules adsorbed within the pores. The presence of weaker framework-water interactions in [Zn(l-L)(Br)] manifests itself in the low value of ∆H for N = 2, when all Br sites on the framework are effectively available for binding. As it will become more evident from the analysis of the spatial distribution of the water molecules inside the pores (Figure 3), the large value of ∆H predicted at N = 8 in [Zn(l-L)(Br)] is indicative of an optimal balance between water packing and water-framework interactions. As N increases, crowding repulsive effects between water molecules confined in the pores dominate and consequently lead to a reduction of the heat of adsorption. The low value of ∆H predicted for N = 24 in [Zn(L)(Br)] indicates a much lower affinity of water for the framework at high loadings. This is in agreement with variable-temperature single-crystal X-ray diffraction measurements that found water evaporating from [Zn(l-L)(Br)] at much lower temperature than from [Zn(l-L)(Cl)].64 Due to weaker framework-water interactions, the spatial arrangement of the H2O molecules in the [Zn(l-L)(Br)] pores is less ordered as demonstrated by the three-dimensional density distributions calculated for the water oxygen which are shown in Figure 3. From the inspection of the front view of the water density at N = 6, 12, 18, and 24, it is immediately apparent that well-defined pockets of dense regions arranged in a hexagonal pattern persist in the vicinity of the Zn—Cl groups at all levels of loading in [Zn(l-L)(Cl)]. The three-dimensional densities also show that, after binding to each of the six Zn—Cl groups within a pore, the water molecules start developing an inner hydrogen-bonded layer that propagates along the MOF channels. The

ACS Paragon Plus Environment

10

Page 11 of 28

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

templating effect of the [Zn(l-L)(Cl)] framework on the water distribution correlates with the heat of adsorption shown in Figure 2, providing direct evidence of different binding motifs within the pores as a function of N. By contrast, due to weaker framework-water interactions, H2O molecules adsorbed in the [Zn(l-L)(Br)] pores do not remain in stable formations, which thus results in more diffuse spatial distributions. Further insights into the spatial arrangement of the water molecules in the MOF pores are l

l

Figure 3. Three-dimensional density distributions of the water oxygen calculated for [Zn(lL)(Cl)] (first two columns) and [Zn(l-L)(Br)] last two columns as a function of the number of water molecules, N, per pore. The first and third columns show front views while the second and fourth columns show side views of the corresponding MOF channels. The coordinates of the framework atoms were averaged over 5 ns of MD simulation in the NVT ensemble. H atoms are in white, C atoms are in dark grey, N atoms are in blue, O atoms are in red, Cl atoms are in green, Br atoms are in orange, Zn atoms are light violet. The water densities (shown at an isosurface value of 0.16 Å-3) are in light blue. 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 28

gained from the analysis of the three-dimensional density distributions along the MOF channels. As mentioned above, the water molecules adsorbed in the [Zn(l-L)(Cl)] pores arrange in consecutive layers that are templated by the Zn—Cl groups and propagate along the MOF channels, effectively forming one-dimensional hydrogen-bonded chains that follow the l-helicity of the framework. While a roughly similar spatial arrangement is also found in [Zn(l-L)(Br)], the water molecules in this case fail to maintain a high level of order, resulting in more scattered density distributions. The present MD simulations also predict a higher degree of distortion of the [Zn(l-L)(Br)] framework upon water adsorption due to the intrinsic MOF flexibility as well as to the larger size of the Br atoms which effectively reduces the pore volume available to the water molecules, thus resulting in higher internal pressures for the same water loading. A detailed analysis of the effects of framework flexibility on the properties of MOFs upon water adsorption is currently being investigated by our lab. Quantitative information on the spatial distribution of the water molecules in the two homochiral MOFs can be obtained from the analysis of the radial distribution functions (RDFs) shown in Figure 4, which directly report on unique structural features depending on the Zn—X groups present on the framework. The Ow-Cl RDFs, describing the spatial correlations between the oxygen atoms (Ow) of the H2O molecules and the Cl atoms of the framework, display distinct peaks at ~3.4 Å, ~5.8 Å and ~8.0 Å, arising from the ordered arrangement of the water molecules inside the [Zn(l-L)(Cl)] pores which is templated by the helical periodicity of the M— Cl groups along the MOF channels. In contrast, the analogous Ow-Br RDFs are less structured, displaying a main peak at ~3.4 Å and a smaller peak at ~4.7 Å, which gradually disappears as N increases. The dependence of the two Ow-X RDFs on water loading directly reflects the relative strengths of the underlying framework-water interactions, with the stronger Zn—Cl ⋯ H—OH

ACS Paragon Plus Environment

12

Page 13 of 28

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

interaction being capable of effectively holding the water molecules at fixed distances from each other. As shown in Figure 3, once all Zn—X sites are hydrated, the additional water molecules are unable of interacting directly with the framework and begin filling the inner region of the MOF channels. As a consequence, the two Ow-X RDFs become more similar at higher loadings. The different hydrogen-bonding patterns that the water molecules establish in the two [Zn(lL)(X)] MOFs as a function of loading clearly emerge from the analysis of the Ow-Ow RDFs. Specifically, a single peak at ~6.2 Å is found at low loadings in [Zn(l-L)(Cl)], corresponding to the shortest distance between water molecules bound to adjacent Zn—Cl groups. In contrast, two

Figure 4. Ow-Cl (panel a), Ow-Br (panel b), and Ow-Ow (panels c and d) radial distribution functions calculated for [Zn(l-L)(Cl)] (left column) and [Zn(l-L)(Br)] (right column) as a function of the number of water molecules, N, per pore. See main text for details.

