Thermal Conductivity of Covalent Organic Frameworks as a Function

Nov 15, 2017 - Single-Site Photocatalytic H2 Evolution from Covalent Organic Frameworks with Molecular Cobaloxime Co-Catalysts. Journal of the America...
0 downloads 10 Views 883KB Size
Subscriber access provided by READING UNIV

Article

Thermal Conductivity of Covalent Organic Frameworks as a Function of Their Pore Size Sunny K S Freitas, Raquel Silveira Borges, Claudia Merlini, Guilherme M.O. Barra, and Pierre Mothé Esteves J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b10487 • Publication Date (Web): 15 Nov 2017 Downloaded from http://pubs.acs.org on November 16, 2017

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 7

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 Covalent Organic Frameworks as a Function of Their Pore Size Sunny K. S. Freitasa; Raquel S. Borgesb; Claudia Merlinic, Guilherme M. O. Barrad, Pierre M. Estevesa,* a.

Instituto de Química - Universidade Federal do Rio de Janeiro - Cidade Universitária - Ilha do Fundão - Centro de Tecnologia - Bloco A - Rio de Janeiro – RJ, 21949-900 Brazil

b.

FMC Technologies, Rio de Janeiro, Brazil

c.

Engineering Department, Federal University of Santa Catarina, Blumenau, SC, Brazil

d.

Mechanic Engineering Department, Federal University of Santa Catarina, Admar Gonzaga Avenue, 1346, 88034001, Florianópolis, SC, Brazil email: [email protected] ABSTRACT: Covalent Organic Frameworks (COFs) are a relatively new class of organic nanostructured porous materials, formed by covalent bonds between light elements, with high surface area. The thermal conductivity values (k) for some COFs (COF-300, RIO-1, RIO-4, RIO-20) were measured using the modified transient plane source (MTPS) technique. Values ranging from 0.038 to 0.048 Wm-1K-1 were measured depending on the COF pore structure and surface area. Thermal conductivities correlate linearly with the inverse of the cross sectional area of the pores.

Introduction The knowledge of thermal properties of materials is fundamental for energy conservation, therefore, for sustainability.1,2 The design of good thermal insulation, for example, can be used to prevent heat from being exchanged between the rooms of a house and its exterior, saving energy. Typical good thermal insulators often are porous, and this leads to low mechanical strength and little flexibility, which may limit their use. Materials with nanometric pores have been considered as good candidates for thermal insulation. For example, aerogels have very low values of thermal conductivity, k, (0.010 W m-1K-1),3,4 since more than 95% of their volume consists of empty spaces.5,6 Other classes of porous solids have been developed and include zeolites and, more recently, metalorganic frameworks (MOFs).7–11 Zeolites are porous aluminosilicates,12,13 normally used in adsorption, ion exchange, catalysis, etc. The thermal conductivity of some zeolites are around 0.030 Wm-1K-1 for powder samples.14,15 Metal-organic frameworks (MOFs) are porous coordination polymers, formed by the union of metallic agglomerates and organic linkers. There are only a few examples of their thermal properties, for instance the thermal conductivity of a single crystal of MOF-5 was measured to be 0.32 Wm-1K-1.16,17 Recently an emerging class of micro- and mesoporous materials, called Covalent Organic Frameworks (COFs), has been proposed18–26 These ordered structures, formed

through condensation reactions between organic molecules, have high chemical stability and high porosity. Depending on their topology and functionality, these reticular materials are promising candidates for various applications, such as gas adsorption27, catalysis28–31, proton conduction32, energy storage27,33,34, and optoelectronics.35,36 Therefore, knowing their thermophysical properties would improve these and other applications. The thermal conductivity (k) is an important parameter in this context. In porous materials, heat is propagated by thermal conductance through the solid, and both radiation and convection through the pores. Therefore, a dependence of k is expected as a function of the pore size, and possibly with the nature of the material. Considering that COFs have low densities (0.17 g / cm3 for COF-108)22 and many voids, these porous materials must therefore conduct very little heat i.e. should have a very low thermal conductivity. Its cage-like structures, potentially interlaced, forms bridges that connect the pores, can define or delimit the amount of energy that can be transmitted through the structure in the microscopic scale. These characteristics also make the COFs and their hierarchical materials excellent candidates as materials used for thermal insulation. As they are a relatively new class of material, no data with respect to their thermal conductivity, to our knowledge, has been reported. In this contribution we present the measurements of the thermal conductivity of some of these materials.

