Benzotrithiophene-Based Covalent Organic Frameworks: Construction

Sep 4, 2018 - Hexacene Diimides. Journal of the American Chemical Society. Cui, Xiao, Winands, Koch, Li, Zhang, Doltsinis, and Wang. 2018 140 (38), pp...
0 downloads 0 Views 2MB Size
Subscriber access provided by University of South Dakota

Communication

Benzotrithiophene-based Covalent Organic Frameworks: Construction and Structure Transformation under Ionothermal Condition Hongtao Wei, Jing Ning, Xingdi Cao, Xuehui Li, and Long Hao J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b08282 • Publication Date (Web): 04 Sep 2018 Downloaded from http://pubs.acs.org on September 4, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 9 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

Journal of the American Chemical Society

Benzotrithiophene-based Covalent Organic Frameworks: Construction and Structure Transformation under Ionothermal Condition Hongtao Wei,†,§ Jing Ning,†,§ Xingdi Cao,†,§ Xuehui Li,† Long Hao*,† †

College of Chemistry and Pharmaceutical Sciences, Qingdao Agricultural University, No.700 Changcheng Road, Qingdao 266109, China

Supporting Information Placeholder ABSTRACT:

A C3-symmetric benzotrithiophene tricarbaldehyde (BTT) is synthesized for the first time with a facile method, which is used to construct BTT-based covalent organic frameworks (COFs) with different pore sizes. Meanwhile, the structure transformations of the COFs under high-temperature ionothermal condition are studied systematically, and the relationships between the regularly changed structures of the further-crosslinked COFs and their supercapacitor performances are investigated. It turns out that dealing COFs of designated structures under ionothermal condition is a great method to build high-conductive structure-controllable materials for electrochemical applications.

Covalent organic Frameworks (COFs) are crystalline reticular materials constructed through the strong covalent bonds of symmetric organic molecules. The appear of COFs extends the organic reactions beyond small molecules.1 The diversities of organic species and reaction types for building COFs also open up a new direction in the molecular level to precisely control the structures of light-element-constituted materials,2,3 including the

controlling of pore sizes,4-7 surface functionalities,8-11 skeleton structures,12-15 heteroatom doping,16-19 and structure defects.20-23 These unique advantages of COFs make their applications gradually expanding into numerous fields over the past decade.24-26 In the electrochemical fields, such as lithium ion battery,27-29 lithium-sulfur battery,30,31 supercapacitor,32,33 or 34,35 electrocatalysis, COFs have also shown their significant value. Their porous structures with ultrahigh surface areas and intrinsic heteroatom doping are extremely favorable for enhancing the electrochemical performances. Their controllable structures also make it possible to deeply study the mechanism and structure-performance relationships in electrochemistry. However the relatively poor conductivities of COFs always hinder their broader applications. The usual improving method is adding plenty of conductive addictives (e.g. carbon nanotube or conductive carbon), which sometimes still can’t fulfill the high-conductivity demand of some areas including supercapacitor and electrocatalysis. Consequently, the high-temperature treatment of the materials is essential.36,37 Our previous works

Figure 1. (a) Key reaction for synthesizing the C3-symmetric benzotrithiophene tricarbaldehyde, BTT. (b) Construction of the model compound, MC with the Schiff-base reaction. (c) Schematic representation of the syntheses of BTT-based COFs with different pore sizes.

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

have shown that dealing the covalent triazine-based materials (CTFs) with ionothermal method can enhance their conductivity and maintain parts of their inherent structures as well.33,38,39 The structure transformations of heteroatom-contained Schiff-base/imine COFs under ionothermal condition still need to be studied. In all the COF-related researches, designs and syntheses of the symmetric molecular building blocks usually play the most important role. Benzotrithiophene is a classic C3-symmetric construction unit for solar cells,40,41 self-assemblies,42,43 and metal organic frameworks.44 Herein, a new benzotrithiophene tricarbaldehyde (BTT) was synthesized through a facile method

Page 2 of 9

with the reaction between p-dithiane-2,5-diol and 1,3,5-trichloro-2,4,6-tris(dichloromethyl)benzene (Figure 1a, Figure S1).45,46 The supposed mechanism may include a Et3N-catalytic nucleophilic substitution (SNAr) reaction followed by an intramolecular Adol reaction (Figure S2).47 The as-prepared BTT was first reacted with 4-tert-Butyl benzenamine to test the reactivity of its aldehyde groups (Figure 1b), which revealed that the Schiff-base reaction happened quickly to get the model compound (MC). Then, three aminobenzene derivatives (DADP, DAB and TAB) were selected to react with BTT, aiming to get Schiff-base BTT-COFs with different pore sizes (BTT-DADP COF, BTT-DAB COF and BTT-TAB COF, Figure 1c). All the reaction details are listed in Supporting Information.

