Target Synthesis of an Azo (N N) Based Covalent Organic Framework

Apr 15, 2016 - An azo (N═N) based covalent organic framework (COF-TpAzo) has been synthesized via a Schiff base condensation reaction...
2 downloads 0 Views 1MB Size
Article pubs.acs.org/jced

Target Synthesis of an Azo (NN) Based Covalent Organic Framework with High CO2‑over‑N2 Selectivity and Benign Gas Storage Capability Rile Ge, Dandan Hao, Qi Shi, Bin Dong, Wenguang Leng, Chang Wang, and Yanan Gao* Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian, 116023, P. R. China S Supporting Information *

ABSTRACT: An azo (NN) based covalent organic framework (COF-TpAzo) has been synthesized via a Schiff base condensation reaction. The Braunauer−Emmett−Teller (BET) specific surface area up to 1552 m2·g−1 was obtained for the framework with a pore volume of 0.97 cm3·g−1. The COF-TpAzo exhibits a high selectivity in CO2/N2 (127/145 at 273/298 K, respectively, Henry method) due to the N2-phobic and CO2-philic feature of COF-TpAzo. Furthermore, hydrogen and methane storage capacities of the nitrogen-rich COFTpAzo have been investigated. The COF-TpAzo shows high H2 and CH4 storage capacity at 1 bar (10.6 mg·g−1 at 77 K for H2, 11.2 mg·g−1 at 273 K for CH4, and 5.5 mg·g−1 at 298 K for CH4).



INTRODUCTION Covalent organic frameworks (COFs) are a novel class of highly porous organic microcrystalline materials that are constructed from lightweight elements with well-defined pore sizes that makes them extremely low dense and high specific surface area.1−7 More intriguingly, the building units can be precisely integrated into periodic two- or three-dimensional structures, endowing COFs with excellent flexibility in the design of frameworks with distinct structures and properties. Thus, COFs have been considered as promising materials for gas storage (CO2, CH4, H2, N2, etc.),8−12 optoelectricity,13 and catalysis5,7,14 since they emerged. CO2 capture is an integral part of the energy industry. Although various technologies have been developed by far, these technologies remain inefficient, leading to costly energy penalties. For instance, aqueous alkanolamine solutions have been widely used to adsorb CO2, but strong interaction of the base with CO2 resulted in the difficulty in separation of CO2, which requires a high energy input for cleavage of this interaction. Porous sorbents are of particular interest as they could potentially reduce the cost of CO2 capture from flue gas.15 Zeolites and activated carbons have been used to capture CO2, but the density of CO2 is low in these materials, so that adsorbent bed can not be minimized. CO2 and N2 have similar and small kinetic diameters, 3.30 and 3.64 Å, respectively.16 However, their different properties (quadrupole moments and polarizabilities) make it possible to design CO2-philic materials by introducing strong basic sites such as triazine,17 tetrazole,18 imide,19 and amines20 into the skeleton of materials. Due to the organic nature of the skeleton, COFs can be modified easily for a special purpose. For this reason, the first gas selectivity of COFs was published by Jiang et al., in 2011.21 Their results © XXXX American Chemical Society

indicated that the pore surface group was an important factor that affects the sorption properties of the COFs. The highest selectivity of CO2/N2 reported by them was 16.2 obtained by 100%AcTrz-COF-5, which is about 16-fold higher than unmodified COF-5. However, the low chemical stability of B−O-bond-based COFs limited their effective use in gas separation under practical conditions. A type of CN bond (from Schiff base reaction) COF has been developed to overcome the drawback of instability. An azine-linked COF (ACOF-1) has been published in the gas separation field and high selectivities toward CO2/N2 (40) and CO2/CH4 (37) were obtained at 273 K.12 Lately, the highest CO2 selectivity toward N2 for all porous organic polymers was published by Lotsch et al.22 The selectivity (189 and 502) toward CO2/N2 of the nitrogen rich covalent triazine framework was obtained by Henry method and IAST method, respectively. In 2013, Banerjee and co-workers reported ultra stable COFTpPa-1 and COF-TpPa-2 synthesized by condensation of 1,3,5triformylphloroglucinol (Tp) with p-phenylenediamine (Pa-1) and 2,5-dimethyl-p-phenylenediamine (Pa-2), respectively, with extremely high resistance to acid (9 N HCl), base (9 N NaOH), and boiling water due to the irreversible transform of hydroxyl (−OH) from enol to keto during the reaction.4 Later, they reported an azo (NN) functionalized COF-TpAzo, which was also stable under extreme conditions and exhibited high proton conductivities in the presence of phosphoric acid.23 The same COF-TpAzo was also successfully prepared by our group and a different solvothermal synthesis system was Received: January 25, 2016 Accepted: April 11, 2016

