Thermal Conductivity of 3D Boron-Based Covalent Organic

Jul 14, 2016 - Thermal Conductivity of 3D Boron-Based Covalent Organic Frameworks ... The Supporting Information is available free of charge on the AC...
0 downloads 0 Views 1MB Size
Subscriber access provided by the University of Exeter

Article

Thermal Conductivity of 3D Boron-based Covalent Organic Frameworks from Molecular Dynamics Simulations Yazhou Liu, Yanhui Feng, Zhi Huang, and Xinxin Zhang J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.6b04891 • Publication Date (Web): 14 Jul 2016 Downloaded from http://pubs.acs.org on July 19, 2016

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.

The Journal of Physical Chemistry C 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 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

Thermal Conductivity of 3D Boron-based Covalent Organic Frameworks from Molecular Dynamics Simulations Yazhou Liu †, Yanhui Feng *, †, ‡, Zhi Huang †, ‡, Xinxin Zhang †, ‡ † School of Mechanical Engineering, University of Science and Technology Beijing, Beijing 100083, China ‡ Beijing Key Laboratory of Energy Saving and Emission Reduction for Metallurgical Industry, University of Science and Technology Beijing, Beijing 100083, China

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

ABSTRACT

Covalent organic frameworks (COFs) have been widely investigated for use in gas storage and separation, while their thermal properties have been scarcely studied. In the study reported in this paper, the thermal conductivities of 3D boron-based COFs were investigated for the first time using molecular dynamics simulations (MD) employing the Green-Kubo method. The predicted thermal conductivities of COF-102, COF-103, COF-105 and COF-108 were on the order of 0.1 W/(m⋅K) at 300K. The thermal conductivity decreased by up to 47% with the increase in temperature from 200 to 500K. This resulting low thermal conductivity was due to the short mean free path of the phonon in the COFs, which was deduced to be 2.7-9.2 nm. The lowfrequency phonon modes below 50 THz contributed mostly to heat conduction. By analyzing the phonon vibrational density of states and overlap energy between per two bonded atoms, it was revealed that the connection between phenylene rings in COF-102 and COF-103 weakens the phonon coupling and then harms the energy flow, as did the connection between phenylene and triphenylene rings in COF-105 and COF-108. In addition, COF-105 had the lowest total overlap energy between all atoms, leading to the minimum thermal conductivity in the four COFs. This study provided a quantitative prediction of the thermal conductivities of COFs and microscopic insight into the mechanism of heat transfer.

KEYWORDS Thermal conductivity, COFs, Molecular dynamics simulation, Vibrational density of states, Overlap energy

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

1. INTRODUCTION Covalent organic frameworks (COFs) are an emerging class of crystalline materials, which are composed of light elements, typically H, B, C, N, Si and O that are linked via strong covalent bonds. Two-dimensional (2D) COFs (termed COF-1 and COF-5) and three-dimensional (3D) COFs (termed COF-102, COF-103, COF-105, and COF-108) were first synthesized by Yaghi’s group in 20051 and 20072. Similar to the so-called metal organic frameworks (MOFs)3, COFs have highly ordered internal structures, high porosity, good thermal stability and large surface areas, which make them particularly useful for applications such as gas adsorption. Compared with the MOFs, the molecular frameworks of COFs are composed of light elements that act to lower the relative weight of the structure. The synthesis of the COF materials can be tailored so that the material’s architecture including pore size, volume, and functionality can meet the demands of a particular application. COFs have been considered to be promising candidates for many potential applications such as gas storage, photo electricity and catalysis. In 2010, E. Klontzas et al.4,5 investigated newly designed 3D COFs (COF-102-2, COF-102-3, COF-102-4, and COF-102-5) for enhanced hydrogen storage capacity. In 2015, Fang et al.6 developed 3D porous crystalline polyimide COFs for drug delivery by choosing tetrahedral building units of various sizes. To date, most experimental and simulation studies on COFs have focused primarily on synthesis, gas adsorption7-9 and gas storage.10,11However, the physical properties of COFs have been barely studied. Zhao et al.12 conducted a computational study on the thermal expansion behavior of COFs and found that COFs exhibit negative thermal expansion, similar to MOFs.13 B. Lukose et al.14 performed theoretical studies on 3D COFs, including COF-102, COF-103, COF-105 and COF-108, and used density functional based methods to explore their structural,