Figure 2. Characterizations of the BTT-based COFs. (a) Experimental and simulated XRD patterns with the corresponding structural schemes, stacking ways, lattice parameters and space groups inserted. (b) Typical TEM image of BTT-based COFs (from BTT-DAB COF). (c) Nitrogen adsorption-desorption isotherms with pore size distribution profiles inserted. (d) FT-IR spectra compared with BTT and MC. (e) UV/Vis DRS spectra with Kubelka-Munk-transformed reflectance spectra inserted. The crystalline structures of BTT-based COFs were first characterized by the X-ray diffraction (XRD) measurements. Through comparing the experimental XRD patterns with the simulated ones, it can be seen that the COFs are all of layered structures with eclipsed stacking ways (Figure 2a, the fractional atomic coordinates in Table S1-S3, schematic for the staggered structures in Figure S3). The layered structures of COFs can also be revealed by the typical transmission electron microscope (TEM) image (Figure 2b). The space groups of BTT-DADP COF and BTT-DAB COF belong to P6/m symmetry, and the space group of BTT-TAB COF belongs to P-6 symmetry. From BTT-DADP COF, BTT-DAB COF to BTT-TAB COF, the diffraction angles of the (100) planes gradually get bigger, suggesting that the pore sizes of the COFs gradually become smaller, which is in correspondence with the decreasing simulated lattice parameters (a and b, Figure 2a). This can also be testified by the results of pore size distribution analyses, which are calculated based on the nitrogen adsorption-desorption measurements at 77K (Figure 2c, average pore sizes in Table S4). Besides, the Brunauer-Emmett-Teller (BET) specific surface areas

(SSAs) of the COFs also become smaller along with the decreasing pore sizes (insertion in Figure 2c). The Fourier transform infrared (FT-IR) spectra of the COFs and the MC show that the C=N stretching for imines at 1600 cm-1 emerges, and the C=O stretching for aldehyde groups at 1670 cm-1 disappears compared with BTT (Figure 2d), proving that the COFs are indeed constructed through the Schiff-base reactions. The ultraviolet–visible diffuse reflectance spectra (UV/vis DRS) were also used to test these colored COF powders. From BTT-TAB COF (yellow), BTT-DADP COF (light orange) to BTT-DAB COF (dark red), an obvious red-shift of the optical absorption edge can be observed, and the corresponding optical band gaps decrease from 2.38, 2.25 to 2.04 eV (Figure 2e, digital photographs in Figure S4). The variation tendency of the band gaps is obviously not consistent with the pore size changes of the COFs, they may also be influenced by the different space groups of the COFs (Figure 2a), which is still under study. Nevertheless, the test results of UV/vis DRS here still make BTT-based COFs good candidates for the visible-light photocatalysis.