A

DOI: 10.1021/acs.jced.6b00071 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

employed to optimize the crystal structure and surface area. Different from the proton conduction behavior reported by Banerjee et al., our original intention to make the COF-TpAzo is to introduce the N2-phobic and CO2-philic azo groups24,25 into the framework of the COF for CO2/N2 separation. As expected, the COF-TpAzo exhibited high CO 2 -over-N 2 selectivity. This unique property of COF-TpAzo, along with their thermal and chemical stability, makes it ideal candidate for postcombustion CO2 separations.



EXPERIMENTAL SECTION

Materials and Method. 4,4′-Azodianiline (Azo) and 1,3,5triformylphloroglucinol (Tp) was synthesized according to the reported procedures.26,27 The 1H NMR spectra matched well with those reported results. 4-Aminoacetanilide, anhydrous 1,3,5-trihydroxybenzene, and hexamethylenetetramine was purchased from Alfa Aesar and used as received. Other organic solvents for synthesis were distilled over appropriate drying reagents under nitrogen. Powder X-ray diffraction (PXRD) data were collected with a PANalytical X’Pert Pro Diffractometer operated at 40 kV and 40 mA with Cu Kα radiation (step size, 0.017°; step time, 10.34 s). FT-IR spectra were recorded with a Bruker TENSOR 27 instrument in the region of 400−4000 cm−1. Gas sorption experiments were collected using a Quantachrome AutosorbiQ2 analyzer using adsorbates of UHP grade. Before measurement, the samples were degassed using high-vacuum at 120 °C for 10 h. Brookfield TC-520 temperature controller was used to maintain temperatures of gas sorption experiments. Thermogravimetric analyses were performed on a STA 449 F3 analyzer (Netzsch) under a N2 atmosphere at a heating rate of 10 °C· min−1 with a temperature range of 40−1000 °C. Scanning electron microscopy (SEM) micrographs were taken on a FEI Quanta 200F operating at an accelerating voltage of 20 kV. Preparation of COF-TpAzo. The azo-based COF-TpAzo was synthesized by the condensation reaction between Tp (0.3 mmol, 63.0 mg) and Azo (0.45 mmol, 95.4 mg) in 1,4-dioxane (3 mL). The mixture was sonicated for 10 min. The tube was then flash-frozen at 77 K (liquid N2 bath) and degassed by three freeze−pump−thaw cycles. The tube was sealed and then heated at 120 °C for 3 days. After the mixture was cooled to room temperature, a red precipitate was collected by filtration. The COF was soaked in tetrahydrofuran for 4 h and then washed with tetrahydrofuran (20 mL × 3) and acetone (20 mL × 3) until the filtrated was colorless. Then the powders were dried under vacuum at 100 °C overnight to afford 125 mg (88% yield) of COF-TpAzo. The reaction solvent used in this synthesis system is different from that report by Banerjee group,23 where 1,4-dioxane was used instead of a mixture of dimethylacetamide and o-dichlorobenzene (1:1). The result indicated that COF-TpAzo can be easily synthesized under various conditions. The schematic representation of simulated eclipsed patterns and the morphology of the COF-TpAzo were shown in Figure 1A and B, respectively.