ACS Paragon Plus Environment

13

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 14 of 28

peaks are found at ~2.8 Å and ~5.4 Å for the Ow-Ow RDF in [Zn(l-L)(Br)]. These two peaks result from the competition between framework-water and water-water interactions, with the peak at the shortest distance corresponding to water molecules directly hydrogen-bonded to each other. In both MOFs, the first Ow-Ow peak at ~2.8 Å becomes more pronounced as a function of water loading, reflecting the increasing number of H2O molecules that establish hydrogen bonds between each other. This is accompanied by the simultaneous development of a broad peak between 4.0 Å and 5.0 Å, which is associated with hydrogen-bonded water molecules located within different layers inside the pores (see Figure 3). 3.2 Dynamical properties Despite being homochiral MOFs, the presence of different halogen atoms in the framework results in distinct dynamical properties of the water molecules adsorbed in the pores. The variation of the overall water diffusion coefficient (blue), along with the corresponding components, parallel (red) and perpendicular (green) to the MOF channels, is shown in Figure 5

Figure 5. Water diffusion coefficient (D) and its components perpendicular (Dx,y) and parallel (Dz) to the MOF channels calculated in [Zn(l-L)(Cl)] (panel a) and [Zn(l-L)(Br)] (panel b) as a function of number of water molecules, N, per pore. Statistical uncertainties are within the size of the symbols. See main text for details.

ACS Paragon Plus Environment

14

Page 15 of 28

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

as a function of loading. As a reference, the aSPC/Fw model predicts a self diffusion coefficient of 0.233 Å2 ps-1 for bulk water, in good agreement with the experimental value (0.229 Å2 ps-1).79 The diffusion of water in [Zn(l-L)(Br)] is effectively only contingent upon the amount of water present in the pores and increases as a function of loading (panel a). By contrast, up to N = 6, which corresponds to the minimum number of water molecules necessary to saturate all Zn—Cl groups within a pore, the diffusion coefficient of water in [Zn(l-L)(Cl)] is extremely small, indicating that the water molecules are effectively immobile at low loadings. Once all Zn—Cl groups are saturated, the diffusion coefficient increases incrementally with N as the water molecules begin filling the inner region of the MOF channels where they can move relatively more freely, only impeded by the surrounding hydrogen-bond network. As expected, due to the confining environment provided by the framework, the diffusion of H2O molecules along the MOF channels, Dz, is significantly faster than the diffusion along perpendicular directions, Dx,y. Molecular-level insights into the dynamics of the hydrogen-bond network established inside the MOF pores can be derived from the analysis of the water orientational correlation function,

Figure 6. Orientational correlation functions, C2(t), calculated for water in [Zn(l-L)(Cl)] (panel a) and [Zn(l-L)(Br)] (panel b) as a function of number of H2O molecules, N, per pore. See main text for details.

ACS Paragon Plus Environment

15

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 16 of 28

C2(t) introduced in Section 2, which is proportional to the experimentally measured orientational anisotropy.99 In particular, the intermediate time scale of orientational relaxation directly reflects the breaking and forming of intermolecular hydrogen bonds, whereas the short and long time decays of C2(t) can be related to librational motions and complete structural randomization, respectively. In this study, the time decay of C2(t), shown in Figure 6 for different water loadings in [Zn(l-L)(Cl)] and [Zn(l-L)(Br)], were fitted to a tri-exponential function

C2 ( t ) = A1e −t τ1 + A 2 e−t τ2 + A 3e−t τ3

(2)

where Ai are the amplitudes and τi are the relaxation time constants associated with each exponential, which are reported in Table 1. After an initial fast drop, C2(t) displays a different time decay as a function of the number of water molecules per pore in the two MOFs. In fact, the different behavior of water in the two MOFs at low loadings can be directly determined from the analysis of the second relaxation time, τ2, associated with the formation and cleavage of hydrogen bonds. At low loadings, water molecules within [Zn(l-L)(Cl)] reorient at much slower rates due to the relatively stronger hydrogen bonds formed with the Zn—Cl groups, with τ2 ranging from 50 to 100 ps when N ≤ 6. This critically retarded hydrogen-bond dynamics persists until addition of more water molecules (N ≥ 8), which, as mentioned above, results in the formation of a second concentric and relatively more mobile water layer extending along the MOF channels. In contrast, as a result of the weaker framework-water interactions, the values of τ2 for the water reorientation within the [Zn(l-L)(Br)] pores display a much smaller variation with N, being always less than 25 ps independently of loading. The calculated τ2 values thus show that, in both MOFs, the formation and cleavage of hydrogen bonds between water molecules occurs on significantly slower time scales than in the liquid phase (τ2 = 2.7 ps for aSPC/Fw water at ambient conditions79). The suppressed water mobility inside the pores thus