ACS Paragon Plus Environment

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 7

2 Methods A series of organic, bi- and tri-dimensional porous materials of high porosity and stability, with high specific surface areas and different pore sizes, called RIOs (acronym for Reticular Innovative Organic materials), were synthesized. RIO-1, for example, contains a tridimensional channel system, and is built from tetrahedral tetrapodal (tetrakis-(4-aminophenyl)methane) and triangular and tripodal (triformylphloroglucinol) building blocks; RIO-4 is a related porous material formed by the tetrakis(4-aminophenyl)-methane and 1,2,3triformylphenol blocks, while COF-300 is prepared from the same tetrahedral building block with the linear dipodal terephtalaldehyde (Figure 1).

tained by single crystal X-ray diffraction37, and its 7.8 Å pores. Characterization of the samples was carried out from the physical adsorption of N2 at 77K. The pore size distribution for the materials was calculated from the isotherms using the Non-local Density Functional Theory (NLDFT) method on carbon and the slit/cylindrical pore model, as implemented in the software NovaWin version 10.01 of Quantachrome™ Nova 2200e. The surface area was determined by the Brunauer-Emmett-Teller (BET) method. Depending on the material, adsorption or desorption data were used in the range of 0.01-0.23 P/P0. Micropore volume was determined using the V-t method (t-plot, de Boer) on the N2 isotherms in the region 0.2 < P/Po < 0.5. The pre-treatment details are described in the in Supporting Information.

Figure 2: Structure of COF-300, showing its 7.8 Å pores.

Figure 1: Combination between the building-blocks for preparing COFs.

RIO-1, RIO-4 and COF-300 (Figure 1) were synthesized under solvothermic conditions in mixture of 1,4dioxane and 3M HOAc, in a sealed vessel, at 120°C, adapting the procedure reported by Uribe-Romo et al.37 RIO-1 was prepared in two slightly different ways: RIO1a was prepared under stirring and RIO-1b without stirring. RIO-20 was synthesized under different conditions, using a round bottom flask, coupled to a condenser, at 190°C. The Supporting Information section contains detailed procedures of these syntheses. Figure 2 shows the tridimensional structure of COF-300, ob-

The thermal conductivity k of the powdered samples were measured based on the modified transient plane source (MTPS) technique through the TCi C-therm™ conductivity meter, which uses a unilateral, interfacial heat reflectance sensor that applies a momentary constant heat source in the sample (see Supporting information for details). Shortly, a known current is applied to the heating element of the sensor, generating a small amount of heat. The heat generated results in an increase in the temperature at the interface between the sensor and the sample - typically less than 2 °C. This temperature rise at the interface induces a change in the voltage drop of the sensor element. The rate of increase of voltage in the sensor is used to determine the thermophysical properties of the sample material. The analyses were carried out at 25 oC and reported values were the average of 10 measurements for each COF (see Supporting Information for details). The results obtained for the thermal conductivity were related to the textural properties of the materials.

ACS Paragon Plus Environment

Page 3 of 7

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

3

Figure 3: N2 gas adsorption/desorption isotherms for COFs and their respective pore distribution. (a) RIO-1a, (b) RIO-1b, (c) RIO-4, (d) COF-300.

Results and Discussion Figure 03 shows that type I isotherms have been obtained for these materials indicating the dominant presence of micropores. RIO-1a and RIO-1b show slightly different isotherms and their surface area (BET) showed considerable variation (1320 m²/g for RIO-1a and 820 m²/g for RIO-1b). However, similar values were obtained for pore width (d), determined as 14 Å for both RIO-1a and for RIO-1b. RIO-4 also presents a type I isotherm (micropores), with a BET surface area of 1080 m²/g and a pore diameter of 11.2 Å. For both RIO-4 and RIO-1a and RIO-1b, the pore distribution is monomodal. COF-300, is crystalline, and also shows type I isotherm, with BET area of 1440 m²/g and a pore width of 7.8 Å (from single crystal diffraction).37 The powder X-ray diffraction (XRD) for RIO-1a, RIO-1b and RIO-4 indicates low crystallinity of these materials (Supporting Information, Figures S5-S8). The XRD diffratogram of RIO-1b, synthesized without stirring, showed slightly more defined diffraction peaks in

