Sie sind auf Seite 1von 33

Subscriber access provided by - Access paid by the | UCSB Libraries

Interfaces: Adsorption, Reactions, Films, Forces, Measurement Techniques, Charge


Transfer, Electrochemistry, Electrocatalysis, Energy Production and Storage
Investigation of the Water Adsorption Properties and
Structural Stability of MIL-100(Fe) with Different Anions
Yen-Ru Chen, Kai-Hsin Liou, Dun-Yen Kang, Jiun-Jen Chen, and Li-Chiang Lin
Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.7b04399 • Publication Date (Web): 13 Mar 2018
Downloaded from http://pubs.acs.org on March 15, 2018

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 32 Langmuir

1
2
3 Investigation of the Water Adsorption Properties and
4
5
6 Structural Stability of MIL-100(Fe) with Different Anions
7
8 Yen-Ru Chen,1 Kai-Hsin Liou,1 Dun-Yen Kang,1 Jiun-Jen Chen,2,* and Li-Chiang Lin3,*
9
10
1
11 Department of Chemical Engineering, National Taiwan University
12
13 No. 1, Sec. 4, Roosevelt Rd., Taipei 10617, Taiwan
14
15
16 2
17 Green Energy and Environment Research Laboratories, Industrial Technology Research
18 Institute
19
20 No. 195, Sec. 4, Chung Hsing Rd., Chutung, Hsinchu 31040, Taiwan
21
22
23
24 3
25 William G. Lowrie Department of Chemical and Biomolecular Engineering, The Ohio State
26 University
27
28 151 W. Woodruff Avenue, Columbus, OH 43210, United States
29
30
31 * Corresponding authors: J.-J. Chen (JiunJenChen@itri.org.tw); L.-C. Lin
32
33 (lin.2645@osu.edu)
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 ACS Paragon Plus Environment
1
Langmuir Page 2 of 32

1
2
3 Abstract
4
5
6 Investigating metal-organic frameworks (MOFs) as water adsorbents has drawn increasing
7
8
attention for their potential in energy-related applications such as water production and heat
9
10
11 transformation. A specific MOF, MIL-100(Fe), is of particular interest for its large adsorption
12
13 capacity with the occurrence of water condensation at a relatively low partial pressure. In the
14
15 synthesis of MIL-100(Fe), depending on the reactants, structures with varying anion
16
17 terminals (e.g., F-, Cl-, or OH-) on the metal trimer have been reported. In this study, we
18
19 employed molecular simulations and density functional theory calculations for investigating
20
21
the water adsorption behaviors and the relative structural stability of MIL-100(Fe) with
22
23
24 different anions. We also proposed a possible defective structure and explored its water
25
26 adsorption properties. The results of this study are in good agreement with the experimental
27
28 measurements and are in support of the observations reported in the literature.
29
30 Understandings toward the spatial configurations and energetics of water molecules in these
31
32 materials have also shed light on their adsorption mechanism at the atomic level.
33
34
35
36
37
38
39 Keywords: Water vapor adsorption, Metal-organic frameworks, Structural stability, Monte
40
41 Carlo simulations, Density functional theory calculations
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
2
Page 3 of 32 Langmuir

1
2
3 Introduction
4
5
6 Metal-organic frameworks (MOFs) are an emerging class of crystalline
7
8
9
microporous/mesoporous materials, composed of metal clusters and organic ligands. MOFs
10
11 possess a variety of properties such as high porosity, large surface area, and diverse
12
13 functionality, thus providing opportunities in fields including, but not limited to, gas
14
15 separations,1-5 catalysis,6-9 drug delivery,10-14 gas storage,15-16 and adsorption refrigeration.17-
16
17 21
In recent years, using MOFs as water adsorbents has drawn increasing attention. MOFs
18
19 have been demonstrated for their promise in dehumidification22 and fresh water production.23
20
21
As examples, Seo et al. showed that, in terms of adsorption capacity and kinetics, MIL-
22
23
24 100(Fe) and MIL-101(Cr) can outperform conventional water sorbents such as SAPO-34,
25
26 NaX, and silica gel.22 Furukawa et al. demonstrated that MOF-801 possesses a high water
27
28 uptake at a low humidity of P/P0=0.1 and can be regenerated under mild conditions.24 Such
29
30 properties make MOF-801 a promising material for the production of fresh water from
31
32 ambient air. The large adsorption capacity at a low relative pressure of MOFs also makes
33
34
them potential adsorbents in heat pumps for more efficient heat transformation.20-21
35
36
37
38
39 Among MOFs reported to date as water adsorbents, MIL-100(Fe) is of particular interest.17,25-
40
26
41 The water adsorption capacity of as-synthesized MIL-100(Fe) can reach as much as 0.75 g-
42
43 water/g-MOF at 298 K,17 which transcends many other existing MOF materials (e.g., ZIF-8
44
45 of 0.01 g/g at 308 K and HKUST-1 of 0.60 g/g at 298 K).27 MIL-100(Fe) consists of iron(III)
46
47
and 1,3,5-bezenetricarboxylic (1,3,5-BTC). Its atomic structure can be seen in Figure 1,
48
49
50 which has two distinct types of cages with diameters of 2.5 nm and 2.9 nm.28-29 It has a
51
52 structure formula of FeIII3O(H2O)2X(C6H3(CO2)3)2•nH2O, where X can be F-, or OH-,30
53
54 depending on the synthesis conditions. The high water adsorption capacity of MIL-100(Fe)
55
56 may be attributed to its large surface area (BET surface area is as high as 2300 m2 g-1)31 and
57
58
59
60 ACS Paragon Plus Environment
3
Langmuir Page 4 of 32

1
2
3 open-metal sites that strongly coordinate with water molecules.31 In addition to the large
4
5 capacity, the capillary condensation of water in the framework (i.e., a sharp increase in the
6
7 adsorption isotherm due to the clustering of water molecules) occurs at a low relative
8
9 pressure P/P0 of 0.25-0.4, a range suitable for heat transformation applications.17,20
10
11
12
13
14 The first experimental synthesis of MIL-100(Fe) reported in the literature involves the use of
15
16 hydrofluoric acid (HF),29 a highly toxic chemical.30 Several modified synthesis approaches
17
18 have been therefore developed to avoid using hydrofluoric acid.9,30,32-34 Using the so-called
19
20 HF-free syntheses, no evident change in the structure of MIL-100(Fe) was found, except that
21
22 fluorine at the terminal position of the metal cluster (or so-called metal trimer) is replaced by
23
24
25
another anion.30 As noted above, hydroxide may replace the terminal fluoride in MIL-100(Fe)
26
27 obtained from an HF-free synthesis.35 Although MIL-100(Fe) structures with hydroxide have
28
29 been observed using HF-free conditions, we expect that MIL-100(Fe) with Cl- might be
30
31 possibly obtained as chloride ions are also commonly available in HF-free syntheses.
32
33 Accordingly, in this study, we considered MIL-100(Fe) structures with three different anions
34
35 (F-, Cl-, and OH-). For convenience, these structures were denoted hereafter as MIL-
36
37
100(Fe)_F, MIL-100(Fe)_Cl, and MIL-100(Fe)_OH, respectively.
38
39
40
41
42 Although no apparent structural change has been identified, the anions in MIL-100(Fe) play
43
44 an important role in its stability and crystallinity. The fluoride-functionalized terminal
45
46 appears to help stabilize the formation of trimeric inorganic building units in MIL-100(Fe),
47
48 which leads to more effective crystallization compared to other anions.25 The better
49
50
crystallinity of MIL-100(Fe)_F also results in a high cyclic stability for water adsorption.17,36
51
52
53 Compared to MIL-100(Fe)_F, MIL-100(Fe)_OH has been found to possess a smaller particle
54
55
56
57
58
59
60 ACS Paragon Plus Environment
4
Page 5 of 32 Langmuir