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

electronic, energetic and mechanical properties. The results demonstrated that COFs’ mechanical stability is in accordance with MOFs, and both their bulk modulus do not exceed 20 GPa. In addition, all four COFs in this reported study were semiconductors with band gaps ranging from 2 to 4 eV. Saeed Amirjalayer et al.15 predicted that the low density materials COF-105 and COF108 show bulk moduli below MOF-5. And the ctn (carbon nitride, I 43d ) networks of COF-102 and COF-103 show values for bulk modulus, which are comparable with MOF-5, emphasizing their structural stiffness. The thermal transport properties of COFs play an important role in gas adsorption, because the adsorption in nanoporous materials is an exothermic process. Consequently, it affects the heat transfer in a gas storage system and thereby affects the system temperature as well as gas storage capacity. Consequently, appropriate thermal management is required to retain energy efficiency.16Furthermore, heat transfer in nanoporous materials is also important in applications in the optical, electronic, and optoelectronic fields.17 However, to the best of our knowledge, there have been no studies reported that deal with the thermal conductivity of COFs. For comparison, however, one can consider recent work conducted on the thermal conductivity of porous MOFs. For example, in 2006, Huang et al.18 predicted the thermal conductivity of MOF-5 in the temperature range of 200-400 K using molecular dynamics. The simulated thermal conductivity was found to be low for a crystal (0.31 W/(m⋅K) at a temperature of 300 K) and showed a weak temperature-dependence. Also, Huang et al.19 measured the thermal conductivity of MOF-5 single cubic crystal (1-2 mm in size) over a wide temperature range of 6-300K, using the longitudinal and steady-state heat flow method. These author’s experimental value for the thermal conductivity was 0.32 W/(m⋅K) at 300 K. In 2013, Zhang et al.20 predicted that the thermal conductivity of Zeolitic Imidazolate Framework-8 (ZIF-8) was about 0.165 W/(m⋅K)

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

using equilibrium molecular dynamics (EMD) simulation. Such a low thermal conductivity was due to the short mean free path of phonon in ZIF-8, which was estimated to be less than two unit cells. Qian et al.21 studied the thermal conductivity of a layered organic-inorganic hybrid crystal

β − ZnTe(en) . using the interatomic potential (Morse potential) derived from ab initio simulations. Because of the incorporation of organic ligands, the hybrid crystal β − ZnTe(en) .

is much more flexible than the inorganic crystal ZnTe , and the predicted thermal conductivities of the hybrid crystal were found to be one order of magnitude lower than those of inorganic

ZnTe crystal based on either experimental22 or MD simulation23 results. Wang et al.24 calculated the anisotropic thermal conductivities of water-stable MOF-74 using the Boltzmann transport equation (BTE) and the density-functional-based tight-binding method. They determined that the thermal conductivity to be 0.44 and 0.68 W/(m⋅K) at 300 K, across and along the pore directions. In addition, J. Purewal et al.25 in 2012 reported an effective thermal conductivity of around 0.1 W/(m⋅K) for low-density MOF-5 pellets made of compressed powders, which is close to the result from a later study(0.091W/(m⋅K),T=300K) conducted by Y. Ming et al..26 To improve the thermal conductivity of MOF-5, Liu et al. synthesized a series of high-density MOF-5 composites containing 0-10wt % expanded natural graphite (ENG), which enhanced the thermal conduction and found that the thermal conductivity was increased by a factor of 5. Although the thermal conductivities of some MOFs have been obtained, the relationship between thermal conductivity and microscopic structure of the materials is still not clear. Huang et al. indicated that the C1/C2 atoms act as a bottleneck in MOF-5.18 Zhang et al. considered Zn and N atoms as the key component for the transfer of energy in ZIF-8.20 However, neither of these investigators performed a quantitative analysis of the material in question. Single crystals of MOF materials can often be obtained and directly subjected to the structural identification using single crystal X-