comparison to the ones obtained for RIO-1a. The synthesis without stirring seems to favor the formation of larger crystals. Table 1 shows the thermal conductivity (k) values for these materials. The measured values are as low as those of typical thermal insulators used in industry, civil engineering, among other applications, typically in the range of 0.04 W m-1K-1.4,38 We have also measured the thermal conductivity of some zeolites, in order to compare the thermal properties of such well-known microporous materials with the ones of COFs. Figure 4a shows that, as the pore diameter increases, there is a steady decrease of k. This can be understood as the pore diameter increases, the material has more voids within its pore structures resulting in a conductivity tending to the conductivity of air, which is lower than bulky materials. Note that, among these COFs, COF-300 has the highest crystallinity (Fig. S5), while RIO-1, RIO-4 and RIO-20 have lower crystallinities (Figs. S6-S8). COF-300 showed slightly higher thermal conductivity, 0.048 ± 0.001 Wm-1K-1, which may be due to its catenated wall structures, that

ACS Paragon Plus Environment

The Journal of Physical Chemistry

Page 4 of 7

4 increase the number of routes for the heat being transmitted through phonons in the entangled framework. In other words, the modes of vibration within the crystal network contribute to phonon transport (mainly acoustic). Table 1: Thermal conductivities of COFs and zeolites Thermal conductivity, k

(Å)

(Wm K )

-1

COFs COF-300

1444 [1366]

37

[7.8]

37

0.048 ± 0.001

RIO-1a

1320

14

0.039 ± 0.001

RIO-1b

820

14

0.0395 ± 0.0006

RIO-4

1080

11.2

0.043 ± 0.001

RIO-20

0.0395 ± 0.0005

Zeolites ZSM-5 (MFI)

39

4.7

0.065 ± 0.001

40

5.95

0.058 ± 0.001

41

4.21

[0.08]

[293 ]

Beta (BEA)

[680]

4A (LTA)

[421]

(a)

14

The interesting point that arises from the experiments is a verification of a linear dependence of the thermal conductivity with the inverse of the pore cross sectional area, i.e. 1/(πr2), where r = (pore width)/2 (Figure 4b). In these porous materials, the heat is transferred mainly through the lattice vibration modes, also known as phonons, which are proportional to the number of chain crossings per unit area. This seems to be true even for zeolites, which are reference microporous materials, since their thermal conductivities also correlates with the inverse of their pore cross sectional area. Babaei, McGaughey and Wilmer42 recently reported a theoretical study on the thermal properties of metalorganic frameworks (MOFs), based on molecular dynamics calculations, and also predicted that these related microporous materials should also present such trend. We have experimentally verified their theoretical predictions based upon MOFs with the COFs of the present study. With this linear correlation in mind, an interesting consequence is to use the thermal conductivity of a sample for estimating the unknown/undetermined pore size of a microporous material. Usually, pore size determination is performed from the gas adsorption isotherm data, using several approaches, such as density functional theory (DFT) or Barret, Joyner and Haland method (BJH)43,44, or by diffraction, either by electrons or X-rays, in the case of single crystals. The first approach requires experiments that require long times and has limitations for precisely determining micropores, in which the quadrupole momentum of the probe molecule such as N2 usually affords problems. On the other hand, transmission electron mi-

0.08 0.07 -1

-1

-1

(m /g) [lit.]

Pore width

k (W m K )

2

0.06 0.05 0.04 4

6

8

10

12

14

16

Pore width (d) (Angs) (b) 0.09 LTA

0.08

-1

Surface area

0.07 ZSM-5

-1

Sample

croscopy or X-ray diffraction are expensive techniques that require crystalline materials for analysis, which is often not the case for COFs and related materials. RIO-20, for instance, was prepared from the condensation of melamine and terephtaladehyde and gave a material of low crystallinity and of difficult characterization. Its thermal conductivity however was easily determined and was found to be (0.0395 ± 0.0005) W m-1K-1. From the correlation we can determine, from its thermal conductivity that the pore width of RIO-20 would be 14.4 Å. We eventually have determined the pore size for such material from its N2 adsorption isotherm, using DFT treatment, and we found, that its pore size to be 15.2 Å, which is close to the one obtained by the correlation with thermal conductivity. Thus, this correlation can be useful for estimating pore sizes of such materials. Since thermal conductivity is a non-destructive and relatively fast and easy method, it can be used for screening purposes of such a structural property.

k (W m K )

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

BEA

0.06

y = 0.57*x + 0.03601 Adj. R-Square = 0.97771

0.05 RIO4

0.04

COF-300

RIO1

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 2

-2

1/(π.r ) (Angs )

Figure 4: (a) The effect of the pore size of RIO-1, RIO-4, and COF-300 with their values of k. Linear dependence in (b) of the conductivity with the inverse of the pore cross sectional area.