1
2
3 size and relatively poorly-defined crystal morphology.30 It has also been reported that this
4
5 structure suffers from a low structural stability in aqueous phase.37
6
7
8
9 Although MIL-100(Fe)_F has been widely explored both experimentally and computationally
10
11
12
for its potential as water adsorbents,17,38 the atomic-level understandings toward the influence
13
14 of alternative anions in MIL-100(Fe) on their water adsorption properties and structural
15
16 stability have yet been achieved. In this work, molecular simulations and density functional
17
18 theory (DFT) calculations were employed to systematically investigate MIL-100(Fe)_F,
19
20 MIL-100(Fe)_Cl, and MIL-100(Fe)_OH. Monte Carlo simulations in a grand canonical
21
22 ensemble (GCMC) or canonical ensemble (NVT) were carried out to simulate their water
23
24
25
adsorption isotherms as well as to understand the spatial distribution and energetics of water
26
27 molecules adsorbed in the frameworks. The ground state energy of each structure was also
28
29 computed using DFT to reveal the relative structural stability. Furthermore, we investigated
30
31 the water adsorption properties of a possible MIL-100(Fe) with defects.
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 ACS Paragon Plus Environment
5
Langmuir Page 6 of 32

1
2
3 Computational Details
4
5
6 Structures. The structures used in this work (see Figure 1) were adopted from the Cambridge
7
8
9
Crystallographic Data Centre (CCDC: 640536)29 with some modifications. First, we removed
10
11 residual water molecules inside the pores. Next, the unit cell was converted into a primitive
12
13 cell in order to reduce the computational cost. We then attached F-, Cl-, or OH- ions onto the
14
15 metal trimer, leading to the so-called MIL-100(Fe)_F, MIL-100(Fe)_Cl, or MIL-100(Fe)_OH
16
17 structures, respectively. Note that each trimer only has one anion for the overall charge
18
19 neutrality of the structures. These three structures (containing approximately 2800 atoms per
20
21
primitive unit cell) were subsequently relaxed by a cascade of the steepest descent, adjusted
22
23
24 basis set Newton-Raphson, and quasi-Newton methods. The structural relaxation was
25
26 performed using the Forcite module in the commercial Materials Studio® with the
27
28 DREIDING39 force field. The convergence criteria for energy, force, and displacement were
29
30 set to 2×10-5 kcal mol-1, 0.001 kcal mol-1 Å-1, and 0.001 Å, respectively. The atomic
31
32 coordinates of these three structures can be found in the Supporting Information.
33
34
35
36
37 Adsorption Isotherms. Monte Carlo simulations in a grand canonical ensemble (GCMC),
38
39 implemented in an open-source package RASPA 2.0,40 were carried out to compute the
40
41 adsorption isotherms of water in these three MIL-100(Fe) structures. The temperature of
42
43 adsorption was set to 300 K. Intermolecular potentials were described using both van der
44
45 Waals interactions (6-12 Lennard-Jones potential) and long-range Coulombic interactions.
46
47
The Lennard-Jones parameters of the framework atoms were adopted from the DREIDING39
48
49
50 force field, while their partial charges were assigned by the extended charge equilibrium
51
52 method (EQeq).41 Water molecules were modeled using the widely used 4-site TIP4P-Ew42
53
54 model. The Lorenz-Berthelot mixing rule was applied to determine the Lennard-Jones
55
56 parameters between two dissimilar atoms. The pair-wise Lennard-Jones potential was shifted
57
58
59
60 ACS Paragon Plus Environment
6
Page 7 of 32 Langmuir

1
2
3 and truncated at a cut-off radius of 12 Å without tail corrections, while the long-range
4
5 electrostatic interactions were calculated by the Ewald summation technique with a precision
6
7 of 1×10-6. A total of approximately 110 million Monte Carlo attempts including random
8
9 insertion, deletion, reinsertion, rotation, and translation of water molecules were carried out
10
11
12
in each of these calculations. The relative ratio of these moves was set respectively to 4: 4: 4:
13
14 1: 1. 100 million moves were conducted firstly to ensure that equilibrium had been achieved,
15
16 and the loading reached approximately a constant after a certain number of Monte Carlo
17
18 attempts during the system initialization. Subsequently, additional 10 million Monte Carlo
19
20 steps were conducted in each calculation to calculate the average loading.
21
22
23
24
25
Adsorption Distribution and Energetics. To investigate the spatial distribution of adsorbed
26
27 water molecules within the framework as well as the corresponding adsorption energetics at a
28
29 specific loading, Monte Carlo simulations in a canonical ensemble (NVT) at 300 K were
30
31 conducted. The detailed settings of the NVT simulations were similar to that of GCMC
32
33 calculations but in the absence of random insertion and deletion moves. NVT simulations
34
35 were also used to compute the binding geometry of a single water molecule in MIL-100(Fe).
36
37
In these calculations, we performed simulations at a low temperature of 5 K, ensuring the
38
39
40 adsorbed phase closes to the lowest energy state. Five replicas of the NVT calculations for
41
42 each structure were carried out with different initial configurations, and tens of millions
43
44 Monte Carlo moves were opted in each of the replica. The configuration with the lowest
45
46 energy was identified as the binding geometry.
47
48
49
50
Structural Stability. We used the ground state energies of MIL-100(Fe) structures with
51
52
53 varying anions to probe their relative structural stability. Considering the size of MIL-
54
55 100(Fe) structures (i.e., ~2800 atoms in the primitive unit cell), it is computationally
56
57
58
59
60 ACS Paragon Plus Environment
7
Langmuir Page 8 of 32