Figure 1. Schematic representation of (A) simulated eclipsed patterns and (B) SEM image of COF-TpAzo.

angle of 3.12° (100). Several less intense peaks at 5.40°, 6.24°, and 8.26° could be assigned to (110), (200), and (210) facets, respectively. A slightly broad peak at higher 2θ (∼27.1°) is due to the π−π stacking between the COF layers and corresponds to the (001) plane. The d spacing for COF-TpAzo was estimated to be 3.31 Å, a little smaller than that of the previously reported 2D-COF,4,14,28 which indicates that there exists a stronger interaction between the adjacent layers. Furthermore, the Pawley refinement yielded an XRD pattern (Figure 2A, pink curve) that is in good agreement with the experimentally observed pattern, as evidenced by their negligible difference (Figure 2A, black curve). A monoclinic unit cell (P6) with the parameters of a = b = 32.8333 ± 0.06 Å, c = 3.4271 ± 0.005 Å, α = β = 90°, and γ = 120° were deduced. This is in accordance with the calculation result reported by Banerjee et al.23 The Rwp and Rp values were converged to 2.22% and 1.69%, respectively. The absence of carbonyl stretching band (1639 cm−1) of Tp indicates the total consumption of starting materials. The strong peak at 1623 cm−1 corresponds to the characteristic CN stretching vibrations of imines. The FTIR spectroscopy of COF-TpAzo matches well with that of the literture.23 Thermogravimetric analysis (TGA) of COF-TpAzo was performed to determine the thermal stability up to 350 °C. (Figure 2C) The porosity of the COF-TpAzo was established by Quantachrome Autosorb-iQ2 N2 sorption−desorption measurements. It seems from the isotherm that COF-TpAzo exhibited a combination of type I and type IV sorption curves (Figure 3A). The pore size distribution was calculated by using nonlocal density function theory (NLDFT), and it can be found that a narrow size distribution of 25.8 Å (Figure 3B) was obtained. Considering that this value is close to the definition of microporous materials, that is, 2 nm, we performed the t-Plot Method Micropore Analysis for the COF, and the result indicated that COF-TpAzo has the micropore surface area of 1286 m2·g−1 and external surface area of 266 m2·g−1. This result suggests that COF-TpAzo exhibited more microporous feature in nature, although the pore size is a little larger than 2 nm. This may be the reason why a high CO2/N2 selectivity was obtained in COF-TpAzo (this will be discussed below). CO2 Uptake and Selectivity. The CO2 adsorption of COF-TpAzo was obtained as high as 105.6 mg·g−1 (68.6 mg· g−1 at 298 K) (Figure 4A), which is higher than many previously reported COFs, such as ILCOF-1 (60 mg·g−1),10



RESULTS AND DISCUSSION The successful preparation of COF-TpAzo was confirmed by powder X-ray diffraction (PXRD) (Figure 2A), Fourier transform infrared (FTIR) spectroscopy (Figure 2B) and scanning electron microscopy (SEM) (Figure 1B) measurements. As shown in Figure 2A (blue curve), the experimental PXRD pattern displayed an intense diffraction peak at a 2θ B

DOI: 10.1021/acs.jced.6b00071 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 2. (A) Observed (blue) and simulated (pink) PXRD patterns of COF-TpAzo. (B) FT-IR spectra of COF-TpAzo. (C) TGA curves of the COF TpAzo.

Figure 3. (A) Nitrogen adsorption−desorption isotherms (filled diamond, adsorption; open diamond, desorption). (B) Pore size distribution of COF-TpAzo.