ACS Paragon Plus Environment

16

Page 17 of 28

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

Table 1. Relaxation time constants, τI, obtained from fits to orientational correlation functions calculated for water in [Zn(l-L)(Cl)] and [Zn(l-L)(Br)] as a function of number

of H2O molecules per pore. [Zn(l-L)(Cl)]

[Zn(l-L)(Br)]

N

τ1 (ps)

τ2 (ps)

τ3 (ps)

τ1 (ps)

τ2 (ps)

τ3 (ps)

2

0.06

94

> 1000

< 0.01

22

315

4

0.05

50

300

< 0.01

18

385

6

0.05

100

750

< 0.01

22

480

8

< 0.01

16

180

< 0.01

19

450

10

< 0.01

25

215

< 0.01

15

410

12

< 0.01

14

180

< 0.01

20

390

18

< 0.01

16

210

< 0.01

15

260

24

< 0.01

10

90

< 0.01

10

160

suggests that proton conduction in the two MOFs can mainly occur through proton hopping between adjacent molecules according to the Grotthuss mechanism, which can explain the higher conductivity measured for [Zn(l-L)(Cl)] due to the presence of ordered water chains that extend along the MOF channels. Molecular-level studies of proton conduction in both [Zn(l-L)(Cl)] and [Zn(l-L)(Br)] are currently ongoing in our group. Finally, it is worth noting the large values of τ3, which are consistent with a very slow randomization of the orientation of the water molecules inside the pores due to the extreme confining effects of the frameworks. 4. CONCLUSIONS Understanding the molecular mechanisms that govern water adsorption, structure, and mobility in MOFs is key to the rational design of de novo hydrothermal stable frameworks for large-scale

ACS Paragon Plus Environment

17

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 18 of 28

applications. Recent experimental measurements have shown that water adsorbed in the pores can efficiently mediate proton conduction in MOFs at relatively low temperature and humidity, although the underlying driving forces are not yet fully understood. In this study, molecular dynamics simulations were employed to characterize the behavior of water in two homochiral MOFs with general formula [Zn(l-L)(X)], X = Cl and Br. A direct correlation was found between the relative strength of the framework-water interactions and the structural and dynamical properties of the H2O molecules adsorbed in the MOF pores. In particular, the water molecules adsorbed in the [Zn(l-L)(Cl)] pores arrange in consecutive layers that are templated by the Zn— Cl groups and propagate along the MOF channels, effectively forming one-dimensional hydrogen-bonded chains that follow the l-helicity of the framework. In contrast, the relatively weaker framework-water interactions in [Zn(l-L)(Br)] were found to lead to less ordered water formations inside the pores. The slow dynamics predicted for water molecules adsorbed in the pores suggests that proton conduction in the two MOFs can mainly occur through proton hopping between adjacent molecules according to the Grotthuss mechanism, which supports the view derived in Ref. 64 based on the experimentally measured conductivities. ASSOCIATED CONTENT Supporting Information. Details of the fitting procedure used in the development of the flexible force fields for the two [Zn(l-L)(X)] MOFs, including a complete list of the force field parameters. This material is available free of charge via the Internet at http://pubs.acs.org. AUTHOR INFORMATION Corresponding Author * [email protected]

ACS Paragon Plus Environment

18

Page 19 of 28

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

Author Contributions ‡ These authors contributed equally. Funding Sources National Science Foundation, DMR-1305101 and ACI-1053575 U.S. Department of Energy, DE-FG02- 13ER16387 and DE-AC02-05CH11231

ACKNOWLEDGMENTS We wish to thank Mr. Adil Mohd-Salleh for helpful discussion at the early stages of this study. This research was supported by the National Science Foundation (Award Number DMR1305101) and the U.S. Department of Energy, Office of Science, under Award No. DE-FG0213ER16387, and used resources of the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by the National Science Foundation Grant Number ACI-1053575 (Allocation TG-CHE110009) as well as of the National Energy Research Scientific Computing Center, which is supported by the Office of Science of the U.S. Department of Energy under Contract DE-AC02-05CH11231.

REFERENCES 1.

Wang, X. S., et al., Metal-Organic Frameworks Based on Double-Bond-Coupled DiIsophthalate Linkers with High Hydrogen and Methane Uptakes. Chem. Mater. 2008, 20, 3145-3152.

2.

Wu, H., et al., Metal-Organic Frameworks with Exceptionally High Methane Uptake: Where and How Is Methane Stored? Chem. Eur. J. 2010, 16, 5205-5214.

3.

Getman, R. B.; Bae, Y. S.; Wilmer, C. E.; Snurr, R. Q., Review and Analysis of Molecular Simulations of Methane, Hydrogen, and Acetylene Storage in Metal-Organic Frameworks. Chem. Rev. 2012, 112, 703-723.