1
2
3 prohibited to compute the ground state energy of these structures using DFT. We therefore
4
5 constructed small clusters, comprising of one metal trimer and six ligands, as shown in Figure
6
7 2 to represent the corresponding full periodic structures. To maintain the neutrality of the
8
9 cluster and to reduce computational costs, two carboxyl groups in the 1,3,5-BTC linker were
10
11
12
replaced by hydrogen atoms. The ground state energy of these clusters was calculated using
13
14 DFT implemented in the CASTEP43 module of Materials Studio®. Plane-wave periodic DFT
15
16 that involves few convergence parameters (e.g., cutoff energy and number of k-points) was
17
18 adopted, and many pioneering efforts have been made for the program optimization.44-45 With
19
20 a reasonable vacuum space, periodic DFT could be an effective way to study non-periodic
21
22 cluster systems.46 In these calculations, the generalized gradient approximation (GGA) in the
23
24
25
scheme of Perdew–Burke–Ernzerhof (PBE)47 was opted as the exchange-correlation
26
27 functional without applying additional dispersion corrections. A plane wave energy cutoff of
28
29 340.0 eV and the Monkhorst–Pack k-point 1×1×1 mesh48 were used. At least 7 Å of vacuum
30
31 space was included to minimize the interaction between periodic images. The convergence
32
33 criteria for energy, force, and displacement were 1×10-5 eV/atom, 0.03 eV Å-1, and 0.001 Å,
34
35 respectively. In this study, in order to make a meaningful comparison of the computed ground
36
37
38
state energies of these clusters that possess different chemical compositions, we further
39
40 assumed the following two reactions:
41
42
43
44 MIL-100(Fe)_F + HCl ⇌ MIL-100(Fe)_Cl + HF Eq. (1)
45
46
47
MIL-100(Fe)_F + H2O ⇌ MIL-100(Fe)_OH + HF Eq. (2)
48
49
50
51 From Eq. (1), we defined the relative energy difference between MIL-100(Fe)_F and MIL-
52
53 100(Fe)_Cl to be represented by the difference between E(MIL-100(Fe)_F) – E(HF) and
54
55 E(MIL-100(Fe)_Cl) – E(HCl). Similarly, from Eq. (2), the energy difference between MIL-
56
57
58
59
60 ACS Paragon Plus Environment
8
Page 9 of 32 Langmuir

1
2
3 100(Fe)_F and MIL-100(Fe)_OH can be quantified by the difference between E(MIL-
4
5 100(Fe)_F) – E(HF) and E(MIL-100(Fe)_OH) – E(H2O). E(HF), E(HCl), and E(H2O) are
6
7 respectively the ground state energies of gas phase HF, HCl, and H2O molecules. These
8
9 energies were computed using exactly the same level of theory and the same simulation
10
11
12
dimension as that adopted in the cluster calculations. All of the above calculations were
13
14 carried out in the gas phase instead of in a solvent-mediated environment such as under the
15
16 synthesis condition and/or the condition when the material is filled (or partially filled) with
17
18 water molecules. Nonetheless, given all of the studied structures are very similar, the gas
19
20 phase assumption should serve as a decent first-order approximation. Further, as discussed
21
22 below, our calculation results are in good agreement with those reported experimentally.
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 ACS Paragon Plus Environment
9
Langmuir Page 10 of 32

1
2
3 Results and Discussion
4
5
6 Water Adsorption. Water adsorption isotherms in MIL-100(Fe)_F, MIL-100(Fe)_Cl, and
7
8
9
MIL-100(Fe)_OH at 300 K computed by GCMC simulations are summarized in Figure 3.
10
11 Isotherms are presented as a function of the relative pressure (P/P0) with P0 to be 3750 Pa at a
12
13 room temperature according to TIP4P-Ew water model.49 The calculated water isotherm of
14
15 MIL-100(Fe)_F is in agreement with the experimental ones.17,25 Compared to a previous
16
17 computational study,38 our isotherm resembles the experimental data more accurately, which
18
19 may be attributed to the adopted water models. In this work, the 4-site TIP4P-Ew model was
20
21
adopted, whereas the 3-site SPC/E water model was opted in the previous work. It has been
22
23
24 shown in our previous study that the TIP4P-Ew model may better describe water adsorption
25
26 properties,50 and it has also been widely adopted in the calculations of adsorption in
27
28 nanoporous materials.51-53 Additionally, Figure 3 shows that the isotherm of MIL-100(Fe)_F
29
30 possesses a plateau at an uptake of approximately 0.4 g-water/g-MOF (i.e., corresponding to
31
32 P/P0 of 0.40-0.53 in the simulated isotherm), and this interesting phenomenon can be found
33
34
in both experimental and simulation isotherms. It has been hypothesized that this may be
35
36
37 caused by the difference in water adsorption characteristics between two distinct cages of
38
39 MIL-100(Fe).17,25 As illustrated in Figure 1, two sizes of cages (2.5 and 2.9 nm) are present in
40
41 MIL-100(Fe). At a low pressure, water molecules may preferentially adsorb in the small
42
43 cages (2.5 nm) due to the more favorable interactions between water and the framework until
44
45 saturation. Thereafter the adsorption in the large cages (2.9 nm) occurs, thus leading to the
46
47
observed plateau. To validate this hypothesis, snapshots of the water configurations adsorbed
48
49
50 in MIL-100(Fe)_F at various pressures are taken from our simulations and shown in Figure 4.
51
52 For a clear presentation, the MIL-100(Fe) framework is shown as sticks with green and
53
54 orange colors (represent small and large cage structures, respectively), while the space-filling
55
56 style was used for water molecules. These atomistic configurations clearly indicate that at a
57
58
59
60 ACS Paragon Plus Environment
10
Page 11 of 32 Langmuir

1
2
3 low pressure (P/P0 < 0.27), water molecules indeed first occupied the small cages. The
4
5 condensation of water in the small cages occurred at a P/P0 between 0.27 and 0.40. At a P/P0
6
7 of 0.40, Figure 4 shows that the small cages became completely filled with a density of water
8
9 determined to be around 0.9 g/cm3, whereas nearly no water molecule was adsorbed in the
10
11
12
large cages. This preferential adsorption in different types of cages observed in our
13
14 simulations also agrees with previous findings in related structures.23,25-26,54
15
16
17
18 The water adsorption capability in MIL-100(Fe) with different anions (Figure 3)
19
20 demonstrates that the water uptake in both MIL-100(Fe)_Cl and MIL-100(Fe)_OH at 300 K
21
22 is substantially lower than that of MIL-100(Fe)_F although these three structures possess a
23
24
25
nearly identical saturation uptake due to their similar pore structures. Specifically, the sharp
26
27 increase in the adsorption isotherm of both MIL-100(Fe)_Cl and MIL-100(Fe)_OH occurs at
28
29 a notably higher pressure, thus suggesting a weaker water-framework interaction was
30
31 involved in these two structures. As a consequence, different from MIL-100(Fe)_F, the
32
33 adsorption isotherms of both MIL-100(Fe)_Cl and MIL-100(Fe)_OH do not possess a plateau.
34
35 Because of their relatively weaker water-framework interactions, the difference in the
36
37
adsorption characteristics between two types of cages becomes marginal. Additionally, this
38
39
40 also indicates that both MIL-100(Fe)_Cl and MIL-100(Fe)_OH may be relatively less
41
42 promising adsorbent candidates for applications such as water harvesting and heat
43
44 transformation.
45
46
47
48 Adsorption Energetics and Binding Geometries. To gain insights into the adsorption
49
50
energetics of water in MIL-100(Fe) with different terminal anions, the guest-host and guest-
51
52
53 guest (i.e., guest: water molecules; host: the framework) interaction energies were calculated
54
55 using NVT simulations (see Figure 5). As discussed above, MIL-100(Fe)_F was indeed
56
57
58
59
60 ACS Paragon Plus Environment
11
Langmuir Page 12 of 32

