Thermal transport in interpenetrated metal-organic frameworks

Thermal transport in interpenetrated metal-organic frameworks. Kutay B. Sezginel, Patrick A. Asinger, Hasan Babaei, and Christopher E. Wilmer. Chem. M...
0 downloads 12 Views 1MB Size
Subscriber access provided by READING UNIV

Article

Thermal transport in interpenetrated metal-organic frameworks Kutay B. Sezginel, Patrick A. Asinger, Hasan Babaei, and Christopher E. Wilmer Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.7b05015 • Publication Date (Web): 16 Jan 2018 Downloaded from http://pubs.acs.org on January 16, 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 free 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 accessible to all readers and 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.

Chemistry of Materials 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 12 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

Chemistry of Materials

Thermal transport in interpenetrated metal-organic frameworks Kutay B. Sezginela, Patrick A. Asingera, Hasan Babaei*a and Christopher E. Wilmera a

Department of Chemical and Petroleum Engineering, University of Pittsburgh, 3700 O’Hara St, Pittsburgh, Pennsylvania 15261, USA *

Corresponding author. [email protected]

Abstract: In practice, the usefulness of metal-organic frameworks (MOFs) for many gas storage applications depends on their ability to rapidly dissipate the heat generated during the exothermic adsorption process. MOFs can be precisely designed to have a wide variety of architectures which can allow tuning their thermal conductivity. In this work, we use molecular dynamics simulations to investigate the effect of interpenetration on the thermal conductivity of MOFs. We find that the addition of a parallel thermal transport pathway yields a thermal conductivity nearly the sum of the two constituent frameworks. This relationship holds for a variety of interpenetrating MOFs with different atomic masses and a wide range of interframework interaction parameters. We show that both the strength and range of interactions between constituent frameworks play a significant role in framework mobility as well as framework coupling which can result in deviation from this relationship. We propose a simple model to predict thermal conductivity of the interpenetrated framework based on thermal conductivities of individual frameworks and an interframework interaction parameter which can account for this deviation.

ACS Paragon Plus Environment

Chemistry of Materials 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

Introduction Metal-organic frameworks (MOFs) show promise in a variety of gas adsorption applications such as gas storage1–5, gas separation6,7, sensing8,9, and catalysis10–12 due to their high porosity and internal surface area. A wide array of organic ligands and metal centers are available for the synthesis of MOFs, providing the opportunity for properties by design when the underlying principles are well understood.13 While the focus for structure-property relationships in MOFs has expectedly been gas-uptake capacity and selectivity, limited insight exists on the thermal transport characteristics of MOFs.14–23 Since the adsorption capacities of MOFs are reduced at higher temperatures, the exothermic process of gas adsorption can limit the rate at which tank-filling, separations, or other processes can be performed.24 Thus, investigating design principles for thermal transport in MOFs is critical for allowing these materials to reach their full potential for gas storage and separations applications.4 MOFs with particularly high void fractions are often capable of having interpenetrated25 or interwoven26 structures. Mostly, interpenetration is observed within the same framework (homointerpenetration). However, on rare occasions interpenetration of different frameworks (heterointerpenetration) have been observed as well. Even though interpenetration decreases the available space within the framework and increases heat of adsorption27, it has been shown to enhance framework stability28 and increase gas adsorption selectivity.29 Moreover, interpenetrated MOFs have sometimes shown enhanced gas adsorption capacity for small gases such as H2 and N2 by the introduction of additional adsorption sites.30,31 The formation, or not, of interpenetrated networks can often be controlled either by rational geometric design of the structure or manipulation of experimental conditions. Deliberate modification of organic ligands,32 using infinite secondary building units33,34 and choosing solvents with appropriate molecular size35 have been used as geometrical restraints to control interpenetration whereas manipulating reactant concentration and reaction temperature were used to control interpenetration by regulating reaction kinetics.36 Additionally, reversible interpenetration control was achieved by ligand removal and addition37 and anisotropic entanglement was realized by partial interpenetration.38 Using these experimental methods, advanced control of final architecture can be achieved by controlling degree of interpenetration39,40, reaching up to 54-fold interpenetration41, as well as modifying interpenetration distance to design minimally interpenetrating (interweaving) MOFs.26 It was also shown that computational methods can also be used to rationally design interpenetrated structures.42,43 Recently, we developed an algorithm to design hetero-interpenetrated MOFs with which we were able to discover 18 novel hetero-interpenetrating hypothetical structures.43 As we will show in this Letter, the advanced control over MOF interpenetration demonstrated by others can be leveraged as a valuable way to tune thermal transport behavior. Interpenetrating frameworks provide parallel pathways for thermal transport. The effect of thermal coupling between frameworks, however, has so far been uncertain. In the present work, we use molecular dynamics (MD) simulations to investigate thermal transport in a variety of doubly interpenetrated MOFs. To isolate the effects of interpenetration from other structural parameters, a simple cubic idealized framework was considered. Effects of pore structure on thermal conductivity have been reported elsewhere.18 We find interpenetration yields a system thermal conductivity approximately equal to the sum of the two independent frameworks, which can be used as a “rule of thumb” for design purposes. To account for deviations from this general rule, we present equations that can be used as guidelines for predicting thermal conductivity in doubly interpenetrated MOF structures.