ACS Paragon Plus Environment

Page 3 of 9 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

Journal of the American Chemical Society

Figure 3. Structural characterizations and electrochemical performances of COF-700s. (a) XRD patterns with the JCPDS card of ZnS2 (No. 36-1450) inserted (BTT-TAB COF-700-untreated was derived from BTT-TAB COF-700 untreated with hot HCl solution during workup). (b) Nitrogen adsorption-desorption isotherms with pore size distribution profiles inserted. (c) Nitrogen contents in COF-700s (from XPS analyses) compared with COFs (calculated). (d) The ratios of sp2 hybridized carbon in COF-700s (from XPS analyses) compared with the calculated carbon contents in COFs. (e) the specific capacitances at different current densities with the typical GC curves inserted. (f) Cumulative specific surface area (SSA) with pore size larger than the ion diameter of the electrolyte (effective SSA, E-SSA); the insertion is the specific capacitance at 0.5 A/g versus the E-SSA, and slopes of the lines are proportional to the permittivity, ε in the formula for supercapacitors: C=εS/d. To enhance the conductivity of the COFs for electrochemical applications, the COFs were treated under ionothermal condition with five times mass of ZnCl2 at 700℃ for 20h (see details in Supporting Information). The further cross-linked COFs are named as COF-700s: BTT-DADP COF-700, BTT-DAB COF-700 and BTT-TAB COF-700, respectively. Notably, if ZnCl2 was not added, the prepared sample would have very low BET SSA and irregular pore size distribution (e.g. 15.6 m2/g for the 700℃ -treated BTT-DAB COF without ZnCl2, Figure S5). Here, the ZnCl2 might have the function of padding and supporting the pore structures of the materials during the high-temperature treatment. Compared with the crystalline structures of COFs, COF-700s all change to amorphous (XRD patterns in Figure 3a), which can also be testified by the typical TEM image in Figure S6. If the samples were not washed adequately with hot HCl solution during workup, the XRD peaks indexed to ZnS2 could be observed (Figure 2a), which suggested that the sulfur atoms in the BTT were reacted with ZnCl2 to form ZnS2 under the ionothermal condition. This can also explain why the sulfur contents in COF-700s are all lower than 1 wt.% (XPS results in Table S5). The results of nitrogen adsorption-desorption measurements reveal that the BET SSAs and average pore sizes of COF-700s all become larger in contrast with that of the COFs (Figure 3b, Table S4). Besides, the SSAs and pore sizes gradually become smaller from BTT-DADP COF-700, BTT-DAB COF-700 to BTT-TAB COF-700 (Figure 3b), which have the same variation tendency compared with that of the COFs (Figure 2c). The nitrogen contents in COF-700s and the corresponding COFs maintain almost the same change regularity (Figure 3c). More interestingly, the ratios of sp2 hybridized carbon in COF-700s vary along with

the carbon contents in the corresponding COFs (Figure 3d and Figure S7),48 which means COFs with higher carbon contents finally transform to COF-700s with higher ratios of sp2 carbon. These test results indicate that the structure characteristics of COF-700s are all transformed regularly according to the original COFs. The electrochemical performances of COF-700s were characterized by a symmetric supercapacitor system (a CR2032 coin-type cell) with an ionic liquid, 1-ethyl-3-methylimidazolium tetrafluoroborate (EMIMBF4) as the electrolyte (details in Supporting Information). The quasi-rectangular cyclic voltammetry (CV) curves (Figure S8) and triangular-shape galvanostatic charge-discharge (GC) curves (Figure S9, typically in Figure 3e) show the typical double-layer supercapacitor characteristics of COF-700-based supercapacitors. The specific capacitances caculated based on the discharge slopes of GC curves become smaller from BTT-DADP COF-700, BTT-TAB COF-700 to BTT-DAB COF-700 (Figure 3e). The rate capacity of BTT-TAB COF-700 is worse than that of BTT-DADP COF-700 and BTT-DAB COF-700, which is because of its relatively bigger equivalent series resistance (results of AC impedence in Figure S10).49 The typical cycling perfomance test of BTT-DADP COF-700 at current density of 10 A/g shows that the specific capacitance still remain 77.5% after 10 000 cycles (Figure S11). To further investigate the relationships between the structures of COF-700s and their supercapacitor performances, the cumulative surface areas of the pores bigger than the diameter of the electrolyte (0.75nm for EMIM+) were extracted (Figure 3f), which could be treated as the effective SSA (E-SSA) for the supercapacitors.33,50 According to the fundamental formula for