1
2
3 identified to possess the strongest host-guest interaction among the three MOF structures
4
5 (Figure 5a) in the low-loading region (below 0.04 g-water/g-MOF). Compared to MIL-
6
7 100(Fe)_Cl and MIL-100(Fe)_OH, the host-guest interaction energy for MIL-100(Fe)_F is
8
9 lower by as much as 10 kJ/mol. The strong host-guest interaction in MIL-100(Fe)_F explains
10
11
12
its significantly higher water uptake in the low-pressure region. This is also reflected on the
13
14 absence of the sharp uptake increase in both MIL-100(Fe)_Cl and MIL-100(Fe)_OH at a low
15
16 relative pressure, as the guest-host interaction has been known to facilitate the occurrence of
17
18 water condensation in confined environments. At a higher loading (> 0.2 g-water/g-MOF),
19
20 the guest-host interaction energies among the three MOF structures become comparable.
21
22 Under this condition, most water molecules are anticipated to be at a distance from the
23
24
25
framework surface, thus making the overall difference in host-guest interaction become
26
27 relatively insignificant. For this, we have computed the accumulative number of water
28
29 adsorbed in the two distinct cage types of the MIL-100(Fe)_F structure as a function of the
30
31 distance from the center of the cages (see the Supporting Information Figures S1 and S2 for
32
33 the adsorption in the large and small cages, respectively). For water adsorption in the small
34
35 cages as an example, at a low loading (0.043 g/g) water molecules are preferentially adsorbed
36
37
38 at a distance of 14-17 Å away from the center of the cages (i.e., near the cage surface), as
39
40 suggested in Figure S1. As noted above, this confirms that the host-guest interaction should
41
42 play a critical role in adsorptive behaviors in this low-pressure region. By contrast, at a high
43
44 loading (0.945 g/g), our results show that most of the water are located near the center (within
45
46
14 Å of the center of the cage). Specifically, approximately 80% of the total water molecules
47
48
49 are adsorbed away from the cage surface and therefore, guest-guest interactions should
50
51 dominate the total energies. Interestingly, our analysis shown in Figures S1 and S2 also
52
53 indicates that the density of adsorbed water in both cages can be notably different. It was
54
55 found that the adsorbed amount of water molecules in the small cages is greater than that in
56
57
58
59
60 ACS Paragon Plus Environment
12
Page 13 of 32 Langmuir

1
2
3 the large cages, leading to a denser adsorbed phase in the small cages. In agreement with a
4
5 recent study by Choi et al., the density of adsorbed phase may differ drastically, which
6
7 depends on the pore structure and surface chemistry.55 Furthermore, for the guest-guest
8
9 interactions shown in Figure 5b, it was found that water molecules in MIL-100(Fe)_F have
10
11
12
the least favorable interactions (i.e., highest energy values). This may be attributed to the
13
14 weaker guest-host interaction found in MIL-100(Fe)_Cl and MIL-100(Fe)_OH, thus allowing
15
16 the agglomeration of water molecules with a more favorable configuration. In other words,
17
18 the less constrain imposed by the host-guest interactions may lead to more favorable guest-
19
20 guest contributions.
21
22
23
24
25
Figure 5a shows that there is an interesting local maximum energy in the guest-host
26
27 interaction at the loading of approximately 0.1 g-water/g-MOF for all of the three structures,
28
29 and a corresponding local minimum can be seen in the guest-guest interaction (Figure 5b).
30
31 This behavior indicates a possible configuration change for water molecules adsorbed in the
32
33 framework. At a low loading condition (less than 0.05 g/g, water molecules are dispersedly
34
35 adsorbed in the framework), these water molecules are located in a way to maximize its
36
37
interaction with the framework. When the concentration of adsorbed water increases, a
38
39
40 change in the configuration for those initially adsorbed water molecules may be needed in
41
42 order to gain strong water-water interactions and subsequently lower the overall energy,
43
44
45
46 The binding geometry and distance of water in MIL-100(Fe)_F, MIL-100(Fe)_Cl, and MIL-
47
48 100(Fe)_OH determined by NVT calculations at a low temperature are summarized in Figure
49
50
6. In all of the three MIL-100(Fe) structures, the water binding configuration was found to be
51
52
53 located between an open-metal site and an anion terminal. In particular, one of the water
54
55 hydrogen atoms was in a close proximity to the anion terminal while the oxygen atom was
56
57
58
59
60 ACS Paragon Plus Environment
13
Langmuir Page 14 of 32

1
2
3 located near the open-metal site. These results appear reasonable as, from the classical force
4
5 field point of view, the negatively charged anion can help stabilize the positively charged
6
7 hydrogen atoms of water. Similarly, the same effect exists between the negatively charged
8
9 oxygen atom of water and the positively charged open-metal site. The hydrogen (H2O)-anion
10
11
12 distance is shorter in MIL-100(Fe)_F (1.87 Å) relative to that in MIL-100(Fe)_OH (2.19 Å)
13
14 and MIL-100(Fe)_Cl (2.28 Å), demonstrating that F- possesses the strongest interaction with
15
16 water. The oxygen (H2O)-metal site distance was found to be correspondingly larger in MIL-
17
18 100(Fe)_F (3.49 Å) relative to that in MIL-100(Fe)_OH (3.22 Å) and MIL-100(Fe)_Cl (3.17
19
20
21 Å).
22
23
24
25
26 Structural Stability and Defective MIL-100(Fe). To understand the structural stability of
27
28 MIL-100(Fe)_F, MIL-100(Fe)_Cl, and MIL-100(Fe)_OH, the ground state energy of their
29
30 representative clusters as introduced previously were calculated and compared. The relative
31
32 energy can serve as a proxy to indicate the stability of the basic building unit in MIL-100(Fe).
33
34 As shown in Figure 7, MIL-100(Fe)_F was found to be more stable than MIL-100(Fe)_OH
35
36
37
by approximately 0.8 eV. This value is deemed as significant, which is far beyond the
38
39 thermal energy of 0.025 eV at the room temperature. This finding supports the observation
40
41 reported in the literature that MIL-100(Fe) synthesized in the absence of HF (i.e., structures
42
43 with hydroxide) tends to be notably less stable.30,37 Interestingly, MIL-100(Fe)_Cl was
44
45 identified to be potentially more stable than MIL-100(Fe)_F. However, we should note that
46
47 our energy calculations do not consider the presence of adsorbed molecules. For instance,
48
49
upon water vapor adsorption, the stronger interactions between H2O and metal trimers with F-
50
51
52 relative to that between H2O and trimers with other anions may also influence the structural
53
54 stability. Additionally, although MIL-100(Fe)_Cl is predicted to be more thermodynamically
55
56 stable than MIL-100(Fe)_OH, the fact that MIL-100(Fe)_Cl structures were not observed
57
58
59
60 ACS Paragon Plus Environment
14
Page 15 of 32 Langmuir