ACS Paragon Plus Environment

Page 2 of 12

Page 3 of 12 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

Chemistry of Materials

Methods We used an idealized cubic MOF structure inspired by IRMOF-1 to better understand parameters that affect thermal transport, as we had done in prior studies17,18 (see Figure 1a). The idealized interpenetrated MOF allows us to study the effect of framework interactions parametrically which would not be possible in a real interpenetrated MOF. We use two-body bonded interactions between atoms within each of interpenetrated structures, modeled using the Morse potential (see Figure 1b).44 The force field parameters were chosen so that the thermal conductivity of the simple cubic structure with a pore size of 1 nm was of the same order as a typical MOF (~1 W/m K). The interpenetrated structure is generated by creating a copy of the idealized structure and translating it 5 Å in each dimension which corresponds to the maximal distance between frameworks (see Figure 1c). We also calculated the thermal conductivity for IRMOF-1 and its doubly interpenetrated MOF using UFF45 parameters which were shown to provide accurate values for predicting bulk modulus.46 All thermal conductivity predictions were done using the Green-Kubo approach and equilibrium molecular dynamics (MD) simulations (LAMMPS47). The force field parameters for structures and MD simulation details are provided in Supplementary Information, with sample LAMMPS47 input files provided in therMOF GitHub repository: https://github.com/kbsezginel/thermof.

Figure 1 (a) Idealized porous crystal (8 × 8 × 8 cubic unit cells). (b) Bonding arrangement for a single unit cell using Morse potentials (red bonds are modeled more strongly than blue bonds). (c) Doubly interpenetrated unit cell with framework depicted as red and blue (initial frameworks in each simulation are 5 Å apart in each dimension. (d) Interpenetrated idealized porous crystal (8 × 8 × 8 cubic unit cells). See Supplementary Information for more details about the structures and simulation details.

The non-bonded interactions between interpenetrating frameworks were modeled by a LennardJones potential: 𝜎 12 𝜎 6 𝑉𝐿𝐽 = 4𝜀 [( ) − ( ) ] 𝑟 𝑟

𝐸𝑞. 1

where 𝜀 is the depth of the potential well and 𝜎 is the distance at which the interatomic potential is zero. A range of 𝜀 values between 0.01 – 1 kcal/mol and 𝜎 values between 1 – 6 Å were chosen to systematically understand the effects of framework interactions on thermal transport. Then, an 𝜀 value of 0.105 kcal/mol (corresponding to a carbon atom in UFF45) and 𝜎 values of 3.5 Å and 4.5 Å were selected to represent different interpenetration configurations for further analysis. The 𝜎 influences framework mobility where a value of 3.5 Å allows interpenetrating frameworks to “freely” translate relative to each by 2 Å in each direction, while a 𝜎 value of 4.5 Å models interpenetrated frameworks that are “locked in” and whose distance from each other remains constant. These configurations can be imagined as, for example,

ACS Paragon Plus Environment

Chemistry of Materials 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 12

isoreticular MOFs with more or less bulky ligands (see IRMOF-13 vs. IRMOF-9 in Figure 2) that correspond to interpenetrated frameworks that are “locked in” or “free”, respectively.