COF-5 (59 mg·g−1),8 and COF-103 (76 mg·g−1)8 at 273 K and 1 bar. We reasoned that the cooperative interactions between CO2 molecules and azo group as well as imine group in skeleton wall favor the adsorption of CO2. The selective uptake of CO2 over N2 and CH4 at 273 and 298 K have been investigated to evaluate the potential application of COFTpAzo in the gas separation field. On the basis of initial slope calculations in the pressure range, the ideal adsorption selectivity of CO2/N2 and CO2/CH4 (N2 and CH4 adsorption isotherms were presented in Figure 4B and C) is 127 and 39 at 273 K, 145 and 43 at 298 K (Figure S2), higher than those reported porous polymers (Table S1).25,29−35 These results indicated that the azo-based COF-TpAzo is efficient for CO2/ N2 separation as expected. Interestingly, contrary to the common trend, the CO2/N2 and CO2/CH4 selectivities of COF-TpAzo at 298 K were even higher than 273 K. This unprecedented behavior was explained not only by conventional CO2 affinities but also by the N2 phobicity of azo group.25 Although the COF-TpAzo exhibited good performance with respect to CO2 selectivity over N2, their amenability to recycling and efficacy in the presence of moisture needs be considered. The chemical stability of COF-TpAzo has been measured by Banerjee et al.23 The COF-TpAzo exhibited excellent chemical stability both on crystalline and porous when directly submerged in boiling water and strong acid (9 N HCl). Although the porosity was almost unchanged after these treatments, we believe the CO2 adsorption affinity will be reduced in the presence of water moisture. To the best of our knowledge, many porous polymers reported so far exhibit significant loss in CO2/N2 selectivity due to water moisture excisting.17,35

The heat adsorption (Qst) for CO2 was calculated via the Clausius−Clapeyron equation using CO2 isotherms at 273 and 298 K (Figure 4D) and was found to be 32.0 kJ·mol−1 at zero coverage, which is higher than the values reported for COFs (see Table S1). CH4 and H2 Adsorption in COF-TpAzo. In addition to CO2 storage and separation by COFs, extensive research on methane and hydrogen adsorption has been conducted because of their potential in clean energy applications. CH4 is known as a clean energy with high hydrogen−carbon ratio, low cost, huge reserves around world. The CH4 uptake by COF-TpAzo has been conducted because it is anticipated that aromatic building blocks and lone pairs at the nitrogen atoms will possess optimal CH4 uptake. We measured the CH4 uptake capacity of COFTpAzo (13.4 mg·g−1 at 273 K and 1 bar) (Figure 4B), which is higher than those of ILCOF-1 (9.0 mg·g−1)10 and TDCOF-5 (10.7 mg·g−1)36 under the same conditions. It was also found that CH4 uptake decreased to 8.4 mg·g−1 when temperature was increased to 298 K, which indicates that temperature has a great influence on CH4 uptake (Figure 4B). The Qst for CH4 was found to be 16.5 kJ·mol−1 at zero coverage (Figure 4D). Hydrogen adsorption performance was analyzed at 77 K and 1 bar to examine the functionality of this mesoporous COFTpAzo. The H2 uptake ability is 10.6 mg·g−1 (Figure 4E), which is comparable to the previously reported COFs under the same conditions, like COF-10 (9.0 mg·g−1),8 CoPc-BPDA COF (12.0 mg·g−1),37 and ACOF-1 (9.9 mg·g−1).12 Simulation of H2 Adsorption in COF-TpAzo. To determine the H2 isotherm for COF-TpAzo as a function of pressure (0 to 1 bar) at 77 K, the grand canonical Monte Carlo (GCMC) simulations were carried out using the predicted C

DOI: 10.1021/acs.jced.6b00071 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 4. Gas adsorption−desorption isotherms of COF-TpAzo at 273 K (square) and 298 K (circles) (A) CO2, (B) CO2 and CH4 adsorption enthalpy, Qst, as calculated with adsorption curves at two different temperatures (273 and 298 K), (C) CH4, (D) N2, and (E) H2.