1
2
3 experimentally in HF-free conditions suggests that solvent molecules involved in the
4
5 synthesis may also play a role. Nonetheless, our calculations have, for the first time, shed
6
7 light on the relative structural stability of MIL-100(Fe) with varying anions.
8
9
10
11 As noted above, based on our DFT calculations, the metal trimer cluster of MIL-100(Fe)_OH
12
13 was identified to be notably less stable compared to that of MIL-100(Fe)_F and MIL-
14
15 100(Fe)_Cl. It is therefore expected that MIL-100(Fe)_OH may be potentially more prone to
16
17
structural defects. In this study, considering its relatively unstable trimer clusters, we
18
19
20 proposed a defective MIL-100(Fe)_OH structure that has metal-node missing defects (see
21
22 Figure 8 for illustration) with a density of one missing metal node out of 68 nodes in the
23
24 primitive cell of MOF-100(Fe). The atomic coordinates of the defect structure can be found
25
26 in the Supporting Information. Similar types of defects have been also studied in another
27
28 well-studied MOF, UiO66.56-58 Same as the analysis carried out for defect-free frameworks,
29
30
we have also computed the adsorption isotherms using GCMC simulations in this metal-node
31
32
33 missing structure. Figure 9 shows that the uptake in the defective MIL-100(Fe)_OH structure
34
35 was predicted to be lower than that of the corresponding defect-free one. Specifically, the
36
37 uptake of water in the defective MIL-100(Fe)_OH structure is nearly zero even at a relative
38
39 pressure of approximately 1.0. In the absence of the metal nodes, the interactions between
40
41 adsorbed water molecules and the framework are anticipated to become overall weaker. As
42
43
shown in Supporting Information Figure S3, the host-guest interaction is weaker in the MIL-
44
45
46 100(Fe)_OH structure with metal-node missing defects, in particular in the low-pressure
47
48 region. Given the host-guest interactions at a low loading can facilitate the condensation of
49
50 water adsorbed in the framework, such defective structure accordingly shows a lower
51
52 adsorption ability. Our calculations also infer that MIL-100(Fe)_OH may have a relatively
53
54 poor cyclic stability with a reduction in adsorption ability, which is in good agreement with
55
56
previously reported experimental results that MIL-100(Fe) synthesized in absence of HF
57
58
59
60 ACS Paragon Plus Environment
15
Langmuir Page 16 of 32

1
2
3 suffered from a significant reduction in adsorption ability after aqueous water treatment.37
4
5 We should note that it remains unknown whether the proposed defective structure is stable in
6
7 the presence of water. The defective structure is possibly an intermediate structure before the
8
9 complete degradation of the materials. A detailed investigation of the structural degradation
10
11
12
can be an important subject of future studies.
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 ACS Paragon Plus Environment
16
Page 17 of 32 Langmuir

1
2
3 Conclusion
4
5
6 We employed Monte Carlo techniques (GCMC and NVT simulations) and density functional
7
8
9
theory (DFT) calculations to investigate the water adsorption behaviors and relative structural
10
11 stability of MIL-100(Fe) with different terminal anions (F-, Cl-, or OH-). Our calculated
12
13 adsorption isotherm in MIL-100(Fe)_F is in agreement with that measured experimentally,
14
15 and both simulation and experiment isotherms show an obvious plateau. By visualizing the
16
17 atomic-level adsorption configurations at varying pressures, it was found that the plateau can
18
19 be attributed to the difference in adsorption characteristics of two distinct types of cages in
20
21
MIL-100(Fe). Small cages, which provide a stronger interaction to water molecules
22
23
24 compared to large cages, were found to be completely filled with water at a notably lower
25
26 pressure. Amongst these three structures, due to the strongest interaction between the
27
28 terminal F- anion and water molecules, the adsorption isotherm in MIL-100(Fe)_F has the
29
30 largest adsorption capability with the occurrence of water condensation at a much smaller
31
32 relative pressure. From the ground state energy calculations by DFT, MIL-100(Fe)_F was
33
34
identified to be significantly more stable than MIL-100(Fe)_OH. This result agrees with the
35
36
37 experimental observations where MIL-100(Fe)_F was demonstrated to have higher
38
39 crystallinity with better defined morphology. Considering the relatively poor stability of
40
41 MIL-100(Fe)_OH, a possible defective MIL-100(Fe)_OH structure (i.e., metal-node missing
42
43 defects) was proposed and studied. This proposed defective structure appears to have a lower
44
45 water uptake compared to the defect-free one, inferring that these less stable materials are
46
47
prone to a poor cyclic stability in adsorption. Overall, the outcomes of this study are in
48
49
50 support of the experimental findings reported in the literature, while the atomistic
51
52 understandings achieved have shed light on the water adsorption mechanism in MIL-100(Fe)
53
54 with different anions.
55
56
57
58
59
60 ACS Paragon Plus Environment
17
Langmuir Page 18 of 32

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 Figure 1. Illustration of the cage structure (left) and the building unit (right) of MIL-100(Fe).
22 Three different anions are considered in this study: F-, Cl-, or OH-.
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 ACS Paragon Plus Environment
18
Page 19 of 32 Langmuir

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24 Figure 2. Illustration of the cluster model used for the ground state energy calculations.
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 ACS Paragon Plus Environment
19
Langmuir Page 20 of 32

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
Figure 3. GCMC-calculated and experiment-measured17,25 water adsorption isotherms at 300
31
32
K. The adsorption isotherm for the anhydrous framework of MIL-100(Fe)_F from a
33 computational study is also included for comparison.38
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 ACS Paragon Plus Environment
20
Page 21 of 32 Langmuir

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 Figure 4. (a) Illustration of two distinct types of cages in MIL-100(Fe). (b) Snapshots of
50 water configurations adsorbed in MIL-100(Fe)_F at various relative pressures, P/P0, ranging
51 from 0.27 to 0.8.
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
21
Langmuir Page 22 of 32

1
2
3
4 (a)
5 0
6

Host-Guest Energy
7
-5
8
9

(kJ/mol)
10 -10
11 Original_F
12
13 -15 Original_Cl
14 Original_OH
15
-20
16
17
18 0.0 0.2 0.4 0.6 0.8 1.0
19 Loading (g/g)
20 (b)
21
-25
22
Guest-Guest Energy

23 Original_F
24 -30 Original_Cl
25
(kJ/mol)

Original_OH
26
27 -35
28
29
-40
30
31
32 -45
33 0.0 0.2 0.4 0.6 0.8 1.0
34
35 Loading (g/g)
(c)
36
37 -36
Total Energy (kJ/mol)

38 Original_F
39 Original_Cl
40
-40
41
Original_OH
42 -44
43
44
45 -48
46
47
48 -52
49 0.0 0.2 0.4 0.6 0.8 1.0
50 Loading (g/g)
51
52
53 Figure 5. (a) Host(framework)-guest(water), (b) guest-guest, and (c) total interaction
54 energies as a function of the water uptake in MIL-100(Fe)_F, MIL-100(Fe)_Cl, and MIL-
55 100(Fe)_OH.
56
57
58
59
60 ACS Paragon Plus Environment
22
Page 23 of 32 Langmuir