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

ray diffraction. Unfortunately, to date, no single crystal of COF materials have been successfully synthesized. Consequently, the experimental measurement on the thermal conductivity on COFs is still not practically possible. In a word, to explore the potential applications of COFs, it is important to quantitatively know their thermal properties and understand the underlying mechanism of heat transfer. In this reported study, we performed EMD simulation was performed to predict the thermal conductivity of four types of 3D COFs (COF-102, COF-103, COF-105 and COF-108). The temperature dependence of the thermal conductivity was discussed. The relative contributions of different phonon oscillation frequencies to the thermal conductivity, the phonon vibrational density of states (VDOS) and the overlap phonon energy were analyzed to determine which atoms are mismatched in density of states. The heat transfer bottleneck in the nanostructure was found using quantitative analyses. This work was intended to lead to an exploration of the thermal properties of COFs and provide some fundamental data for the efficient design of COFbased architecture. 2. MODEL AND METHODS For this study, four types of 3D COFs were chosen for simulation, namely, COF-102, COF103, COF-105 and COF-1082. Simulated powder X-ray diffraction in comparison with experimental powder spectrum suggested a ctn topology for COF-102, 103 and 105, and bor (boracite, P 43m ) topology for COF-108.2 In this work, ctn topology was used for the four types of COFs. The crystal structures for COF-102 and COF-108 are illustrated in Fig.1 (a) and (b). If the tetrahedral C atom in COF-102 and COF-108 is replaced by a Si atom, the corresponding structures are of COF-103 and COF-105.

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

COF-102 and COF-103 can be synthesized by the self-condensation of tetra (4-dihydroxyborylphenyl) methane (TBPM) and tetra (4-dihydroxy-borylphenyl) silane (TBPS), respectively,

under formation of planar boroxine (B O ) rings.2 If we further perform the co-condensation of these compounds with triangular hexahydroxytriphenylene (HHTP), COF-105 and COF-108 will result. The optimized cell parameters in comparison with experimental values2 are given in Table 1. (a)

(b)

Fig.1 (a): crystal structure of COF-102; (b): crystal structure of COF-108; carbon, boron, oxygen and hydrogen atoms are represented in gray, red, green and white, respectively. If the tetrahedral C is replaced by Si, the structures are of COF-103 and COF-105, respectively. Table 1 Optimized cell parameters and experimental values of COFs with ctn (carbon nitride, I 43d

) topology. Structure

Building blocks

Cell parameters[Å] Simulation15

Experiment2

Density ρ [gcm-3]

COF-102

TBPM

27.144

27.1771(13)

0.43

COF-103

TBPS

28.353

28.2477(21)

0.39

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

COF-105

TBPS+HHTP

44.596

COF-108

TBPM+HHTP

43.387

Page 8 of 28

44.886(5)

0.18 0.19

* TBPM: tetra (4-dihydroxy-borylphenyl) methane; TBPS: tetra (4-dihydroxy-borylphenyl) silane; HHTP: triangular hexahydroxytriphenylene; In 2012, S. Amirjalayer et al.15 developed a force field for boron based COFs, derived from ab initio simulations. All the simulations in this work are based on this force field. All the details about the force field can be found in our Supporting Information. The MD simulation included two different techniques, equilibrium (EMD) and nonequilibrium (NEMD) molecular dynamics. NEMD requires the imposition of a steady-state temperature gradient along the axes of interest and the thermal conductivity is calculated using Fourier’s law of heat conduction. 27 NEMD requires a relatively larger sized simulation system, to build a reproducible diffusive temperature profile. EMD has significant advantages over NEMD in that it requires much smaller simulation cells to achieve converged results. Therefore, EMD was chosen to predict the thermal conductivity of COFs in the microcanonical ensemble. In the EMD method, small fluctuations of heat current are monitored over time and the Green–Kubo relation is used to determine the thermal conductivity of these perturbations in equilibrium.28 The thermal conductivity is written using the Green–Kubo relation:20,29 k = 