ACS Paragon Plus Environment

Page 5 of 7

The Journal of Physical Chemistry

5 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 thermal conductivities obtained for the materials studied here may be related to several factors. In relation to the porosity, it is noted that the larger the pores present in the COFs, the less heat will be transferred through the material. Pores are constituted mainly of empty spaces, which can be filled by gas or liquid flows (e.g. air). Therefore, the heat conduction in this case is minimal, basically made through the phonons conducted through the walls of the material, since they are not metallic structures. There may also be a contribution from the internal convection effect, which may occur due to the presence of gas adsorbed on the pore walls. It is noteworthy that neither surface area nor other textural characteristics of such materials is correlated in a simple way to their thermal conductivities. However, it may be possible that other structural parameters of these materials, such as crystallite size, dimension order of the channels (i.e. 1D, 2D or 3D-channel system), entanglement degree, etc. are related to their thermophysical properties. However, for the cases investigated in this study, we could only find a relation of the thermal conductivities with the inverse of the pore sectional area and that the (powdered) COFs present relatively low thermal conductivities, on the order of 10-2 W m-1K-1, compatible with good insulators.

REFERENCES (1)

(2) (3)

(4)

(5) (6)

(7)

(8) (9)

(10)

Conclusions In this study we have reported the thermal conductivity values for a series of COFs, namely RIO-1a, RIO-1b, RIO-4, COF-300 and RIO-20 which are respectively 0.0389 (a) e 0.0395 (b), 0.043, 0.048 and 0.0395 W m-1K-1. Correlations between the thermal conductivity values with some structural parameters were investigated, such as the pore diameter. An interesting correlation of the thermal conductivity with the pore cross sectional area, considering cylindrical pores, was found. This appears as a useful, nondestructive and easy method for estimating the pore width for such and related materials. In conclusion, the pore size was the main structural variable that played a major role in the thermal conductivity.

(11)

(12)

(13)

(14)

(15)

(16)

Acknowledgements This work is dedicated to the memory of Prof. George A. Olah. This work was partially financially supported by CNPq, FAPERJ and CAPES. Prof. Simon J. Garden is acknowledged for helpful discussion.

Supporting Information

(17)

(18)

(19)

Synthetic details, characterization of materials, thermal conductivities, experimental details. (20)

(21)