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
Figure 6. The binding configurations of water molecules and the corresponding distances of
16
17
H(H2O)-anions(framework) and O(H2O)-Fe(framework) in (a) MIL-100(Fe)_F, (b) MIL-
18 100(Fe)_Cl, and (c) MIL-100(Fe)_OH.
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 ACS Paragon Plus Environment
23
Langmuir Page 24 of 32

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 Figure 7. The relative energies of MIL-100(Fe) clusters with different terminal anions (i.e.,
22 F-, Cl-, and OH-).
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 ACS Paragon Plus Environment
24
Page 25 of 32 Langmuir

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 Figure 8. Illustration of (a) defect-free MIL-100(Fe) and (b) metal-node missing MIL-100(Fe)
27 structures. Note that hydrogen atoms are not displayed in this figure.
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 ACS Paragon Plus Environment
25
Langmuir Page 26 of 32

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 Figure 9. The GCMC-calculated water adsorption isotherm of the metal-node missing MIL-
30 100(Fe)_OH structure.
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 ACS Paragon Plus Environment
26
Page 27 of 32 Langmuir

1
2
3 Associated Content
4
5 Supporting Information Available. The supporting information includes additional figures
6
7 referred in the main article (i.e., accumulative number of water adsorbed in MIL-100(Fe)_F
8
9 and the host-guest interaction energy of water in the defective MIL-100(Fe)_OH structure) as
10
11 well as the structural information of all structures studied in this work. This material is
12
13
available free of charge via the Internet at http://pubs.acs.org.
14
15
16
17
AUTHOR INFORMATION
18
19 Author Contributions
20
21
22 This study was developed and completed through contributions by all authors.
23
24
25
26
27
28 Acknowledgement
29
30
31 This work was supported by the Bureau of Energy of the Ministry of Economic Affairs,
32
33 Taiwan. The authors also gratefully thank Ohio Supercomputer Center (OSC) for
34
35 computational resources.59
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 ACS Paragon Plus Environment
27
Langmuir Page 28 of 32

1
2
3 References
4
5
(1) Li, J. R.; Kuppler, R. J.; Zhou, H. C., Selective gas adsorption and separation in
6
metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1477-1504.
7
8 (2) Hamon, L.; Heymans, N.; Llewellyn, P. L.; Guillerm, V.; Ghoufi, A.; Vaesen, S.;
9 Maurin, G.; Serre, C.; De Weireld, G.; Pirngruber, G. D., Separation of CO2-CH4
10 mixtures in the mesoporous MIL-100(Cr) MOF: experimental and modelling
11 approaches. Dalton Trans. 2012, 41, 4052-4059.
12 (3) Parkes, M. V.; Sava Gallis, D. F.; Greathouse, J. A.; Nenoff, T. M., Effect of Metal in
13 M-3(btc)(2) and M-2(dobdc) MOFs for O-2/N-2 Separations: A Combined Density
14 Functional Theory and Experimental Study. J. Phys. Chem. C 2015, 119, 6556-6567.
15 (4) Motevalli, B.; Wang, H.; Liu, J. Z., Cooperative Reformable Channel System with
16 Unique Recognition of Gas Molecules in a Zeolitic Imidazolate Framework with
17 Multilevel Flexible Ligands. J. Phys. Chem. C 2015, 119, 16762-16768.
18 (5) Li, W. B.; Zhang, Y. F.; Zhang, C. Y.; Meng, Q.; Xu, Z. H.; Su, P. C.; Li, Q. B.;
19
Shen, C.; Fan, Z.; Qin, L.; Zhang, G. L., Transformation of metal-organic frameworks
20
21
for molecular sieving membranes. Nat. Commun. 2016, 7, 11315.
22 (6) Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T., Metal-
23 organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38, 1450-1459.
24 (7) Laurier, K. G. M.; Vermoortele, F.; Ameloot, R.; De Vos, D. E.; Hofkens, J.;
25 Roeffaers, M. B. J., Iron(III)-Based Metal-Organic Frameworks As Visible Light
26 Photocatalysts. J. Am. Chem. Soc. 2013, 135, 14488-14491.
27 (8) Santos, V. P.; Wezendonk, T. A.; Jaen, J. J. D.; Dugulan, A. I.; Nasalevich, M. A.;
28 Islam, H. U.; Chojecki, A.; Sartipi, S.; Sun, X.; Hakeem, A. A.; Koeken, A. C. J.;
29 Ruitenbeek, M.; Davidian, T.; Meima, G. R.; Sankar, G.; Kapteijn, F.; Makkee, M.;
30 Gascon, J., Metal organic framework-mediated synthesis of highly active and stable
31 Fischer-Tropsch catalysts. Nat. Commun. 2015, 6, 6451.
32
(9) Zhang, F.; Shi, J.; Jin, Y.; Fu, Y.; Zhong, Y.; Zhu, W., Facile synthesis of MIL-
33
34
100(Fe) under HF-free conditions and its application in the acetalization of aldehydes
35 with diols. Chem. Eng. J. 2015, 259, 183-190.
36 (10) Horcajada, P.; Chalati, T.; Serre, C.; Gillet, B.; Sebrie, C.; Baati, T.; Eubank, J. F.;
37 Heurtaux, D.; Clayette, P.; Kreuz, C.; Chang, J. S.; Hwang, Y. K.; Marsaud, V.;
38 Bories, P. N.; Cynober, L.; Gil, S.; Ferey, G.; Couvreur, P.; Gref, R., Porous metal-
39 organic-framework nanoscale carriers as a potential platform for drug delivery and
40 imaging. Nat. Mater. 2010, 9, 172-178.
41 (11) Huxford, R. C.; Della Rocca, J.; Lin, W. B., Metal-organic frameworks as potential
42 drug carriers. Curr. Opin. Chem. Biol. 2010, 14, 262-268.
43 (12) Cunha, D.; Ben Yahia, M.; Hall, S.; Miller, S. R.; Chevreau, H.; Elkaim, E.; Maurin,
44 G.; Horcajada, P.; Serre, C., Rationale of Drug Encapsulation and Release from
45
Biocompatible Porous Metal-Organic Frameworks. Chem. Mater. 2013, 25, 2767-
46
2776.
47
48 (13) Bernini, M. C.; Fairen-Jimenez, D.; Pasinetti, M.; Ramirez-Pastor, A. J.; Snurr, R. Q.,
49 Screening of bio-compatible metal-organic frameworks as potential drug carriers
50 using Monte Carlo simulations. J. Mater. Chem. B 2014, 2, 766-774.
51 (14) Bellido, E.; Guillevic, M.; Hidalgo, T.; Santander-Ortega, M. J.; Serre, C.; Horcajada,
52 P., Understanding the colloidal stability of the mesoporous MIL-100(Fe)
53 nanoparticles in physiological media. Langmuir 2014, 30, 5911-5920.
54 (15) Li, Y.; Yang, R. T., Gas adsorption and storage in metal-organic framework MOF-
55 177. Langmuir 2007, 23, 12937-12944.
56
57
58
59
60 ACS Paragon Plus Environment
28
Page 29 of 32 Langmuir