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 28

4.

Kennedy, R. D., et al., Carborane-Based Metal-Organic Framework with High Methane and Hydrogen Storage Capacities. Chem. Mater. 2013, 25, 3539-3543.

5.

Peng, Y.; Krungleviciute, V.; Eryazici, I.; Hupp, J. T.; Farha, O. K.; Yildirim, T., Methane Storage in Metal-Organic Frameworks: Current Records, Surprise Findings, and Challenges. J. Am. Chem. Soc. 2013, 135, 11887-11894.

6.

Wilmer, C. E.; Farha, O. K.; Yildirim, T.; Eryazici, I.; Krungleviciute, V.; Sarjeant, A. A.; Snurr, R. Q.; Hupp, J. T., Gram-Scale, High-Yield Synthesis of a Robust Metal-Organic Framework for Storing Methane and Other Gases. Energy. Environ. Sci. 2013, 6, 11581163.

7.

Gandara, F.; Furukawa, H.; Lee, S.; Yaghi, O. M., High Methane Storage Capacity in Aluminum Metal-Organic Frameworks. J. Am. Chem. Soc. 2014, 136, 5271-5274.

8.

Gomez-Gualdron, D. A.; Gutov, O. V.; Krungleviciute, V.; Borah, B.; Mondloch, J. E.; Hupp, J. T.; Yildirim, T.; Farha, O. K.; Snurr, R. Q., Computational Design of MetalOrganic Frameworks Based on Stable Zirconium Building Units for Storage and Delivery of Methane. Chem. Mater. 2014, 26, 5632-5639.

9.

He, Y. B.; Zhou, W.; Qian, G. D.; Chen, B. L., Methane Storage in Metal-Organic Frameworks. Chem. Soc. Rev. 2014, 43, 5657-5678.

10.

Yan, Y.; Yang, S. H.; Blake, A. J.; Schroder, M., Studies on Metal-Organic Frameworks of Cu(Ii) with Isophthalate Linkers for Hydrogen Storage. Acc. Chem. Res. 2014, 47, 296-307.

11.

Suh, M. P.; Park, H. J.; Prasad, T. K.; Lim, D. W., Hydrogen Storage in Metal-Organic Frameworks. Chem. Rev. 2012, 112, 782-835.

12.

Britt, D.; Furukawa, H.; Wang, B.; Glover, T. G.; Yaghi, O. M., Highly Efficient Separation of Carbon Dioxide by a Metal-Organic Framework Replete with Open Metal Sites. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 20637-20640.

13.

Demessence, A.; D'Alessandro, D. M.; Foo, M. L.; Long, J. R., Strong Co2 Binding in a Water-Stable, Triazolate-Bridged Metal-Organic Framework Functionalized with Ethylenediamine. J. Am. Chem. Soc. 2009, 131, 8784-8786.

14.

Queen, W. L., et al., Comprehensive Study of Carbon Dioxide Adsorption in the MetalOrganic Frameworks M-2(Dobdc) (M = Mg, Mn, Fe, Co, Ni, Cu, Zn). Chem. Sci. 2014, 5, 4569-4581.

15.

Mason, J. A.; Veenstra, M.; Long, J. R., Evaluating Metal-Organic Frameworks for Natural Gas Storage. Chem. Sci. 2014, 5, 32-51.

ACS Paragon Plus Environment

20

Page 21 of 28

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

16.

Lin, L. C.; Kim, J.; Kong, X. Q.; Scott, E.; McDonald, T. M.; Long, J. R.; Reimer, J. A.; Smit, B., Understanding Co2 Dynamics in Metal-Organic Frameworks with Open Metal Sites. Angew. Chem. Int. Ed. 2013, 52, 4410-4413.

17.

McDonald, T. M., et al., Cooperative Insertion of Co2 in Diamine-Appended MetalOrganic Frameworks. Nature 2015, 519, 303-308.

18.

Bae, Y. S.; Lee, C. Y.; Kim, K. C.; Farha, O. K.; Nickias, P.; Hupp, J. T.; Nguyen, S. T.; Snurr, R. Q., High Propene/Propane Selectivity in Isostructural Metal-Organic Frameworks with High Densities of Open Metal Sites. Angew. Chem. Int. Ed. 2012, 51, 1857-1860.

19.

Herm, Z. R.; Wiers, B. M.; Mason, J. A.; van Baten, J. M.; Hudson, M. R.; Zajdel, P.; Brown, C. M.; Masciocchi, N.; Krishna, R.; Long, J. R., Separation of Hexane Isomers in a Metal-Organic Framework with Triangular Channels. Science 2013, 340, 960-964.

20.

Duren, T.; Snurr, R. Q., Assessment of Isoreticular Metal-Organic Frameworks for Adsorption Separations: A Molecular Simulation Study of Methane/N-Butane Mixtures. J. Phys. Chem. B 2004, 108, 15703-15708.

21.

Ma, L. Q.; Abney, C.; Lin, W. B., Enantioselective Catalysis with Homochiral MetalOrganic Frameworks. Chem. Soc. Rev. 2009, 38, 1248-1256.