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

supercapacitors: C=εS/d, d is the thickness of the electric double layer, which is a constant when the electrolyte and testing conditions are fixed. Consequently, the permittivity (ε), which represents the interaction between the electrolyte and electrode material is proportional to the C/S (specific capacitance/E-SSA, slopes of the lines in the inserted figure of Figure 3f). Through comparing the slopes, it can be concluded that the interactions between the COF-700s and the electrolyte gradually become weaker form BTT-TAB COF-700, BTT-DADP COF-700 to BTT-DAB COF-700. This variation tendency is inconsonant with the changes of pore sizes and nitrogen contents of COF-700s (Figure 3b and 3c), but is accordant with the change of the ratios of sp2 carbon in COF-700s (Figure 3d), which means COF-700 with the higher ratio of sp2 carbon has the stronger interaction with the electrolyte. This also indicates that the E-SSA and the interaction between electrode materials and the electrolyte (ε) should be carefully balanced, and besides the pore sizes and heteroatom-doping, the ratio of sp2 carbon in the materials should also be concerned when designing and synthesizing the materials for supercapacitors. In summary, a new C3-symmetric benzotrithiophene tricarbaldehyde (BTT) and a series of BTT-based COFs have been constructed for the first time. These COFs display gradually increased pore sizes and SSAs, and also show good potential for visible-light photocatalysis with optical band gaps of 2.04-2.08 eV. More importantly, the further-crosslinked COFs under ionothermal condition (COF-700s) reveal regularly changed pore sizes, SSAs, nitrogen contents and ratios of sp2 carbon according to the structures of the corresponding COFs. And the test results of COF-700-based supercapacitors show that the interaction between COF-700s and the electrolyte is related to the ratios of sp2 carbon in COF-700s. In short, dealing COFs of designated structures under ionothermal condition is a great method to build high-conductive structure-controllable materials for electrochemical applications. This method is extremely beneficial for both the in-depth study of the structure-performance relationships in electrochemistry and the further enhancement of electrochemical performances.

ASSOCIATED CONTENT Supporting Information Supporting Information consists of experimental section, Figures S1-S11, and Tables S1-S5. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author L.H., E-mail: [email protected]

Author Contributions §

H.W., J.N. and X.C. contribute equally to this work.

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT This work was supported by National Natural Science Foundation of China (51603114), Natural Science Foundation of Shandong Province (ZR2016EMQ03), Science and Technology Foundation of Qingdao City (16-5-1-43-jch) and Doctorial Fund of Qingdao Agriculture University (663-1115046, 663-1117016 and 663/1113309). The authors acknowledge School of Material

Page 4 of 9

Science and Engineering, Ocean University of China for the support in XRD simulations.