1
2
3 (16) Li, L. J.; Tang, S. F.; Wang, C.; Lv, X. X.; Jiang, M.; Wu, H. Z.; Zhao, X. B., High
4 gas storage capacities and stepwise adsorption in a UiO type metal-organic
5 framework incorporating Lewis basic bipyridyl sites. Chem. Commun. 2014, 50,
6 2304-2307.
7 (17) Jeremias, F.; Khutia, A.; Henninger, S. K.; Janiak, C., MIL-100(Al, Fe) as water
8 adsorbents for heat transformation purposes-a promising application. J. Mater. Chem.
9 2012, 22, 10148-10151.
10
(18) Ehrenmann, J.; Henninger, S. K.; Janiak, C., Water Adsorption Characteristics of
11
12
MIL-101 for Heat-Transformation Applications of MOFs. Eur. J. Inorg. Chem. 2011,
13 2011, 471-474.
14 (19) Henninger, S. K.; Habib, H. A.; Janiak, C., MOFs as Adsorbents for Low
15 Temperature Heating and Cooling Applications. J. Am. Chem. Soc. 2009, 131, 2776-
16 2777.
17 (20) de Lange, M. F.; Verouden, K. J. F. M.; Vlugt, T. J. H.; Gascon, J.; Kapteijn, F.,
18 Adsorption-Driven Heat Pumps: The Potential of Metal–Organic Frameworks. Chem.
19 Rev. 2015, 115, 12205-12250.
20 (21) de Lange, M. F.; van Velzen, B. L.; Ottevanger, C. P.; Verouden, K. J. F. M.; Lin, L.-
21 C.; Vlugt, T. J. H.; Gascon, J.; Kapteijn, F., Metal–Organic Frameworks in
22 Adsorption-Driven Heat Pumps: The Potential of Alcohols as Working Fluids.
23
Langmuir 2015, 31, 12783-12796.
24
25
(22) Seo, Y. K.; Yoon, J. W.; Lee, J. S.; Hwang, Y. K.; Jun, C. H.; Chang, J. S.; Wuttke,
26 S.; Bazin, P.; Vimont, A.; Daturi, M.; Bourrelly, S.; Llewellyn, P. L.; Horcajada, P.;
27 Serre, C.; Ferey, G., Energy-Efficient Dehumidification over Hierachically Porous
28 Metal-Organic Frameworks as Advanced Water Adsorbents. Adv. Mater. 2012, 24,
29 806-810.
30 (23) Kim, S. I.; Yoon, T. U.; Kim, M. B.; Lee, S. J.; Hwang, Y. K.; Chang, J. S.; Kim, H.
31 J.; Lee, H. N.; Lee, U. H.; Bae, Y. S., Metal-organic frameworks with high working
32 capacities and cyclic hydrothermal stabilities for fresh water production. Chem. Eng.
33 J. 2016, 286, 467-475.
34 (24) Furukawa, H.; Gándara, F.; Zhang, Y.-B.; Jiang, J.; Queen, W. L.; Hudson, M. R.;
35 Yaghi, O. M., Water Adsorption in Porous Metal–Organic Frameworks and Related
36
Materials. J. Am. Chem. Soc. 2014, 136, 4369-4381.
37
(25) Küsgens, P.; Rose, M.; Senkovska, I.; Fröde, H.; Henschel, A.; Siegle, S.; Kaskel, S.,
38
39 Characterization of metal-organic frameworks by water adsorption. Microporous
40 Mesoporous Mater. 2009, 120, 325-330.
41 (26) de Lange, M. F.; Gutierrez-Sevillano, J. J.; Hamad, S.; Vlugt, T. J. H.; Calero, S.;
42 Gascon, J.; Kapteijn, F., Understanding Adsorption of Highly Polar Vapors on
43 Mesoporous MIL-100(Cr) and MIL-101 (Cr): Experiments and Molecular
44 Simulations. J. Phys. Chem. C 2013, 117, 7613-7622.
45 (27) Zhang, K.; Lively, R. P.; Dose, M. E.; Brown, A. J.; Zhang, C.; Chung, J.; Nair, S.;
46 Koros, W. J.; Chance, R. R., Alcohol and water adsorption in zeolitic imidazolate
47 frameworks. Chem. Commun. 2013, 49, 3245-3247.
48 (28) Vimont, A.; Goupil, J.-M.; Lavalley, J.-C.; Daturi, M.; Surblé, S.; Serre, C.; Millange,
49 F.; Férey, G.; Audebrand, N., Investigation of Acid Sites in a Zeotypic Giant Pores
50
Chromium(III) Carboxylate. J. Am. Chem. Soc. 2006, 128, 3218-3227.
51
52
(29) Horcajada, P.; Surble, S.; Serre, C.; Hong, D. Y.; Seo, Y. K.; Chang, J. S.; Greneche,
53 J. M.; Margiolaki, I.; Ferey, G., Synthesis and catalytic properties of MIL-100(Fe), an
54 iron(III) carboxylate with large pores. Chem. Commun. 2007, 2820-2822.
55 (30) Seo, Y. K.; Yoon, J. W.; Lee, J. S.; Lee, U. H.; Hwang, Y. K.; Jun, C. H.; Horcajada,
56 P.; Serre, C.; Chang, J. S., Large scale fluorine-free synthesis of hierarchically porous
57
58
59
60 ACS Paragon Plus Environment
29
Langmuir Page 30 of 32