22.

Ma, L. Q.; Falkowski, J. M.; Abney, C.; Lin, W. B., A Series of Isoreticular Chiral MetalOrganic Frameworks as a Tunable Platform for Asymmetric Catalysis. Nat. Chem. 2010, 2, 838-846.

23.

Wang, C.; Zheng, M.; Lin, W., Asymmetric Catalysis with Chiral Porous Metal-Organic Frameworks: Critical Issues. J. Phys. Chem. Lett. 2011, 2, 1701-1709.

24.

Yoon, M.; Srirambalaji, R.; Kim, K., Homochiral Metal-Organic Frameworks for Asymmetric Heterogeneous Catalysis. Chem. Rev. 2012, 112, 1196-1231.

25.

Liu, J. W.; Chen, L. F.; Cui, H.; Zhang, J. Y.; Zhang, L.; Su, C. Y., Applications of MetalOrganic Frameworks in Heterogeneous Supramolecular Catalysis. Chem. Soc. Rev. 2014, 43, 6011-6061.

26.

Zhao, M.; Ou, S.; Wu, C. D., Porous Metal-Organic Frameworks for Heterogeneous Biomimetic Catalysis. Acc. Chem. Res. 2014, 47, 1199-1207.

27.

McGuirk, C. M.; Katz, M. J.; Stern, C. L.; Sarjeant, A. A.; Hupp, J. T.; Farha, O. K.; Mirkin, C. A., Turning on Catalysis: Incorporation of a Hydrogen-Bond-Donating Squaramide Moiety into a Zr Metal-Organic Framework. J. Am. Chem. Soc. 2015, 137, 919-925.

28.

Sheberla, D.; Sun, L.; Blood-Forsythe, M. A.; Er, S.; Wade, C. R.; Brozek, C. K.; AspuruGuzik, A.; Dinca, M., High Electrical Conductivity in Ni-3(2,3,6,7,10,11-

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 28

Hexaiminotriphenylene)(2), a Semiconducting Metal-Organic Graphene Analogue. J. Am. Chem. Soc. 2014, 136, 8859-8862. 29.

Talin, A. A., et al., Tunable Electrical Conductivity in Metal-Organic Framework ThinFilm Devices. Science 2014, 343, 66-69.

30.

Hurd, J. A.; Vaidhyanathan, R.; Thangadurai, V.; Ratcliffe, C. I.; Moudrakovski, I. L.; Shimizu, G. K. H., Anhydrous Proton Conduction at 150 Degrees C in a Crystalline MetalOrganic Framework. Nat. Chem. 2009, 1, 705-710.

31.

Shimizu, G. K. H.; Taylor, J. M.; Kim, S., Proton Conduction with Metal-Organic Frameworks. Science 2013, 341, 354-355.

32.

Liang, X. Q.; Zhang, F.; Feng, W.; Zou, X. Q.; Zhao, C. J.; Na, H.; Liu, C.; Sun, F. X.; Zhu, G. S., From Metal-Organic Framework (Mof) to Mof-Polymer Composite Membrane: Enhancement of Low-Humidity Proton Conductivity. Chem. Sci. 2013, 4, 983-992.

33.

Okawa, H.; Sadakiyo, M.; Yamada, T.; Maesato, M.; Ohba, M.; Kitagawa, H., ProtonConductive Magnetic Metal-Organic Frameworks, {NR3(CH2COOH)} [MaIIMbIII(ox)3]: Effect of Carboxyl Residue Upon Proton Conduction. J. Am. Chem. Soc. 2013, 135, 22562262.

34.

Paesani, F., Molecular Mechanisms of Water-Mediated Proton Transport in Mil-53 MetalOrganic Frameworks. J. Phys. Chem. C 2013, 117, 19508-19516.

35.

Taylor, J. M.; Dawson, K. W.; Shimizu, G. K. H., A Water-Stable Metal-Organic Framework with Highly Acidic Pores for Proton-Conducting Applications. J. Am. Chem. Soc. 2013, 135, 1193-1196.

36. Nagarkar, S. S.; Unni, S. M.; Sharma, A.; Kurungot, S.; Ghosh, S. K., Two-in-One: Inherent Anhydrous and Water-Assisted High Proton Conduction in a 3d Metal-Organic Framework. Angew. Chem. Int. Ed. 2014, 53, 2638-2642. 37.

Ramaswamy, P.; Wong, N. E.; Shimizu, G. K. H., Mofs as Proton Conductors - Challenges and Opportunities. Chem. Soc. Rev. 2014, 43, 5913-5932.

38.

Bureekaew, S.; Horike, S.; Higuchi, M.; Mizuno, M.; Kawamura, T.; Tanaka, D.; Yanai, N.; Kitagawa, S., One-Dimensional Imidazole Aggregate in Aluminium Porous Coordination Polymers with High Proton Conductivity. Nat. Mater. 2009, 8, 831-836.

39.

Sadakiyo, M.; Yamada, T.; Kitagawa, H., Rational Designs for Highly Proton-Conductive Metal-Organic Frameworks. J. Am. Chem. Soc. 2009, 131, 9906-9907.