Figure 2 Different interpenetration configurations based on cubic IRMOF series 48: (a) metal center distribution (b) BPDC linker (IRMOF-9) and PDC linker (IRMOF-13) representing less and more bulky linkers, respectively. Color scheme is as follows: Zn (blue spheres), O (red spheres), C (gray spheres).

Results and Discussion Framework interaction dependency of constituent position and thermal conductivity The effect of inter-framework interactions on thermal conductivity and framework mobility was investigated by calculating thermal conductivity and MSD for a wide range of ε and σ values (see Figure 3). On average, the thermal conductivity for the interpenetrated framework was found to be around 1.6 W/mK which is approximately double that of the single framework (0.82 W/mK). Consequently, this led us to explore the idea that the thermal conductivity of the interpenetrating frameworks can be predicted by a simple linear combination rule: 𝑘𝐼𝑃 = 𝑘1 + 𝑘2

𝐸𝑞. 2

where kIP, k1, and k2 are the thermal conductivities of the interpenetrating framework, single framework 1 and single framework 2, respectively. In order to understand the conditions under which this relationship holds, percentage error for predicting thermal conductivity using Eq. 2 was calculated (Figure 3b). As seen in Figure 3a and 3b the thermal conductivity is relatively constant in the region 0.01 < ε < 0.2 kcal/mol and 1 < σ < 4.5 Å. However, increasing ε above 0.1 kcal/mol decreases thermal conductivity for σ values lower than 4.5 Å as apparent from the dark blue region in Figure 3a. This low thermal conductivity region described by 1.5 ≤ σ ≤ 4.5 Å and 0.1 ≤ ε ≤ 1.0 corresponds to increased coupling of the frameworks, which likely leads to higher phonon scattering rates introduced by the interactions between two frameworks. We then modified Eq. 2 by adding a coupling constant (γ) that depends on the interframework interactions as follows: 𝑘𝐼𝑃 = (𝑘1 + 𝑘2 )(1 − 𝛾)

𝐸𝑞. 3

We defined individual coupling constants for σ and ε as 𝛾 = 𝛾𝜀 𝛾𝜎 and derived equations to represent the coupling seen for the region 1.5 ≤ σ ≤ 4.5 Å and 0.1 ≤ ε ≤ 1.0. Derivation of the equations for coupling constants are given in section 3b of Supplementary Information. As seen in Figure 3c the

ACS Paragon Plus Environment

Page 5 of 12 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

Chemistry of Materials

coupling constant was modeled to increase linearly with ε and change quadratically with σ between 0 – 0.25. Including the coupling constant in Eq. 3 decreases the average error for the whole interaction range from 13.9 % to 7.8 % (Fig 3d). We limited the region of the coupling constant because we believe the deviations seen from Eq. 2 outside this region are not related to framework coupling. It can be seen in Figure 3a that increasing σ above 4.5 Å increases the system thermal conductivity especially at higher ε values. This increase is a result of stronger interactions between frameworks providing better thermal transport pathways. As for σ < 1.5 Å we observe very high framework mobility (Fig 3e) which is attained to unphysically low repulsive forces causing frameworks to slide through each other. As a proof of concept, we calculated thermal conductivity for single and interpenetrated IRMOF-1. The thermal conductivity for the interpenetrated IRMOF-1 was found to be 0.68 W/mK with a standard deviation of 0.04 W/mK over 10 runs. The single framework thermal conductivity was found to be 0.34 W/mK again with a standard deviation of 0.04 W/mK. According to Eq. 2 the thermal conductivity for the interpenetrated framework is predicted as 0.68 W/mK, which is in excellent agreement with the simulated value and suggests no significant interframework coupling.

Figure 3 Effect of framework interaction on thermal conductivity and framework mobility in terms of forcefield parameters. For each plot, each bin represents a simulation with corresponding ε and σ values. The black bins on the upper right-hand side are for simulations where high repulsive forces caused frameworks to collapse. (a) Thermal conductivity (W/m K). (b) Thermal conductivity prediction error (%) for interpenetrated framework using the relationship in equation 2. (c) Coupling constant (γ). (d) Thermal conductivity prediction error (%) for interpenetrated framework using the relationship in equation 2. (e) Mean squared displacement, MSD (Å2).