〈(0) ∙ ()〉 



Where ! is volume; "# is the Boltzmann constant; T is temperature;  is heat flux; and ⟨ ⟩

denotes time average. The heat flux  was calculated by

&  =  '∑* )* --- +, − ∑* ---- ./ --- +, 0



---- / is the symmetric Where )* is the total energy of atom 1 ; --- is +, the velocity of atom 1 ; and . stress tensor of atom 1 defined as

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

.33 ---- ./ = 2.34 .35

.34 .44 .45

.35 .45 6 .55



with @ C .78 = − 9 ∑AB& :;& "RSTU&  > "RSTU& W > "RSTU& . This result indicates that larger total overlap energy between all species of atoms results in higher thermal conductivity for the material. Larger total overlap energy implies a weaker mismatch while stronger phonon coupling exists in the material, which benefits heat transport47.

We focused on four sets of comparison, i.e. COF-102 vs. COF-103, COF-105 vs. COF-108, COF-102 vs. COF-108 and COF-103 vs. COF-105. Each set has a tiny structural difference, representing an exchange of only one tetrahedral Si atom and C atom in sets of COF-102 vs. COF-103 and COF-105 vs. COF-108. The latter had the HHTP rather than the former in sets of COF-102 vs. COF-108 and COF-103 vs. COF-105. When the tetrahedral C atom was replaced by

ACS Paragon Plus Environment

20

Page 21 of 28

a Si atom this mismatch of atoms increased and the phonon coupling was weakened. As a result, the thermal conductivities of COF-102/COF-108 were higher than those of COF-103/ COF-105. Similarly, compared with COF-102 /COF-103, energy localization worsens and phonon coupling was weakened in COF-108/COF-105 owing to the introduction of HHTP with additional thermal resistance. 0.06

C0 C1 C2 C3 C4 B O Overlap

0.00024

0.04 0.03

Overlap energy(ev)

(a)

0.05 VDOS(1/THz)

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

The Journal of Physical Chemistry

0.02 0.01

0.00020

COF-102

(b) COF-103 COF-108

0.00016 0.00012

COF-105

0.00008 0.00004

0.00 0

20

40 60 Frequency(THz)

80

0.00000 111

Fig.9 (a) Vibrational density of states for all species of atoms of COF-102 and their overlap at 300 K; (b) Overlap energy between all species of atoms of COF-102, COF-103, COF-105 and COF-108.

CONCLUSIONS In summary, the thermal conductivities of COF-102, COF-103, COF-105 and COF-108 were studied using equilibrium molecular dynamics with the Green-Kubo method. The relative contributions of different phonon oscillation frequencies to thermal conductivity, the phonon vibrational density of states (VDOS) and the overlap phonon energy were analyzed to estimate the heat transfer bottleneck and energy localization in the nanostructure. The main conclusions of this study are as follows: (1) The estimated thermal conductivity of four types of COFs is on the order of 0.1 W/(m⋅K) at