Bentz, D.; Peltz, M.; Durán-Herrera, A.; Valdez, P.; Juárez, C. Thermal Properties of High-Volume Fly Ash Mortars and Concretes. J. Build. Phys. 2011, 34, 263–275. Berge, A.; Johansson, P. Ä. R. Literature Review of High Performance Thermal Insulation. Build. Phys. 2012, 40. Fricke, J.; Tillotson, T. Aerogels: Production, Characterization, and Applications. Thin Solid Films 1997, 297, 212–223. Cuce, E.; Cuce, P. M.; Wood, C. J.; Riffat, S. B. Toward Aerogel Based Thermal Superinsulation in Buildings: A Comprehensive Review. Renew. Sustain. Energy Rev. 2014, 34, 273–299. Kistler, S. S. Coherent Expanded-Aerogels. J. Phys. Chem. 1932, 36, 52–64. Nguyen, S. T.; Feng, J.; Ng, S. K.; Wong, J. P. W.; Tan, V. B. C.; Duong, H. M. Advanced Thermal Insulation and Absorption Properties of Recycled Cellulose Aerogels. Colloids Surfaces A Physicochem. Eng. Asp. 2014, 445, 128– 134. Rowsell, J. L. C.; Yaghi, O. M. Metal-Organic Frameworks: A New Class of Porous Materials. Microporous Mesoporous Mater. 2004, 73, 3–14. Zhou, H.-C.; Kitagawa, S. Metal–Organic Frameworks (MOFs). Chem. Soc. Rev. 2014, 43, 5415–5418. Chen, B. Interwoven Metal-Organic Framework on a Periodic Minimal Surface with Extra-Large Pores. Science 2001, 291, 1021–1023. Tranchemontagne, D. J.; Mendoza-Cortés, J. L.; O’Keeffe, M.; Yaghi, O. M. Secondary Building Units, Nets and Bonding in the Chemistry of Metal–Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1257-1283. Cheetham, A. K.; Férey, G.; Loiseau, T. Open-Framework Inorganic Materials. Angew. Chemie Int. Ed. 1999, 38, 3268– 3292. Li, Y.; Yu, J. New Stories of Zeolite Structures: Their Descriptions, Determinations, Predictions, and Evaluations. Chem. Rev. 2014, 114, 7268–7316. Rollmann, L. D.; Valyocsik, E. W.; Shannon, R. D. Zeolite Molecular Sieves. In Inorganic Syntheses; John Wiley & Sons, Inc.: Hoboken, NJ, USA; pp 227–234. Murashov, V. V.; White, M. A. Thermal Properties of Zeolites: Effective Thermal Conductivity of Dehydrated Powdered Zeolite 4A. Mater. Chem. Phys. 2002, 75, 178–180. Schnell, S. K.; Vlugt, T. J. H. Thermal Conductivity in Zeolites Studied by Non-Equilibrium Molecular Dynamics Simulations. Int. J. Thermophys. 2013, 34, 1197–1213. 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, 405–411. 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, 393–404. Diercks, C. S.; Yaghi, O. M. The Atom, the Molecule, and the Covalent Organic Framework. Science 2017, 355, eaal1585. Waller, P. J.; Gándara, F.; Yaghi, O. M. Chemistry of Covalent Organic Frameworks. Acc. Chem. Res. 2015, 48, 3053–3063. Ockwig, N. W.; Cote, A. P.; Keeffe, M. O.; Matzger, A. J.; Yaghi, O. M. Porous , Crystalline , Covalent Organic Frameworks. Science 2005, 310, 1166–1171. Calik, M.; Sick, T.; Dogru, M.; Döblinger, M.; Datz, S.; Budde, H.; Hartschuh, A.; Auras, F.; Bein, T. From Highly

ACS Paragon Plus Environment

The Journal of Physical Chemistry

Page 6 of 7

6 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

(22)

(23)

(24) (25)

(26)

(27)

(28)

(29)

(30)

(31)

(32)

(33)

(34)

(35)

(36)

(37)

(38)

(39)