eclipsed structure, as shown in Figure 1. The nonbonded interactions of H2 and COF-TpAzo were described using the Morse functional form. The van der Waals FF parameters38 were shown in Table S2. To obtain an accurate measure of H2 loading, we constructed 10 000 000 configurations to compute the average loading for each condition. For the simulation, we used COF-TpAzo eclipsed structures (Figure 1). The sorbent model is an infinite three-dimensional periodic super cell, 2 × 2 × 2 for COF-TpAzo of each unit cell to eliminate boundary effects. Also, through the GCMC simulation, we investigated H2 adsorption sites in COF-TpAzo at 77 K. Figure 5 shows snapshots of structures of COF-TpAzo with adsorbed H2 at 1 bar. In this case, one can find that the most favored adsorption site of H2 is on azo and benzene rings in the organic linkers. The H2 adsorption of COF-TpAzo was predicted as 13.7 mg· g−1 at 1 bar and 77 K (Table S3), which is in reasonable agreement with experimental result (10.6 mg·g−1 at 1 bar and 77 K)

Figure 5. GCMC snapshots of the structures of COF-TpAzo with adsorbed H2 at 0.7 bar. The atoms are colored as follows: gray = C, red = O, blue = N, and cyan = H.



separation. As expected, the COF-TpAzo exhibited high adsorption selectivity of CO2/N2 is 127 (145) and CO2/CH4 is 39 (43) at 273 K (298 K). Moreover, it is worth mentioning that COF-TpAzo has high adsorption capacities for both H2 and CH4 due to the excellent porosity with a high BET surface area and large pore volume. We believe that this strategy could

CONCLUSIONS In summary, we have synthesized a stable “CO2-philic” and “N2-phobic” COF material, COF-TpAzo, by introducing azo function group in the skeleton of framework for CO2/N2 D

DOI: 10.1021/acs.jced.6b00071 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(10) Rabbani, M. G.; Sekizkardes, A. K.; Kahveci, Z.; Reich, T. E.; Ding, R. S.; El-Kaderi, H. M. A 2D Mesoporous Imine-Linked Covalent Organic Framework for High Pressure Gas Storage Applications. Chem. - Eur. J. 2013, 19, 3324−3328. (11) Tilford, R. W.; Mugavero, S. J.; Pellechia, P. J.; Lavigne, J. J. Tailoring microporosity in covalent organic frameworks. Adv. Mater. 2008, 20, 2741−2746. (12) Li, Z. P.; Feng, X.; Zou, Y. C.; Zhang, Y. W.; Xia, H.; Liu, X. M.; Mu, Y. A 2D azine-linked covalent organic framework for gas storage applications. Chem. Commun. 2014, 50, 13825−13828. (13) Wan, S.; Guo, J.; Kim, J.; Ihee, H.; Jiang, D. L. A Belt-Shaped, Blue Luminescent, and Semiconducting Covalent Organic Framework. Angew. Chem., Int. Ed. 2008, 47, 8826−8830. (14) Xu, H.; Chen, X.; Gao, J.; Lin, J. B.; Addicoat, M.; Irle, S.; Jiang, D. L. Catalytic covalent organic frameworks via pore surface engineering. Chem. Commun. 