300K. And they were in the order of "RSTU& 9 > "RSTU&  > "RSTU& W > "RSTU& . As the 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

simulated system temperature was increased from 200 to 500 K, the thermal conductivities decrease between 26.6% and 47.0% for all four COFs. (2) It is deduced that the phonon mean free paths of four COFs are less than two unit cells, i.e. 2.7-9.2 nm, similar to some MOFs. This short mean free path for phonons results in the low thermal conductivity for the COFs. The heat conduction in the COFs can be primarily attributed to the low frequency phonon modes in the range of 0-50 THz. (3) Energy localization occurs on the C2 and C3 atoms in phenylene ring of the structure, as well as on C5, C6 and C7 atoms in HHTP. The C0-C1 or Si-C1 bonds act as bottlenecks to energy transport, followed by B-O atoms. In other words, the connection section between phenylene rings in COF-102 and COF-103, and the connection between phenylene and triphenylene in COF-105 and COF-108, are detrimental to the phonon coupling and energy flow. (4) The total overlap energy in the VDOS of all the species atoms was found to be the most serious mismatched structure in COF-105, resulting in its thermal conductivity being the lowest of four COFs examined. When the tetrahedral C atom was replaced by a Si atom, the atom mismatch increased, so that the thermal conductivities of COF-102/COF-108 were higher than COF-103/COF-105. Compared with COF-102/COF-103, the energy localization and phonon coupling was weaker in COF-108/COF-105 due to the introduction of HHTP with additional thermal resistance. This study is focused on 3D boron-based COFs, but the methodology employed can be extended to other COFs. To date, reports of the thermal properties of these complex nanoporous materials have been very scarce. More fundamental and in-depth studies will facilitate the design of COFs with desired thermal properties, which will enhance their use in practical applications in the marketplace.

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

ASSOCIATED CONTENT Supporting Information Validation of the simulation method, all the details about the force field in section 2 and overlap energy in section 3.6. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author * Yanhui Feng. E-mail: [email protected]. Tel: 86-010-62334971

ACKNOWLEDGEMENTS This work was supported by the National Natural Science Foundation of China (51422601 and 51436001), and the Fundamental Research Funds for the Central Universities (FRF-TP-15003C1).

REFERENCES (1)

Cote, A. P.; Benin, A. I.; Ockwig, N. W.; O’Keeffe, M.; Matzger, A. J.;Yaghi, O. M. Porous, Crystalline, Covalent Organic Frameworks. Science 2005, 310, 1166–1170.

(2) El-Kaderi, H. M.; Hunt, J. R.; Mendoza-Cortés, J. L.; Côté, A. P.; Taylor, R. E.; O’Keeffe, M.; Yaghi, O. M. Designed synthesis of 3D covalent organic frameworks. Science 2007, 316, 268–272. (3) Rowsell, J. L. C.; Yaghi, O. M. Metal–organic frameworks: a new class of porous materials. Microporous Mesoporous Mater. 2004, 73, 3–14.

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

(4) Klontzas, E.; Tylianakis, E.; Froudakis, G. E. Designing 3D COFs with enhanced hydrogen storage capacity. Nano Lett. 2010, 10,452–454. ( 5 ) Tylianakis, E.; Klontzas, E.; Froudakis, G. E. Multi-Scale Theoretical Investigation of Hydrogen Storage in Covalent Organic Frameworks. Nanoscale 2011, 3, 856– 869. (6)

Fang, Q.; Wang, J.; Gu, S.; Kaspar, R.B.; Zhuang, Z.; Zheng, J.; Guo, H.; Qiu, S.; Yan, Y.3D porous crystalline polyimide covalent organic frameworks for drug delivery. J. Am. Chem. Soc. 2015, 137, 8352–8355.