REFERENCES (1) Diercks, C. S.; Yaghi, O. M. Science 2017, 355. (2) Huang, N.; Wang, P.; Jiang, D. Nat. Rev. Mater. 2016, 1, 16068. (3) Das, S.; Heasman, P.; Ben, T.; Qiu, S. Chem. Rev. 2017, 117, 1515. (4) Guan, X.; Ma, Y.; Li, H.; Yusran, Y.; Xue, M.; Fang, Q.; Yan, Y.; Valtchev, V.; Qiu, S. J. Am. Chem. Soc. 2018, 140, 4494. (5) Zhai, L.; Huang, N.; Xu, H.; Chen, Q.; Jiang, D. Chem. Commun. 2017, 53, 4242. (6) Pang, Z.-F.; Xu, S.-Q.; Zhou, T.-Y.; Liang, R.-R.; Zhan, T.-G.; Zhao, X. J. Am. Chem. Soc. 2016, 138, 4710. (7) Fang, Q.; Zhuang, Z.; Gu, S.; Kaspar, R. B.; Zheng, J.; Wang, J.; Qiu, S.; Yan, Y. Nat. Commun. 2014, 5, 4503. (8) Lu, Q.; Ma, Y.; Li, H.; Guan, X.; Yusran, Y.; Xue, M.; Fang, Q.; Yan, Y.; Qiu, S.; Valtchev, V. Angew. Chem. Int. Edit. 2018, 57, 6042. (9) Diercks, C. S.; Lin, S.; Kornienko, N.; Kapustin, E. A.; Nichols, E. M.; Zhu, C.; Zhao, Y.; Chang, C. J.; Yaghi, O. M. J. Am. Chem. Soc. 2018, 140, 1116. (10) Lohse, M. S.; Stassin, T.; Naudin, G.; Wuttke, S.; Ameloot, R.; De Vos, D.; Medina, D. D.; Bein, T. Chem. Mater. 2016, 28, 626. (11) Ding, S.-Y.; Dong, M.; Wang, Y.-W.; Chen, Y.-T.; Wang, H.-Z.; Su, C.-Y.; Wang, W. J. Am. Chem. Soc. 2016, 138, 3031. (12) Pachfule, P.; Acharjya, A.; Roeser, J.; Langenhahn, T.; Schwarze, M.; Schomäcker, R.; Thomas, A.; Schmidt, J. J. Am. Chem. Soc. 2018, 140, 1423. (13) Waller, P. J.; Lyle, S. J.; Osborn Popp, T. M.; Diercks, C. S.; Reimer, J. A.; Yaghi, O. M. J. Am. Chem. Soc. 2016, 138, 15519. (14) Ascherl, L.; Sick, T.; Margraf, J. T.; Lapidus, S. H.; Calik, M.; Hettstedt, C.; Karaghiosoff, K.; Döblinger, M.; Clark, T.; Chapman, K. W.; Auras, F.; Bein, T. Nat. Chem. 2016, 8, 310. (15) Vyas, V. S.; Haase, F.; Stegbauer, L.; Savasci, G.; Podjaski, F.; Ochsenfeld, C.; Lotsch, B. V. Nat. Commun. 2015, 6, 8508. (16) Huang, N.; Zhai, L.; Xu, H.; Jiang, D. J. Am. Chem. Soc. 2017, 139, 2428. (17) Sun, Q.; Aguila, B.; Perman, J.; Nguyen, N.; Ma, S. J. Am. Chem. Soc. 2016, 138, 15790. (18) Medina, D. D.; Petrus, M. L.; Jumabekov, A. N.; Margraf, J. T.; Weinberger, S.; Rotter, J. M.; Clark, T.; Bein, T. ACS Nano 2017, 11, 2706. (19) Zhang, D. S.; Chang, Z.; Lv, Y. B.; Hu, T. L.; Bu, X. H. Rsc Adv. 2012, 2, 408. (20) Kissel, P.; Erni, R.; Schweizer, W. B.; Rossell, M. D.; King, B. T.; Bauer, T.; Gotzinger, S.; Schluter, A. D.; Sakamoto, Nat. Chem. 2012, 4, 287. (21) Evans, A. M.; Parent, L. R.; Flanders, N. C.; Bisbey, R. P.; Vitaku, E.; Kirschner, M. S.; Schaller, R. D.; Chen, L. X.; Gianneschi, N. C.; Dichtel, W. R. Science 2018, 361, 52. (22) Ma, T.; Kapustin, E. A.; Yin, S. X.; Liang, L.; Zhou, Z.; Niu, J.; Li, L.-H.; Wang, Y.; Su, J.; Li, J.; Wang, X.; Wang, W. D.; Wang, W.; Sun, J.; Yaghi, O. M. Science 2018, 361, 48. (23) Auras, F.; Ascherl, L.; Haldmioun, A. H.; Margraf, J. T.; Hanusch, F. C.; Reuter, S.; Bessinger, D.; Doblinger, M.; Hettstedt, C.; Karaghiosoff, K.; Herbert, S.; Knochel, P.; Clark, T.; Bein, T. J. Am. Chem. Soc. 2016, 138, 16703. (24) Slater, A. G.; Cooper, A. I. Science 2015, 348, 988. (25) Ding, S. Y.; Wang, W. Chem. Soc. Rev. 2013, 42, 548. (26) Xiang, Z.; Cao, D. J. Mater. Chem. A 2013, 1, 2691. (27) Xu, F.; Jin, S.; Zhong, H.; Wu, D.; Yang, X.; Chen, X.; Wei, H.; Fu, R.; Jiang, D. Sci. Rep. 2015, 5, 8225. (28) Wang, S.; Wang, Q.; Shao, P.; Han, Y.; Gao, X.; Ma, L.; Yuan, S.; Ma, X.; Zhou, J.; Feng, X.; Wang, B. J. Am. Chem. Soc. 2017, 139, 4258. (29) Yang, D. H.; Yao, Z. Q.; Wu, D.; Zhang, Y. H.; Zhou, Z.; Bu, X. H. J. Mater. Chem. A 2016, 4, 18621. (30) Talapaneni, S. N.; Hwang, T. H.; Je, S. H.; Buyukcakir, O.; Choi, J. W.; Coskun, A. Angew. Chem. Int. Edit. 2016, 55, 3106. (31) Ghazi, Z. A.; Zhu, L. Y.; Wang, H.; Naeem, A.; Khattak, A. M.; Liang, B.; Khan, N. A.; Wei, Z. X.; Li, L. S.; Tang, Z. Y. Adv. Energy Mater. 2016, 6, 1601250. (32) Xu, F.; Xu, H.; Chen, X.; Wu, D.; Wu, Y.; Liu, H.; Gu, C.; Fu, R.; Jiang, D. Angew. Chem. Int. Edit. 2015, 54, 6814.