Framework mobility is also significantly affected by the force field parameters as evident from the MSD shown in Figure 3e. In addition to MSD, the framework mobility was investigated by calculating the relative distance between reference atoms selected from interpenetrating frameworks (see Fig 4). The relative distance of one corner atom of the second framework (Fig 4d shown in red) to the 8 corner atoms of the surrounding first framework (Fig 4d shown in blue) were calculated from MD trajectories. The resulting 2D histograms given in Figure 4 show where the framework spends time during the simulation. The effect of ε is shown in Figure 4a for σ 3.5 Å and Figure 4b for σ 4.5 Å and the effect of σ is shown in Figure 4c.

ACS Paragon Plus Environment

Chemistry of Materials 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

Generally, σ is inversely related to the total area the frameworks can translate relative to each other whereas ε is directly related with the amount of time frameworks spend in a given configuration. Consequently, MSD is generally higher at lower ε and σ values. Above σ value of 4 Å, the frameworks are tightly locked therefore any change in force field parameters do not reflect directly on the MSD. Below 4 Å, however, MSD increases with decreasing σ as seen in Figure 3e. For lower σ values, there are multiple potential wells in between the frameworks. As a result, frameworks reside in different potential wells during each run and can jump between wells during simulation (see Figure 4a). With increasing ε, the energy barrier to jump between potential wells increases and frameworks become more closely coupled. Depending on the initial velocities, frameworks are trapped (for the duration of the simulation) in a position relative to each other and stay there during simulation (Fig 4c). For low ε values, however, frameworks easily move relative to each other (Fig 4a). Given an infinite simulation time, the average distance between frameworks should correspond to optimal distance since, by symmetry, each corner would be visited equally many times resulting in complete sampling of the configuration space.

Figure 4 Effect of interpenetration interaction on framework distances during simulation. Framework interactions are modeled using Lennard-Jones potential. (a) σ: 3.5 Å and (b) σ: 4.5 Å with varying ε values in kcal/mol shown in figure. Grid size: 6 Å and bin size: 1 Å (c) ε: 0.105 kcal/mol and varying σ values shown in figure. Grid size: 12 Å and bin size: 2 Å (d) Framework distances are measured according to middle red atom with reference to corner atoms of the surrounding blue framework. 5 Å × 5 Å × 5 Å cube shown as dashed box in 2 dimensions. The yellow-red bins represent the position of the other framework. (e) Time spent in given location are given in percentage. Color scale is limited to 50 % of simulation time for clarity. All results are for x and y directions from a single run. Methods used for calculating relative distance are given in section 3c of Supplementary Information.

For low σ values, the frameworks have increased mobility as the amount of space they can translate with respect to each other with the same energy penalty is increased. This is also complemented by the MSD (Fig 3e) where the displacement generally increases towards lower σ values. As seen in Fig 3e, between σ values of 1 – 2 Å the displacement is quite high and between 2 – 4 Å the displacement is moderate, decreasing as σ goes higher. For σ ≥ 2 Å increasing ε slightly decreases MSD since fewer potential wells are visited during a simulation. As the MSD is calculated in reference to the initial position of the frameworks, simulations with higher ε values result in frameworks spending more time on a single potential well distant from their initial position. For lower ε values, however, as the frameworks travel

ACS Paragon Plus Environment

Page 6 of 12

Page 7 of 12 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

Chemistry of Materials

between potential wells they get closer to their initial position, consequently decreasing the MSD. Again, with better sampling of the configuration space these differences would be minimal. The high MSD for σ < 2 Å is due to frameworks sliding through each other since interatomic distances within a framework are 3.3 Å. In this region increasing ε decreases the MSD as the interaction strength is increased. However, these extraordinary framework interactions are not expected to be observed in real MOFs. Mass dependency of thermal conductivity Initially thermal conductivities for a single framework were investigated for different atomic masses between 4 to 200 Da (i.e., changing the mass of every atom in the framework, see Figure 5). As seen in Figure 5a, thermal conductivity as a function of atomic mass shows a 1/√𝑀 relation. As previously shown by others,49 this is because frequency and group velocity of acoustic phonons in the lattice scales as 1/√𝑀 whereas the mean free path and specific heat do not change considerably with mass. We then investigated interpenetrated frameworks, where the atomic mass of one of the frameworks was changed and the other kept constant. We found the same relationship to hold for both “free” and “locked” interpenetration (Fig 5b).