2014, 50, 1292−1294. (15) Deng, H. X.; Grunder, S.; Cordova, K. E.; Valente, C.; Furukawa, H.; Hmadeh, M.; Gandara, F.; Whalley, A. C.; Liu, Z.; Asahina, S.; Yaghi, O. M.; et al. Large-Pore Apertures in a Series of Metal-Organic Frameworks. Science 2012, 336, 1018−1023. (16) D’Alessandro, D. M.; Smit, B.; Long, J. R. Carbon Dioxide Capture: Prospects for New Materials. Angew. Chem., Int. Ed. 2010, 49, 6058−6082. (17) Thomas, A. Functional Materials: From Hard to Soft Porous Frameworks. Angew. Chem., Int. Ed. 2010, 49, 8328−8344. (18) Du, N. Y.; Park, H. B.; Robertson, G. P.; Dal-Cin, M. M.; Visser, T.; Scoles, L.; Guiver, M. D. Polymer nanosieve membranes for CO2capture applications. Nat. Mater. 2011, 10, 372−375. (19) Farha, O. K.; Spokoyny, A. M.; Hauser, B. G.; Bae, Y. S.; Brown, S. E.; Snurr, R. Q.; Mirkin, C. A.; Hupp, J. T. Synthesis, Properties, and Gas Separation Studies of a Robust Diimide-Based Microporous Organic Polymer. Chem. Mater. 2009, 21, 3033−3035. (20) Patel, H. A.; Karadas, F.; Canlier, A.; Park, J.; Deniz, E.; Jung, Y.; Atilhan, M.; Yavuz, C. T. High capacity carbon dioxide adsorption by inexpensive covalent organic polymers. J. Mater. Chem. 2012, 22, 8431−8437. (21) Nagai, A.; Guo, Z.; Feng, X.; Jin, S.; Chen, X.; Ding, X.; Jiang, D. Pore surface engineering in covalent organic frameworks. Nat. Commun. 2011, 2, 536−543. (22) Hug, S.; Stegbauer, L.; Oh, H.; Hirscher, M.; Lotsch, B. V. Nitrogen-Rich Covalent Triazine Frameworks as High-Performance Platforms for Selective Carbon Capture and Storage. Chem. Mater. 2015, 27, 8001−8010. (23) 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. (24) Patel, H. A.; Je, S. H.; Park, J.; Jung, Y.; Coskun, A.; Yavuz, C. T. Directing the Structural Features of N-2-Phobic Nanoporous Covalent Organic Polymers for CO2 Capture and Separation. Chem. - Eur. J. 2014, 20, 772−780. (25) Patel, H. A.; Je, S. H.; Park, J.; Chen, D. P.; Jung, Y.; Yavuz, C. T.; Coskun, A. Unprecedented high-temperature CO2 selectivity in N2-phobic nanoporous covalent organic polymers. Nat. Commun. 2013, 4, 1357−1364. (26) Santurri, P.; Robbins, F.; Stubbings, R. 4,4′-Diaminoazobenzene. Organic Syntheses 2003, 40, 18. (27) Chong, J. H.; Sauer, M.; Patrick, B. O.; MacLachlan, M. J. Highly stable keto-enamine salicylideneanilines. Org. Lett. 2003, 5, 3823−3826. (28) Cote, A. P.; El-Kaderi, H. M.; Furukawa, H.; Hunt, J. R.; Yaghi, O. M. Reticular synthesis of microporous and mesoporous 2D covalent organic frameworks. J. Am. Chem. Soc. 2007, 129, 12914− 12915. (29) Wang, L.; Dong, B.; Ge, R.; Jiang, F.; Xiong, J.; Gao, Y.; Xu, J. A thiadiazole-functionalized covalent organic framework for efficient CO2 capture and separation. Microporous Mesoporous Mater. 2016, 224, 95−99.