40.

Yamada, T.; Sadakiyo, M.; Kitagawa, H., High Proton Conductivity of One-Dimensional Ferrous Oxalate Dihydrate. J. Am. Chem. Soc. 2009, 131, 3144-3145.

ACS Paragon Plus Environment

22

Page 23 of 28

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

41.

Sen, S.; Nair, N. N.; Yamada, T.; Kitagawa, H.; Bharadwaj, P. K., High Proton Conductivity by a Metal Organic Framework Incorporating Zn8o Clusters with Aligned Imidazolium Groups Decorating the Channels. J. Am. Chem. Soc. 2012, 134, 19432-19437.

42.

Yamada, T.; Otsubo, K.; Makiura, R.; Kitagawa, H., Designer Coordination Polymers: Dimensional Crossover Architectures and Proton Conduction. Chem. Soc. Rev. 2013, 42, 6655-6669.

43.

Ohba, M., et al., Bidirectional Chemo-Switching of Spin State in a Microporous Framework. Angew. Chem. Int. Ed. 2009, 48, 4767-4771.

44.

Okawa, H.; Shigematsu, A.; Sadakiyo, M.; Miyagawa, T.; Yoneda, K.; Ohba, M.; Kitagawa, H., Oxalate-Bridged Bimetallic Complexes {NH(prol)3}MCr(ox)3 (M = MnII, FeII, CoII; NH(prol)3+ = Tri(3-hydroxypropyl)ammonium) Exhibiting Coexistent Ferromagnetism and Proton Conduction. J. Am. Chem. Soc. 2009, 131, 13516-13522.

45.

Ohtani, R.; Yoneda, K.; Furukawa, S.; Horike, N.; Kitagawa, S.; Gaspar, A. B.; Munoz, M. C.; Real, J. A.; Ohba, M., Precise Control and Consecutive Modulation of Spin Transition Temperature Using Chemical Migration in Porous Coordination Polymers. J. Am. Chem. Soc. 2011, 133, 8600-8605.

46.

Aravena, D., et al., Guest Modulation of Spin-Crossover Transition Temperature in a Porous Iron(II) Metal-Organic Framework: Experimental and Periodic Dft Studies. Chem. Eur. J. 2014, 20, 12864-12873.

47.

Cirera, J.; Babin, V.; Paesani, F., Theoretical Modeling of Spin Crossover in Metal-Organic Frameworks: Fe(Pz)2Pt(CN)4 as a Case Study. Inorg. Chem. 2014, 53, 11020-11028.

48.

Bartual-Murgui, C.; Akou, A.; Thibault, C.; Molnar, G.; Vieu, C.; Salmon, L.; Bousseksou, A., Spin-Crossover Metal-Organic Frameworks: Promising Materials for Designing Gas Sensors. J. Mater. Chem. C 2015, 3, 1277-1285.

49.

Allendorf, M. D.; Bauer, C. A.; Bhakta, R. K.; Houk, R. J. T., Luminescent Metal-Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1330-1352.

50.

Canivet, J.; Fateeva, A.; Guo, Y. M.; Coasne, B.; Farrusseng, D., Water Adsorption in Mofs: Fundamentals and Applications. Chem. Soc. Rev. 2014, 43, 5594-5617.

51.

Burtch, N. C.; Jasuja, H.; Walton, K. S., Water Stability and Adsorption in Metal-Organic Frameworks. Chem. Rev. 2014, 114, 10575-10612.

52.

Horcajada, P., et al., Porous Metal-Organic-Framework Nanoscale Carriers as a Potential Platform for Drug Delivery and Imaging. Nat. Mater. 2010, 9, 172-178.

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

Page 24 of 28

53.

McKinlay, A. C.; Morris, R. E.; Horcajada, P.; Ferey, G.; Gref, R.; Couvreur, P.; Serre, C., Biomofs: Metal-Organic Frameworks for Biological and Medical Applications. Angew. Chem. Int. Ed. 2010, 49, 6260-6266.

54.

Miller, S. R.; Heurtaux, D.; Baati, T.; Horcajada, P.; Greneche, J. M.; Serre, C., Biodegradable Therapeutic Mofs for the Delivery of Bioactive Molecules. Chem. Commun. 2010, 46, 4526-4528.

55.

Horcajada, P.; Gref, R.; Baati, T.; Allan, P. K.; Maurin, G.; Couvreur, P.; Ferey, G.; Morris, R. E.; Serre, C., Metal-Organic Frameworks in Biomedicine. Chem. Rev. 2012, 112, 1232-1268.

56.

Soubeyrand-Lenoir, E.; Vagner, C.; Yoon, J. W.; Bazin, P.; Ragon, F.; Hwang, Y. K.; Serre, C.; Chang, J. S.; Llewellyn, P. L., How Water Fosters a Remarkable 5-Fold Increase in Low-Pressure Co2 Uptake within Mesoporous Mil-100(Fe). J. Am. Chem. Soc. 2012, 134, 10174-10181.