Crystalline to Outer Surface-Functionalized Covalent Organic Frameworks – A Modulation Approach. J. Am. Chem. Soc. 2016, 138, 1234–1239. 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. Jackson, K. T.; Reich, T. E.; El-Kaderi, H. M. Targeted Synthesis of a Porous Borazine-Linked Covalent Organic Framework. Chem. Commun. 2012, 48, 8823–8825. Jiang, J.; Zhao, Y.; Yaghi, O. M. Covalent Chemistry Beyond Molecules. J. Am. Chem. Soc. 2016, 138, 3255-3265. Jin, Y.; Zhu, Y.; Zhang, W. Development of Organic Porous Materials Through Schiff-Base Chemistry. CrystEngComm 2013, 15, 1484–1499. Zhang, Y.; Su, J.; Furukawa, H.; Yun, Y.; Ga, F.; Duong, A.; Zou, X.; Yaghi, O. M. Single-Crystal Structure of a Covalent Organic Framework. 2013, 135, 16336-16339. Furukawa, H.; Yaghi, O. M. Storage of Hydrogen, Methane, and Carbon Dioxide in Highly Porous Covalent Organic Frameworks for Clean Energy Applications. J. Am. Chem. Soc. 2009, 131, 8875–8883. Gonçalves, R. S. B.; de Oliveira, A. B. V.; Sindra, H. C.; Archanjo, B. S.; Mendoza, M. E.; Carneiro, L. S. A.; Buarque, C. D.; Esteves, P. M. Heterogeneous Catalysis by Covalent Organic Frameworks (COF): Pd(OAc) 2 @COF-300 in Cross-Coupling Reactions. ChemCatChem 2016, 8, 743–750. Ding, S. Y.; Gao, J.; Wang, Q.; Zhang, Y.; Song, W. G.; Su, C. Y.; Wang, W. Construction of Covalent Organic Framework for Catalysis: Pd/COF-LZU1 in Suzuki-Miyaura Coupling Reaction. J. Am. Chem. Soc. 2011, 133, 19816–19822. Kaleeswaran, D.; Antony, R.; Sharma, A.; Malani, A.; Murugavel, R. Catalysis and CO₂ Capture by Palladium Incorporated Covalent Organic Frameworks. Chempluschem 2017, 82, 1253–1265. Mu, M.; Wang, Y.; Qin, Y.; Yan, X.; Li, Y.; Chen, L. TwoDimensional Imine-Linked Covalent Organic Frameworks as a Platform for Selective Oxidation of Olefins. ACS Appl. Mater. Interfaces 2017, 9, 22856–22863. Chandra, S.; Kundu, T.; Kandambeth, S.; BabaRao, R.; Marathe, Y.; Kunjir, S. M.; Banerjee, R. Phosphoric Acid Loaded Azo (− N=N−) Based Covalent Organic Framework for Proton Conduction. J. Am. Chem. Soc. 2014, 136, 6570– 6573. Han, S. S.; Mendoza-Cortés, J. L.; Goddard III, W. A. Recent Advances on Simulation and Theory of Hydrogen Storage in Metal–Organic Frameworks and Covalent Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1460-1476. DeBlase, C. R.; Silberstein, K. E.; Truong, T.-T.; Abruña, H. D.; Dichtel, W. R. B-Ketoenamine-Linked Covalent Organic Frameworks Capable of Pseudocapacitive Energy Storage. J. Am. Chem. Soc. 2013, 135, 16821-16824. Bertrand, G. H. V.; Michaelis, V. K.; Ong, T.-C.; Griffin, R. G.; Dinca, M. Thiophene-Based Covalent Organic Frameworks. Proc. Natl. Acad. Sci. 2013, 110, 4923–4928. Cai, S.-L.; Zhang, Y.-B.; Pun, A. B.; He, B.; Yang, J.; Toma, F. M.; Sharp, I. D.; Yaghi, O. M.; Fan, J.; Zheng, S.-R.; et al. Tunable Electrical Conductivity in Oriented Thin Films of Tetrathiafulvalene-Based Covalent Organic Framework. Chem. Sci. 2014, 5, 4693–4700. Uribe-romo, F. J.; Hunt, J. R.; Furukawa, H.; Keeffe, M. O.; Yaghi, O. M. A Crystalline Imine-Linked 3-D Porous Covalent Organic Framework. J. Am. Chem. Soc. 2009, 131, 4570–4571. Jelle, B. P.; Gustavsen, A.; Baetens, R. The Path to the High Performance Thermal Building Insulation Materials and Solutions of Tomorrow. J. Build. Phys. 2010, 34, 99-123. Sang, S.; Chang, F.; Liu, Z.; He, C.; He, Y.; Xu, L. Difference

(40)

(41)

(42)

(43)

(44)

of ZSM-5 Zeolites Synthesized with Various Templates. Catal. Today 2004, 93–95, 729–734. Shanjiao, K.; Yanjun, G.; Tao, D.; Ying, Z.; Yanying, Z. Preparation and Characterization of Zeolite Beta with Low SiO2/Al2O3 Ratio. Pet. Sci. 2007, 4, 70–74. Esmaeili, N.; Kazemian, H.; Bastani, D. Controlled Crystallization of LTA Zeolitic Nanoparticles from a Clear Solution Using Organic Template. Iran. J. Chem. Chem. Eng. 2011, 30, 9–14. Babaei, H.; McGaughey, A. J. H.; Wilmer, C. E. Effect of Pore Size and Shape on the Thermal Conductivity of MetalOrganic Frameworks. Chem. Sci. 2017, 8, 583–589. Rouquerol, F.; Rouquerol, J.; Sing, K. Adsorption by Powders & Porous Solids: Principles, Methodology and Applications.; Provence, 1999; Vol. 1. Thommes, M.; Kaneko, K.; Neimark, A. V; Olivier, J. P.; Rodriguez-reinoso, F.; Rouquerol, J.; Sing, K. S. W. Physisorption of Gases , with Special Reference to the Evaluation of Surface Area and Pore Size Distribution (IUPAC Technical Report). 2015, 87, 1051–1069.

ACS Paragon Plus Environment

7 Page 7 of 7

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

TOC Graphic

ACS Paragon Plus Environment

7