be popularized to design other functionalized COFs for a special purpose.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jced.6b00071. Summary of CO2 uptake, CO2/N2 and CO2/CH4 selectivity and Qst for CO2, of the reported COFs and other porous polymers; figure of BET Fitting of COFTpAzo; figure of initial slope selectivity of CO2/N2 and CO2/CH4 for COF-TpAzo at 273 and 298 K; the parameters of force field, predicted H2 adsorption amount for COF-TpAzo at 77 K from 0 to 1 bar. (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Fax: +86 411 84379992. Tel.: +86 411 84379992. Funding

This work was supported National Natural Science Foundation of China (21406215, 21273235, and 21303076). Notes

The authors declare no competing financial interest.

■ ■

ACKNOWLEDGMENTS We are thankful to Dr. Lei Shi (DICP) for the discussion on synthesis. 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) Spitler, E. L.; Koo, B. T.; Novotney, J. L.; Colson, J. W.; UribeRomo, F. J.; Gutierrez, G. D.; Clancy, P.; Dichtel, W. R. A 2D Covalent Organic Framework with 4.7-nm Pores and Insight into Its Interlayer Stacking. J. Am. Chem. Soc. 2011, 133, 19416−19421. (3) Feng, X.; Liu, L.; Honsho, Y.; Saeki, A.; Seki, S.; Irle, S.; Dong, Y.; Nagai, A.; Jiang, D. High-rate charge-carrier transport in porphyrin covalent organic frameworks: switching from hole to electron to ambipolar conduction. Angew. Chem., Int. Ed. 2012, 51, 2618−2622. (4) Kandambeth, S.; Mallick, A.; Lukose, B.; Mane, M. V.; Heine, T.; Banerjee, R. Construction of crystalline 2D covalent organic frameworks with remarkable chemical (acid/base) stability via a combined reversible and irreversible route. J. Am. Chem. Soc. 2012, 134, 19524−19527. (5) 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. (6) El-Kaderi, H. M.; Hunt, J. R.; Mendoza-Cortes, J. L.; Cote, A. P.; Taylor, R. E.; O’Keeffe, M.; Yaghi, O. M. Designed synthesis of 3D covalent organic frameworks. Science 2007, 316, 268−272. (7) Fang, Q. R.; Gu, S.; Zheng, J.; Zhuang, Z. B.; Qiu, S. L.; Yan, Y. S. 3D Microporous Base-Functionalized Covalent Organic Frameworks for Size-Selective Catalysis. Angew. Chem., Int. Ed. 2014, 53, 2878− 2882. (8) 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. (9) 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. E

DOI: 10.1021/acs.jced.6b00071 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(30) Li, Z.; Zhi, Y.; Feng, X.; Ding, X.; Zou, Y.; Liu, X.; Mu, Y. An Azine-Linked Covalent Organic Framework: Synthesis, Characterization and Efficient Gas Storage. Chem. - Eur. J. 2015, 21, 12079− 12084. (31) Stegbauer, L.; Hahn, M. W.; Jentys, A.; Savasci, G.; Ochsenfeld, C.; Lercher, J. A.; Lotsch, B. V. Tunable Water and CO2 Sorption Properties in Isostructural Azine-Based Covalent Organic Frameworks through Polarity Engineering. Chem. Mater. 2015, 27, 7874−7881. (32) Mondal, S.; Das, N. Triptycene based 1,2,3-triazole linked network polymers (TNPs): small gas storage and selective CO2 capture. J. Mater. Chem. A 2015, 3, 23577−23586. (33) Lu, J.; Perez-Krap, C.; Suyetin, M.; Alsmail, N. H.; Yan, Y.; Yang, S. H.; Lewis, W.; Bichoutskaia, E.; Tang, C. C.; Blake, A. J.; et al. Robust Binary Supramolecular Organic Framework (SOF) with High CO2 Adsorption and Selectivity. J. Am. Chem. Soc. 2014, 136, 12828− 12831. (34) Reich, T. E.; Behera, S.; Jackson, K. T.; Jena, P.; El-Kaderi, H. M. Highly selective CO2/CH4 gas uptake by a halogen-decorated borazine-linked polymer. J. Mater. Chem. 2012, 22, 13524−13528. (35) Rabbani, M. G.; El-Kaderi, H. M. Template-free synthesis of a highly porousbenzimidazole-linked polymer for CO2 capture and H2 storage. Chem. Mater. 2011, 23, 1650−1653. (36) Neti, V. S. P. K.; Wu, X. F.; Hosseini, M.; Bernal, R. A.; Deng, S. G.; Echegoyen, L. Synthesis of a phthalocyanine 2D covalent organic framework. CrystEngComm 2013, 15, 7157−7160. (37) Kahveci, Z.; Islamoglu, T.; Shar, G. A.; Ding, R.; El-Kaderi, H. M. Targeted synthesis of a mesoporous triptycene-derived covalent organic framework. CrystEngComm 2013, 15, 1524−1527. (38) Han, S. S.; Choi, S. H.; Goddard, W. A. Zeolitic Imidazolate Frameworks as H2 Adsorbents: Ab Initio Based Grand Canonical Monte Carlo Simulation. J. Phys. Chem. C 2010, 114, 12039−12047.

F

DOI: 10.1021/acs.jced.6b00071 J. Chem. Eng. Data XXXX, XXX, XXX−XXX