Figure 5 Effect of atomic mass on thermal conductivity (a) Single MOF (b) Free interpenetration (σ: 3. 5 Å) Dashed lines are equations shown in the figure. (c) Parity plot for predicted (Eq. 1) and simulated thermal conductivities for “free” and “locked” interpenetrated MOFs for different masses. For interpenetrated structures, atomic mass of the second framework (M2) is modified whereas first framework atoms are always chosen as Ar (M1~40 Da). Error bars represent one standard deviation of uncertainty.

Moreover, thermal conductivity values for the interpenetrated framework is approximately the sum of thermal conductivity of individual frameworks. For instance, thermal conductivity for individual frameworks made of atoms that have masses of either, 39.95 Da (Ar), or 83.80 Da (Kr), have thermal conductivities of 0.83 W/mK and 0.57 W/mK, respectively. Summing their individual thermal conductivities gives 1.40 W/m K, which is between 1.49 and 1.34 W/m K given by the interpenetrated frameworks having σ values of 4.5 Å and 3.5 Å respectively. As shown in Figure 5c, the predicted thermal conductivities (using Eq. 2) for different masses match closely with simulated thermal conductivity values. This suggests that a linear combination of single-framework thermal conductivities is a good model for the conductivities of interpenetrated frameworks, even when their corresponding masses are different. Conclusion The work presented here shows that, based on MD simulations, doubly interpenetrated MOFs will have thermal conductivity approximately the sum of the constituent frameworks. We show that the relation holds for both cubic idealized MOFs and IRMOF-1 as presented here. For rare cases where

ACS Paragon Plus Environment

Chemistry of Materials 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

frameworks are not maximally interpenetrating and interframework attraction is high, the effect of coupling might result in lower thermal conductivity than the sum of individual frameworks. This suggests that the nature of interpenetration is quite important for thermal conductivity. We show that the decrease in thermal conductivity can be approximated by calculating a coupling factor using force field parameters (σ and ε) that define interframework interactions. For interactions described by 0.01 < ε < 0.1 kcal/mol and 1 < σ < 4.5 Å the effect of coupling was found to be minimal. We believe these results provide important insights into the gas adsorption application of interpenetrated MOFs especially for small gasses such as H2. The inclusion of additional frameworks could increase storage capacity of the gas by increasing adsorption sites and provide additional thermal transport pathways which allows the generated heat to dissipate more quickly. Associated Content Supporting Information Generation of idealized MOF structures, force field parameters for idealized MOFs and IRMOF-1. Molecular dynamics simulation details, construction of framework coupling model, framework mobility calculations (relative distance between frameworks and mean squared displacement). Acknowledgements C. E. W., H. B. and K. B. S. gratefully acknowledge the Donors of the American Chemical Society Petroleum Research Fund for support of this research. They also acknowledge both the Swanson School of Engineering and the Center for Simulation and Modeling (SAM) at the University of Pittsburgh for early financial support, and for providing computational resources, respectively.

ACS Paragon Plus Environment

Page 8 of 12

Page 9 of 12 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

Chemistry of Materials

References (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13)

(14)

(15)

(16) (17) (18) (19)

(20)

Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. The Chemistry and Applications of MetalOrganic Frameworks. Science 2013, 341 (6149), 1230444. Rosi, N. L.; Eckert, J.; Eddaoudi, M.; Vodak, D. T.; Kim, J.; O’Keeffe, M.; Yaghi, O. M. Hydrogen Storage in Microporous Metal-Organic Frameworks. Science 2003, 300 (5622), 1127–1129. He, Y.; Zhou, W.; Qian, G.; Chen, B. Methane Storage in Metal–organic Frameworks. Chem. Soc. Rev. 2014, 43 (16), 5657–5678. Mason, J. A.; Veenstra, M.; Long, J. R. Evaluating Metal–organic Frameworks for Natural Gas Storage. Chem. Sci. 2014, 5 (1), 32–51. Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm, Z. R.; Bae, T.-H.; Long, J. R. Carbon Dioxide Capture in Metal–organic Frameworks. Chem. Rev. 2011, 112 (2), 724–781. Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective Gas Adsorption and Separation in Metal–organic Frameworks. Chem. Soc. Rev. 2009, 38 (5), 1477–1504. Qiu, S.; Xue, M.; Zhu, G. Metal–organic Framework Membranes: From Synthesis to Separation Application. Chem. Soc. Rev. 2014, 43 (16), 6116–6140. Kreno, L. E.; Leong, K.; Farha, O. K.; Allendorf, M.; Van Duyne, R. P.; Hupp, J. T. Metal–organic Framework Materials as Chemical Sensors. Chem. Rev. 2011, 112 (2), 1105–1125. Hu, Z.; J. Deibert, B.; Li, J. Luminescent Metal–organic Frameworks for Chemical Sensing and Explosive Detection. Chem. Soc. Rev. 2014, 43 (16), 5815–5840. Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Metal–organic Framework Materials as Catalysts. Chem. Soc. Rev. 2009, 38 (5), 1450–1459. Liu, J.; Chen, L.; Cui, H.; Zhang, J.; Zhang, L.; Su, C.-Y. Applications of Metal–organic Frameworks in Heterogeneous Supramolecular Catalysis. Chem. Soc. Rev. 2014, 43 (16), 6011–6061. Zhang, T.; Lin, W. Metal–organic Frameworks for Artificial Photosynthesis and Photocatalysis. Chem. Soc. Rev. 2014, 43 (16), 5982–5993. Lu, W.; Wei, Z.; Gu, Z.-Y.; Liu, T.-F.; Park, J.; Park, J.; Tian, J.; Zhang, M.; Zhang, Q.; Gentle III, T. Tuning the Structure and Function of Metal–organic Frameworks via Linker Design. Chem. Soc. Rev. 2014, 43 (16), 5561–5593. Huang, B. L.; McGaughey, A. J. H.; Kaviany, M. Thermal Conductivity of Metal-Organic Framework 5 (MOF-5): Part I. Molecular Dynamics Simulations. Int. J. Heat Mass Transf. 2007, 50 (3–4), 393– 404. Huang, B. L.; Ni, Z.; Millward, A.; McGaughey, A. J. H.; Uher, C.; Kaviany, M.; Yaghi, O. Thermal Conductivity of a Metal-Organic Framework (MOF-5): Part II. Measurement. Int. J. Heat Mass Transf. 2007, 50 (3), 405–411. Zhang, X.; Jiang, J. Thermal Conductivity of Zeolitic Imidazolate Framework-8: A Molecular Simulation Study. J. Phys. Chem. C 2013, 117 (36), 18441–18447. Babaei, H.; Wilmer, C. E. Mechanisms of Heat Transfer in Porous Crystals Containing Adsorbed Gases: Applications to Metal-Organic Frameworks. Phys. Rev. Lett. 2016, 116 (2), 25902. Babaei, H.; McGaughey, A. J.; Wilmer, C. E. Effect of Pore Size and Shape on the Thermal Conductivity of Metal-Organic Frameworks. Chem. Sci. 2017, 8 (1), 583–589. Wang, X.; Guo, R.; Xu, D.; Chung, J.; Kaviany, M.; Huang, B. Anisotropic Lattice Thermal Conductivity and Suppressed Acoustic Phonons in MOF-74 from First Principles. J. Phys. Chem. C 2015, 119 (46), 26000–26008. Cui, B.; Audu, C. O.; Liao, Y.; Nguyen, S. T.; Farha, O. K.; Hupp, J. T.; Grayson, M. Thermal Conductivity of ZIF-8 Thin-Film under Ambient Gas Pressure. ACS Appl. Mater. Interfaces 2017, 9 (34), 28139–28143.