1
2
3 iron(III) trimesate MIL-100(Fe) with a zeolite MTN topology. Microporous
4 Mesoporous Mater. 2012, 157, 137-145.
5 (31) Yoon, J. W.; Seo, Y. K.; Hwang, Y. K.; Chang, J. S.; Leclerc, H.; Wuttke, S.; Bazin,
6 P.; Vimont, A.; Daturi, M.; Bloch, E.; Llewellyn, P. L.; Serre, C.; Horcajada, P.;
7 Greneche, J. M.; Rodrigues, A. E.; Ferey, G., Controlled reducibility of a metal-
8 organic framework with coordinatively unsaturated sites for preferential gas sorption.
9 Angew. Chem., Int. Ed. Engl. 2010, 49, 5949-5952.
10
(32) Marquez, A. G.; Demessence, A.; Platero-Prats, A. E.; Heurtaux, D.; Horcajada, P.;
11
12
Serre, C.; Chang, J. S.; Ferey, G.; de la Pena-O'Shea, V. A.; Boissiere, C.; Grosso, D.;
13 Sanchez, C., Green Microwave Synthesis of MIL-100(Al, Cr, Fe) Nanoparticles for
14 Thin-Film Elaboration. Eur. J. Inorg. Chem. 2012, 5165-5174.
15 (33) Duan, S. X.; Li, J. X.; Liu, X.; Wang, Y. N.; Zeng, S. Y.; Shao, D. D.; Hayat, T., HF-
16 Free Synthesis of Nanoscale Metal-Organic Framework NMIL-100(Fe) as an
17 Efficient Dye Adsorbent. ACS Sustainable Chem. Eng. 2016, 4, 3368-3378.
18 (34) Jeremias, F.; Henninger, S. K.; Janiak, C., Ambient pressure synthesis of MIL-
19 100(Fe) MOF from homogeneous solution using a redox pathway. Dalton Trans.
20 2016, 45, 8637-8644.
21 (35) Canioni, R.; Roch-Marchal, C.; Secheresse, F.; Horcajada, P.; Serre, C.; Hardi-Dan,
22 M.; Ferey, G.; Greneche, J. M.; Lefebvre, F.; Chang, J. S.; Hwang, Y. K.; Lebedev,
23
O.; Turner, S.; Van Tendeloo, G., Stable polyoxometalate insertion within the
24
25
mesoporous metal organic framework MIL-100(Fe). J. Mater. Chem. 2011, 21, 1226-
26 1233.
27 (36) Soubeyrand-Lenoir, E.; Vagner, C.; Yoon, J. W.; Bazin, P.; Ragon, F.; Hwang, Y. K.;
28 Serre, C.; Chang, J.-S.; Llewellyn, P. L., How Water Fosters a Remarkable 5-Fold
29 Increase in Low-Pressure CO2 Uptake within Mesoporous MIL-100(Fe). J. Am.
30 Chem. Soc. 2012, 134, 10174-10181.
31 (37) Bezverkhyy, I.; Weber, G.; Bellat, J. P., Degradation of fluoride-free MIL-100(Fe)
32 and MIL-53(Fe) in water: Effect of temperature and pH. Microporous Mesoporous
33 Mater. 2016, 219, 117-124.
34 (38) Kolokathis, P. D.; Pantatosaki, E.; Papadopoulos, G. K., Atomistic Modeling of Water
35 Thermodynamics and Kinetics within MIL-100(Fe). J. Phys. Chem. C 2015, 119,
36
20074-20084.
37
(39) Mayo, S. L.; Olafson, B. D.; Goddard, W. A., Dreiding - a Generic Force-Field for
38
39 Molecular Simulations. J. Phys. Chem. 1990, 94, 8897-8909.
40 (40) Dubbeldam, D.; Calero, S.; Ellis, D. E.; Snurr, R. Q., RASPA: molecular simulation
41 software for adsorption and diffusion in flexible nanoporous materials. Mol. Simul.
42 2016, 42, 81-101.
43 (41) Wilmer, C. E.; Kim, K. C.; Snurr, R. Q., An Extended Charge Equilibration Method.
44 J. Phys. Chem. Lett. 2012, 3, 2506-2511.
45 (42) Horn, H. W.; Swope, W. C.; Pitera, J. W.; Madura, J. D.; Dick, T. J.; Hura, G. L.;
46 Head-Gordon, T., Development of an improved four-site water model for
47 biomolecular simulations: TIP4P-Ew. J. Chem. Phys. 2004, 120, 9665-9678.
48 (43) Clark, S. J.; Segall, M. D.; Pickard, C. J.; Hasnip, P. J.; Probert, M. J.; Refson, K.;
49 Payne, M. C., First principles methods using CASTEP. Z. Kristallogr. 2005, 220,
50
567-570.
51
52
(44) Neugebauer, J.; Hickel, T., Density functional theory in materials science. Wiley
53 Interdiscip. Rev. Comput. Mol. Sci. 2013, 3, 438-448.
54 (45) Kresse, G.; Furthmüller, J., Efficient iterative schemes for ab initio total-energy
55 calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169-11186.
56
57
58
59
60 ACS Paragon Plus Environment
30
Page 31 of 32 Langmuir

1
2
3 (46) Payne, M. C.; Teter, M. P.; Allan, D. C.; Arias, T. A.; Joannopoulos, J. D., Iterative
4 minimization techniques for ab initio total-energy calculations: molecular dynamics
5 and conjugate gradients. Rev. Mod. Phys. 1992, 64, 1045-1097.
6 (47) Perdew, J. P.; Burke, K.; Ernzerhof, M., Generalized gradient approximation made
7 simple. Phys. Rev. Lett. 1996, 77, 3865-3868.
8 (48) Monkhorst, H. J.; Pack, J. D., Special Points for Brillouin-Zone Integrations. Phys.
9 Rev. B 1976, 13, 5188-5192.
10
(49) Horn, H. W.; Swope, W. C.; Pitera, J. W., Characterization of the TIP4P-Ew water
11
12
model: Vapor pressure and boiling point. J. Chem. Phys. 2005, 123, 194504.
13 (50) Peng, X.; Lin, L.-C.; Sun, W.; Smit, B., Water adsorption in metal–organic
14 frameworks with open-metal sites. AIChE J. 2015, 61, 677-687.
15 (51) Paranthaman, S.; Coudert, F.-X.; Fuchs, A. H., Water adsorption in hydrophobic
16 MOF channels. Phys. Chem. Chem. Phys. 2010, 12, 8124-8130.
17 (52) Mercado, R.; Vlaisavljevich, B.; Lin, L.-C.; Lee, K.; Lee, Y.; Mason, J. A.; Xiao, D.
18 J.; Gonzalez, M. I.; Kapelewski, M. T.; Neaton, J. B.; Smit, B., Force Field
19 Development from Periodic Density Functional Theory Calculations for Gas
20 Separation Applications Using Metal–Organic Frameworks. J. Phys. Chem. C 2016,
21 120, 12590-12604.
22 (53) Lin, L.-C.; Lee, K.; Gagliardi, L.; Neaton, J. B.; Smit, B., Force-Field Development
23
from Electronic Structure Calculations with Periodic Boundary Conditions:
24
25
Applications to Gaseous Adsorption and Transport in Metal–Organic Frameworks. J.
26 Chem. Theory Comput. 2014, 10, 1477-1488.
27 (54) Akiyama, G.; Matsuda, R.; Kitagawa, S., Highly Porous and Stable Coordination
28 Polymers as Water Sorption Materials. Chem. Lett. 2010, 39, 360-361.
29 (55) Choi, J.; Lin, L.-C.; Grossman, J. C., The Role of Structural Defects in the Water
30 Adsorption Properties of MOF-801. J. Phys. Chem. C 2018.
31 (56) Cliffe, M. J.; Wan, W.; Zou, X. D.; Chater, P. A.; Kleppe, A. K.; Tucker, M. G.;
32 Wilhelm, H.; Funnell, N. P.; Coudert, F. X.; Goodwin, A. L., Correlated defect
33 nanoregions in a metal-organic framework. Nat. Commun. 2014, 5.
34 (57) Fang, Z. L.; Bueken, B.; De Vos, D. E.; Fischer, R. A., Defect-Engineered Metal-
35 Organic Frameworks. Angew. Chem., Int. Ed. 2015, 54, 7234-7254.
36
(58) Shearer, G. C.; Chavan, S.; Bordiga, S.; Svelle, S.; Olsbye, U.; Lillerud, K. P., Defect
37
Engineering: Tuning the Porosity and Composition of the Metal-Organic Framework
38
39 UiO-66 via Modulated Synthesis. Chem. Mater. 2016, 28, 3749-3761.
40 (59) Ohio Supercomputer Center. http://osc.edu/ark:/19495/f5s1ph73 (accessed Mar 6,
41 2018)
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
31
Langmuir Page 32 of 32

1
2
3 Table of Contents Graphic
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 ACS Paragon Plus Environment
32

Das könnte Ihnen auch gefallen