57.

Yazaydin, A. O.; Benin, A. I.; Faheem, S. A.; Jakubczak, P.; Low, J. J.; Willis, R. R.; Snurr, R. Q., Enhanced Co2 Adsorption in Metal-Organic Frameworks Via Occupation of Open-Metal Sites by Coordinated Water Molecules. Chem. Mater. 2009, 21, 1425-1430.

58.

Chen, Y. F.; Babarao, R.; Sandler, S. I.; Jiang, J. W., Metal Organic Framework Mil-101 for Adsorption and Effect of Terminal Water Molecules: From Quantum Mechanics to Molecular Simulation. Langmuir 2010, 26, 8743-8750.

59.

Serre, C.; Millange, F.; Thouvenot, C.; Nogues, M.; Marsolier, G.; Louer, D.; Ferey, G., Very Large Breathing Effect in the First Nanoporous Chromium(III)-Based Solids: MIL-53 or Cr-III(OH)·{O2C-C6H4-CO2}·{HO2C-C6H4-CO2H}x·H2Oy. J. Am. Chem. Soc. 2002, 124, 13519-13526.

60.

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.

61.

Cirera, J.; Sung, J. C.; Howland, P. B.; Paesani, F., The Effects of Electronic Polarization on Water Adsorption in Metal-Organic Frameworks: H2O in MIL-53(Cr). J. Chem. Phys. 2012, 137.

62.

Paesani, F., Water in Metal-Organic Frameworks: Structure and Diffusion of H2o in Mil53(Cr) from Quantum Simulations. Mol. Simul. 2012, 38, 631-641.

63.

Shigematsu, A.; Yamada, T.; Kitagawa, H., Wide Control of Proton Conductivity in Porous Coordination Polymers. J. Am. Chem. Soc. 2011, 133, 2034-2036.

ACS Paragon Plus Environment

24

Page 25 of 28

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

64.

Sahoo, S. C.; Kundu, T.; Banerjee, R., Helical Water Chain Mediated Proton Conductivity in Homochiral Metal-Organic Frameworks with Unprecedented Zeolitic Unh-Topology. J. Am. Chem. Soc. 2011, 133, 17950-17958.

65.

Greathouse, J. A.; Allendorf, M. D., The Interaction of Water with Mof-5 Simulated by Molecular Dynamics. J. Am. Chem. Soc. 2006, 128, 10678-10679.

66.

Han, S. S.; Choi, S. H.; van Duin, A. C. T., Molecular Dynamics Simulations of Stability of Metal-Organic Frameworks against H2o Using the ReaxFF Reactive Force Field. Chem. Commun. 2010, 46, 5713-5715.

67.

De Lange, M. F.; Gutierrez-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.

68.

Medders, G. R.; Paesani, F., Water Dynamics in Metal-Organic Frameworks: Effects of Heterogeneous Confinement Predicted by Computational Spectroscopy. J. Phys. Chem. Lett. 2014, 5, 2897-2902.

69.

Haigis, V.; Coudert, F. X.; Vuilleumier, R.; Boutin, A., Investigation of Structure and Dynamics of the Hydrated Metal-Organic Framework MIL-53(Cr) Using First-Principles Molecular Dynamics. Phys. Chem. Chem. Phys. 2013, 15, 19049-19056.

70.

Salazar, J. M.; Weber, G.; Simon, J. M.; Bezverkhyy, I.; Bellat, J. P., Characterization of Adsorbed Water in MIL-53(Al) by Ftir Spectroscopy and Ab-Initio Calculations. J. Chem. Phys. 2015, 142, 124702.

71.

Grosch, J. S.; Paesani, F., Molecular-Level Characterization of the Breathing Behavior of the Jungle-Gym-Type Dmof-1 Metal-Organic Framework. J. Am. Chem. Soc. 2012, 134, 4207-4215.

72.

Wang, J. M.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A., Development and Testing of a General Amber Force Field. J. Comput. Chem. 2004, 25, 1157-1174.

73.

Frisch, M. J., et al. Gaussian 09, Gaussian, Inc.: Wallingford, CT, USA, 2009.

74.

Zhao, Y.; Truhlar, D. G., The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120, 215-241.

75.

Dunning, T. H., Gaussian-Basis Sets for Use in Correlated Molecular Calculations .1. The Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90, 1007-1023.

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 28

76.

Peterson, K. A.; Puzzarini, C., Systematically Convergent Basis Sets for Transition Metals. Ii. Pseudopotential-Based Correlation Consistent Basis Sets for the Group 11 (Cu, Ag, Au) and 12 (Zn, Cd, Hg) Elements. Theor. Chem. Acc. 2005, 114, 283-296.

77.

Goldberg, D. E., Genetic Algorithms in Search, Optimization and Machine Learning, 1st ed.; Addison-Wesley Longman Publishing Co., Inc.: Boston, MA, U.S.A., 1989.

78.

Breneman, C. M.; Wiberg, K. B., Determining Atom-Centered Monopoles from Molecular Electrostatic Potentials - the Need for High Sampling Density in Formamide Conformational-Analysis. J. Comput. Chem. 1990, 11, 361-373.