(7) Han, S. S.; Furukawa, H.; Yaghi, O. M.; Goddard III, W. A. Covalent Organic Frameworks as Exceptional Hydrogen Storage Materials.J. Am. Chem. Soc. 2008, 130, 11580–11581. (8) Garberoglio, G. Computer simulation of the adsorption of light gases in covalent organic frameworks. Langmuir 2007, 23, 12154–12158. (9) Choi, Y. J.; Choi, J. H.; Choi, K. M.; Kang, J. K. Covalent organic frameworks for extremely high reversible CO2 uptake capacity: a theoretical approach. J. Mater. Chem. 2011, 21, 1073–1078. (10) Feng, X.; Ding, X.; Jiang, D. Covalent organic frameworks. Chem. Soc. Rev. 2012, 41, 6010–6022. (11) Ding, S. Y.; Wang, W. Covalent organic frameworks (COFs): from design to applications. Chem. Soc. Rev. 2013, 42, 548–568. (12) Zhao, L.; Zhong, C. Negative Thermal Expansion in Covalent Organic Framework COF102. J. Phys. Chem. C 2009, 113, 16860–16862. (13) Han, S. S., and Goddard, W. A. Metal−Organic Frameworks Provide Large Negative Thermal Expansion Behavior. J. Phys. Chem. C 2007, 111, 15185–15191. (14) Lukose, B.; Kuc, A.; Heine, T. Stability and electronic properties of 3D covalent organic frameworks. J. Mol. Model. 2013, 19, 2143–2148. (15) Amirjalayer, S.; Snurr, R. Q.; Schmid, R. Prediction of structure and properties of boronbased covalent organic frameworks by a first-principles derived force field. J. Phys. Chem. C 2012, 116, 4921–4929. (16) Zhang, J.; Fisher, T. S.; Ramachandran, P. V.; Gore, J. P.; Mudawar, I. A Review of Heat Transfer Issues in Hydrogen Storage Technologies. J. Heat Transfer 2005, 127, 1391– 1399. (17)

Luo, T.; Chen, G. Nanoscale Heat Transfer—from Computation to Experiment. Phys. Chem. Chem. Phys. 2013, 15, 3389–3412.

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

(18) 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 Transfer 2007, 50, 393–404. (19) 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 Transfer 2007, 50, 405–411. (20) Zhang, X.; Jiang, J. Thermal conductivity of zeolitic imidazolate framework-8: a molecular simulation study. J. Phys. Chem. C 2013, 117, 18441–18447. (21) Qian, X.; Gu, X.; Yang, R. Anisotropic Thermal Transport in Organic–Inorganic Hybrid Crystal β-ZnTe(en)0.5. J. Phys. Chem. C 2015, 119, 28300–28308. (22) Slack, G. A. Thermal Conductivity of II-VI Compounds and Phonon Scattering by Fe2+ Impurities. Phys. Rev. B 1972, 6, 3791–3800. (23) Wang, H.; Chu, W. Thermal Conductivity of ZnTe Investigated by Molecular Dynamics. J. Alloys Compd. 2009, 485, 488–492. (24) 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, 26000–26008. (25) Purewal, J.; Liu, D.; Sudik, A.; Veenstra, M.; Yang, J.; Maurer, S.; Müller, U.; Siegel, D. J. Improved Hydrogen Storage and Thermal Conductivity in High-Density MOF-5 Composites. J. Phys. Chem. C 2012, 116, 20199– 20212. (26) Ming, Y.; Purewal, J.; Liu, D.a.; Sudik, A.; Xu, C.; Yang, J.; Veenstra, M.; Rhodes, K.; Soltis, R.; Warner, J.; Gaab, M.; Müller, U.; Siegel, D. J. Thermophysical properties of MOF-5 powders. Microporous Mesoporous Mater. 2014, 185, 235– 244. (27) Zhang, X.; Xie, H.; Hu, M.; Bao, H.; Yue, S.; Qin, G.; Su, G. Thermal Conductivity of Silicene Calculated Using an Optimized Stillinger-Weber Potential. Phys. Rev. B 2014, 89, 054310. (28) Salaway, R. N.; Zhigilei, L. V. Molecular Dynamics Simulations of Thermal Conductivity of Carbon Nanotubes: Resolving the Effects of Computational Parameters. Int. J. Heat Mass Transfer 2014, 70, 954–964. (29) Schelling, P. K.; Phillpot, S. R.; Keblinski, P. Comparison of atomic-level simulation methods for computing thermal conductivity. Phys. Rev. B 2002, 65, 144306. (30)

Plimpton, S. J. Fast parallel algorithms for short-range molecular dynamics. Journal of computational physics. J. Comput. Phys. 1995, 117, 1–19.