ACS Paragon Plus Environment

Chemistry of Materials 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

(21) (22)

(23)

(24)

(25) (26) (27) (28) (29)

(30) (31) (32)

(33) (34)

(35)

(36)

(37)

(38)

Han, L.; Budge, M.; Greaney, P. A. Relationship between Thermal Conductivity and Framework Architecture in MOF-5. Comput. Mater. Sci. 2014, 94, 292–297. Liu, D.; Purewal, J. J.; Yang, J.; Sudik, A.; Maurer, S.; Mueller, U.; Ni, J.; Siegel, D. J. MOF-5 Composites Exhibiting Improved Thermal Conductivity. Int. J. Hydrog. Energy 2012, 37 (7), 6109– 6117. Ming, Y.; Purewal, J.; Sudik, A.; Xu, C.; Yang, J.; Veenstra, M.; Rhodes, K.; Soltis, R.; Warner, J.; Gaab, M. Thermophysical Properties of MOF-5 Powders. Microporous Mesoporous Mater. 2014, 185, 235–244. Zhang, H.; Deria, P.; K. Farha, O.; T. Hupp, J.; Q. Snurr, R. A Thermodynamic Tank Model for Studying the Effect of Higher Hydrocarbons on Natural Gas Storage in Metal–organic Frameworks. Energy Environ. Sci. 2015, 8 (5), 1501–1510. Gong, Y.-N.; Zhong, D.-C.; Lu, T.-B. Interpenetrating Metal–organic Frameworks. CrystEngComm 2016, 18 (15), 2596–2606. Chen, B.; Eddaoudi, M.; Hyde, S.; O’keeffe, M.; Yaghi, O. Interwoven Metal-Organic Framework on a Periodic Minimal Surface with Extra-Large Pores. Science 2001, 291 (5506), 1021–1023. Kesanli, B.; Cui, Y.; Smith, M. R.; Bittner, E. W.; Bockrath, B. C.; Lin, W. Highly Interpenetrated Metal–Organic Frameworks for Hydrogen Storage. Angew. Chem. Int. Ed. 2005, 44 (1), 72–75. Ma, L.; Lin, W. Unusual Interlocking and Interpenetration Lead to Highly Porous and Robust Metal– organic Frameworks. Angew. Chem. Int. Ed. 2009, 48 (20), 3637–3640. Ma, S.; Wang, X.-S.; Yuan, D.; Zhou, H.-C. A Coordinatively Linked Yb Metal–Organic Framework Demonstrates High Thermal Stability and Uncommon Gas-Adsorption Selectivity. Angew. Chem. Int. Ed. 2008, 47 (22), 4130–4133. Kim, H.; Das, S.; Kim, M. G.; Dybtsev, D. N.; Kim, Y.; Kim, K. Synthesis of Phase-Pure Interpenetrated MOF-5 and Its Gas Sorption Properties. Inorg. Chem. 2011, 50 (8), 3691–3696. Jiang, H.-L.; Makal, T. A.; Zhou, H.-C. Interpenetration Control in Metal–organic Frameworks for Functional Applications. Coord. Chem. Rev. 2013, 257 (15–16), 2232–2249. Farha, O. K.; Malliakas, C. D.; Kanatzidis, M. G.; Hupp, J. T. Control over Catenation in Metal− Organic Frameworks via Rational Design of the Organic Building Block. J. Am. Chem. Soc. 2009, 132 (3), 950–952. Rosi, N. L.; Eddaoudi, M.; Kim, J.; O’Keeffe, M.; Yaghi, O. M. Infinite Secondary Building Units and Forbidden Catenation in Metal-Organic Frameworks. Angew. Chem. Int. Ed. 2002, 41 (2), 284–287. Deng, H.; Grunder, S.; Cordova, K. E.; Valente, C.; Furukawa, H.; Hmadeh, M.; Gándara, F.; Whalley, A. C.; Liu, Z.; Asahina, S. Large-Pore Apertures in a Series of Metal-Organic Frameworks. science 2012, 336 (6084), 1018–1023. Song, F.; Wang, C.; Falkowski, J. M.; Ma, L.; Lin, W. Isoreticular Chiral Metal− Organic Frameworks for Asymmetric Alkene Epoxidation: Tuning Catalytic Activity by Controlling Framework Catenation and Varying Open Channel Sizes. J. Am. Chem. Soc. 2010, 132 (43), 15390–15398. Rankine, D.; Avellaneda, A.; Hill, M. R.; Doonan, C. J.; Sumby, C. J. Control of Framework Interpenetration for in Situ Modified Hydroxyl Functionalised IRMOFs. Chem. Commun. 2012, 48 (83), 10328–10330. Choi, S. B.; Furukawa, H.; Nam, H. J.; Jung, D.-Y.; Jhon, Y. H.; Walton, A.; Book, D.; O’Keeffe, M.; Yaghi, O. M.; Kim, J. Reversible Interpenetration in a Metal–Organic Framework Triggered by Ligand Removal and Addition. Angew. Chem. Int. Ed. 2012, 51 (35), 8791–8795. Ferguson, A.; Liu, L.; Tapperwijn, S. J.; Perl, D.; Coudert, F.-X.; Van Cleuvenbergen, S.; Verbiest, T.; van der Veen, M. A.; Telfer, S. G. Controlled Partial Interpenetration in Metal–organic Frameworks. Nat. Chem. 2016, 8 (3), 250–257.