79.

Park, K.; Lin, W.; Paesani, F., A Refined Ms-Evb Model for Proton Transport in Aqueous Environments. J. Phys. Chem. B 2012, 116, 343-352.

80.

Leach, A. R., Molecular Modeling: Principles and Applications, 2nd ed.; Pearson Prentice Hall, 2001.

81.

Bukowski, R.; Szalewicz, K.; Groenenboom, G. C.; van der Avoird, A., Predictions of the Properties of Water from First Principles. Science 2007, 315, 1249-1252.

82.

Bukowski, R.; Szalewicz, K.; Groenenboom, G. C.; van der Avoird, A., Polarizable Interaction Potential for Water from Coupled Cluster Calculations. I. Analysis of Dimer Potential Energy Surface. J. Chem. Phys. 2008, 128, 094313.

83.

Bukowski, R.; Szalewicz, K.; Groenenboom, G. C.; van der Avoird, A., Polarizable Interaction Potential for Water from Coupled Cluster Calculations. Ii. Applications to Dimer Spectra, Virial Coefficients, and Simulations of Liquid Water. J. Chem. Phys. 2008, 128.

84.

Wang, Y. M.; Shepler, B. C.; Braams, B. J.; Bowman, J. M., Full-Dimensional, Ab Initio Potential Energy and Dipole Moment Surfaces for Water. J. Chem. Phys. 2009, 131.

85.

Babin, V.; Medders, G. R.; Paesani, F., Toward a Universal Water Model: First Principles Simulations from the Dimer to the Liquid Phase. J. Phys. Chem. Lett. 2012, 3, 3765-3769.

86.

Babin, V.; Leforestier, C.; Paesani, F., Development of a "First Principles" Water Potential with Flexible Monomers: Dimer Potential Energy Surface, Vrt Spectrum, and Second Virial Coefficient. J. Chem. Theory. Comput. 2013, 9, 5395-5403.

87.

Babin, V.; Medders, G. R.; Paesani, F., Development of a "First Principles" Water Potential with Flexible Monomers. Ii: Trimer Potential Energy Surface, Third Virial Coefficient, and Small Clusters. J. Chem. Theory. Comput. 2014, 10, 1599-1607.

88.

Medders, G. R.; Babin, V.; Paesani, F., Development of a "First-Principles" Water Potential with Flexible Monomers. Iii. Liquid Phase Properties. J. Chem. Theory. Comput. 2014, 10, 2906-2910.

ACS Paragon Plus Environment

26

Page 27 of 28

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

89.

Morales, M. A.; Gergely, J. R.; McMinis, J.; McMahon, J. M.; Kim, J.; Ceperley, D. M., Quantum Monte Carlo Benchmark of Exchange-Correlation Functionals for Bulk Water. J. Chem. Theory Comput. 2014, 10, 2355-2362.

90.

Todorov, I. T.; Smith, W.; Trachenko, K.; Dove, M. T., Dl_Poly_3: New Dimensions in Molecular Dynamics Simulations Via Massive Parallelism. J. Mater. Chem. 2006, 16, 1911-1918.

91.

Tuckerman, M. E., Statistical Mechanics: Theory and Molecular Simulation; Oxford University Press, 2010.

92.

Chen, Z. X.; Xiang, S. C.; Zhao, D. Y.; Chen, B. L., Reversible Two-Dimensional-Three Dimensional Framework Transformation within a Prototype Metal-Organic Framework. Cryst. Growth Des. 2009, 9, 5293-5296.

93.

Bakker, H. J.; Kropman, M. F.; Omta, A. W., Effect of Ions on the Structure and Dynamics of Liquid Water. J. Phys.: Condens. Matt. 2005, 17, S3215-S3224.

94.

Marcus, Y., Effect of Ions on the Structure of Water. Pure Appl. Chem. 2010, 82, 18891899.

95.

Grossfield, A., Dependence of Ion Hydration on the Sign of the Ion's Charge. J. Chem. Phys. 2005, 122.

96.

Brammer, L.; Bruton, E. A.; Sherwood, P., Fluoride Ligands Exhibit Marked Departures from the Hydrogen Bond Acceptor Behavior of Their Heavier Halogen Congeners. New J. Chem. 1999, 23, 965-968.

97.

Brammer, L.; Bruton, E. A.; Sherwood, P., Understanding the Behavior of Halogens as Hydrogen Bond Acceptors. Cryst. Growth Des. 2001, 1, 277-290.

98.

Aullon, G.; Bellamy, D.; Brammer, L.; Bruton, E. A.; Orpen, A. G., Metal-Bound Chlorine Often Accepts Hydrogen Bonds. Chem. Commun. 1998, 653-654.

99.

Bakker, H. J.; Skinner, J. L., Vibrational Spectroscopy as a Probe of Structure and Dynamics in Liquid Water. Chem. Rev. 2010, 110, 1498-1517.

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

Page 28 of 28

Table of Contents Graphic

ACS Paragon Plus Environment

28