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

(31)

McGaughey, A. J. H.; Kaviany, M. Observation and description of phonon interactions in molecular dynamics simulations. Phys. Rev. B 2005, 71, 184305.

(32)

Cui, L.; Feng, Y.; Zhang, X. Enhancement of Heat Conduction in Carbon Nanotubes with Filled Fullerene Molecules. Phys. Chem. Chem. Phys. 2015, 17, 27520–27526.

(33)

Berber, S.; Kwon, Y.-K.; Tomanek, D. Unusually high thermal conductivity of carbon nanotubes. Phys. Rev. Lett. 2000, 84, 4613−4616.

(34)

Ziman, J. M. Principles of the Theory of Solids; Cambridge University Press: New York, 1972.

(35)

Zhang, H.; Lee, G.; Cho, K. Thermal transport in graphene and effects of vacancy defects. Physical Review B 2011, 84, 115460.

(36)

Chen, J.; Zhang, G.; Li, B. How to improve the accuracy of equilibrium molecular dynamics for computation of thermal conductivity? Phys. Lett. A 2010, 374, 2392– 2396.

(37) Roufosse, M.; Klemens, P. G. Thermal conductivity of complex dielectric crystals. Phys. Rev. B 1973, 7, 5379. (38)

Slack, G. A. The thermal conductivity of nonmetallic crystals. Solid State Phys. 1979, 34, 1–71.

(39)

Shiomi, J.; Esfarjani, K.; Chen, G. Thermal conductivity of half-Heusler compounds from first-principles calculations. Phys. Rev. B. 2011, 84, 104302.

(40)

Pisoni, A.; Jaćimović, J.; Barišić, O. S.; Spina, M.; Gaál, R.; Forró, L.; Horváth, E. J. Ultra-low thermal conductivity in organic–inorganic hybrid perovskite CH3NH3PbI3. J. Phys. Chem. Lett. 2014, 5, 2488–2492.

(41) Toberer, E. S.; May, A. F.; Snyder, G. Zintl chemistry for designing high efficiency thermoelectric materials†‡. J. Chem. Mater. 2010, 22, 624– 634. ( 42 )

Wan, C. L.; Wang, Y. F.; Wang, N.; Norimatsu, W.; Kusunoki, M.; Koumoto, K. Development of novel thermoelectric materials by reduction of lattice thermal conductivity. Sci. Technol. Adv. Mater. 2010, 11, 044306.

(43)

Loong, C.; Vashishta, P.; Kalia, R. K.; Jin, W.; Degani, M. H.; Hinks, D. G.; Price, D. L.; Jorgensen, J. D.; Dabrowski, B.; Mitchell, A. W. et al. Phonon density of states and oxygen-isotope effect in Ba1-xKxBiO3. Phys. Rev. B 1992, 45, 8052.

(44)

Guo, R.; Huang, B. Thermal Transport in Nanoporous Si: Anisotropy and Junction Effects Int. J. Heat Mass Transfer 2014, 77, 131–139.

(45)

Pop, E.; Varshney, V.; Roy, A. K. Thermal properties of graphene: Fundamentals and applications. MRS Bull. 2012, 37, 1273–1281.

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

(46)

McGaughey, A. J. H.; Kaviany, M. Thermal conductivity decomposition and analysis using molecular dynamics simulations. Part I. Lennard-Jones argon. Int. J. Heat Mass Transfer 2004, 47, 1783–1798.

(47) Cui, L.; Feng Y. H.; Zhang, X. X. Dependence of Thermal Conductivity of Carbon Nanopeapods on Filling Ratios of Fullerene Molecules. J. Phys. Chem. A 2015, 119, 11226–11232.

Table of Contents (TOC) Image The TOC graphic in the actual size was submitted as an attachment in File Upload.

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

TOC graphic 46x36mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 28 of 28