ACS Paragon Plus Environment

Page 10 of 12

Page 11 of 12 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

Chemistry of Materials

(39)

(40)

(41)

(42) (43) (44) (45)

(46) (47) (48)

(49)

Reineke, T. M.; Eddaoudi, M.; Moler, D.; O’keeffe, M.; Yaghi, O. Large Free Volume in Maximally Interpenetrating Networks: The Role of Secondary Building Units Exemplified by Tb2 (ADB) 3 [(CH3) 2SO] 4⊙ 16 [(CH3) 2SO] 1. J. Am. Chem. Soc. 2000, 122 (19), 4843–4844. Long, D.-L.; Hill, R. J.; Blake, A. J.; Champness, N. R.; Hubberstey, P.; Wilson, C.; Schröder, M. Anion Control over Interpenetration and Framework Topology in Coordination Networks Based on Homoleptic Six-Connected Scandium Nodes. Chem.- Eur. J. 2005, 11 (5), 1384–1391. Wu, H.; Yang, J.; Su, Z.-M.; Batten, S. R.; Ma, J.-F. An Exceptional 54-Fold Interpenetrated Coordination Polymer with 103-Srs Network Topology. J. Am. Chem. Soc. 2011, 133 (30), 11406– 11409. Kwon, O.; Park, S.; Zhou, H.-C.; Kim, J. Computational Prediction of Hetero-Interpenetration in Metal–organic Frameworks. Chem. Commun. 2017, 53 (12), 1953–1956. Sezginel, K. B.; Feng, T.; Wilmer, C. E. Discovery of Hypothetical Hetero-Interpenetrated MOFs with Arbitrarily Dissimilar Topologies and Unit Cell Shapes. CrystEngComm 2017, 19 (31), 4497–4504. Morse, P. M. Diatomic Molecules according to the Wave Mechanics. II. Vibrational Levels. Phys. Rev. 1929, 34 (1), 57. Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard, W. A.; Skiff, W. M. UFF, a Full Periodic Table Force Field for Molecular Mechanics and Molecular Dynamics Simulations. J. Am. Chem. Soc. 1992, 114 (25), 10024–10035. Boyd, P. G.; Moosavi, S. M.; Witman, M.; Smit, B. Force-Field Prediction of Materials Properties in Metal-Organic Frameworks. J. Phys. Chem. Lett. 2017, 8 (2), 357–363. Plimpton, S. Fast Parallel Algorithms for Short-Range Molecular Dynamics. J. Comput. Phys. 1995, 117 (1), 1–19. Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’Keeffe, M.; Yaghi, O. M. Systematic Design of Pore Size and Functionality in Isoreticular MOFs and Their Application in Methane Storage. Science 2002, 295 (5554), 469–472. Che, J.; Çağın, T.; Deng, W.; Goddard III, W. A. Thermal Conductivity of Diamond and Related Materials from Molecular Dynamics Simulations. J. Chem. Phys. 2000, 113 (16), 6888–6900.

ACS Paragon Plus Environment

Chemistry of Materials 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

Table of Content Graphic

ACS Paragon Plus Environment

Page 12 of 12