4

ACS Paragon Plus Environment

Page 5 of 9 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

Journal of the American Chemical Society (33) Hao, L.; Ning, J.; Luo, B.; Wang, B.; Zhang, Y.; Tang, Z.; Yang, J.; Thomas, A.; Zhi, L. J. Am. Chem. Soc. 2015, 137, 219. (34) Lin, S.; Diercks, C. S.; Zhang, Y.-B.; Kornienko, N.; Nichols, E. M.; Zhao, Y.; Paris, A. R.; Kim, D.; Yang, P.; Yaghi, O. M.; Chang, C. J. Science 2015, 349, 1208. (35) Nandi, S.; Singh, S. K.; Mullangi, D.; Illathvalappil, R.; George, L.; Vinod, C. P.; Kurungot, S.; Vaidhyanathan, R. Adv. Energy Mater. 2016, 6, 1601189. (36) Kou, Y.; Xu, Y.; Guo, Z.; Jiang, D. Angew. Chem. Int. Edit. 2011, 50, 8753. (37) Zhang, J.; Zhao, Z.; Xia, Z.; Dai, L. Nat. Nanotechnol. 2015, 10, 444. (38) Hao, L.; Luo, B.; Li, X.; Jin, M.; Fang, Y.; Tang, Z.; Jia, Y.; Liang, M.; Thomas, A.; Yang, J.; Zhi, L. Energ. Environ. Sci. 2012, 5, 9747. (39) Hao, L.; Zhang, S.; Liu, R.; Ning, J.; Zhang, G.; Zhi, L. Adv. Mater. 2015, 27, 3190. (40) Molina-Ontoria, A.; Zimmermann, I.; Garcia-Benito, I.; Gratia, P.; Roldán-Carmona, C.; Aghazada, S.; Graetzel, M.; Nazeeruddin, M. K.; Martín, N. Angew. Chem. Int. Edit. 2016, 55, 6270. (41) Gu, C.; Huang, N.; Chen, Y.; Qin, L.; Xu, H.; Zhang, S.; Li, F.; Ma, Y.; Jiang, D. Angew. Chem. Int. Edit. 2015, 54, 13594.

(42) Xiao, Q.; Sakurai, T.; Fukino, T.; Akaike, K.; Honsho, Y.; Saeki, A.; Seki, S.; Kato, K.; Takata, M.; Aida, T. J. Am. Chem. Soc. 2013, 135, 18268. (43) Ikeda, T.; Adachi, H.; Fueno, H.; Tanaka, K.; Haino, T. J. Org. Chem. 2017, 82, 10062. (44) Feng, D.; Liu, T.-F.; Su, J.; Bosch, M.; Wei, Z.; Wan, W.; Yuan, D.; Chen, Y.-P.; Wang, X.; Wang, K.; Lian, X.; Gu, Z.-Y.; Park, J.; Zou, X.; Zhou, H.-C. Nat. Commun. 2015, 6, 5979/1. (45) Taerum, T.; Lukoyanova, O.; Wylie, R. G.; Perepichka, D. F. Org. Lett. 2009, 11, 3230. (46) Wei, H.; Sun, M.; Ji, M. J. Chem. Res. 2009, 2009, 359. (47) Yang, L.-L.; Hu, X.-L.; Tang, Z.-Q.; Li, X.-F. Chem. Lett. 2015, 44, 1515. (48) Zhang, C.; Fu, L.; Liu, N.; Liu, M.; Wang, Y.; Liu, Z. Adv. Mater. 2011, 23, 1020. (49) Ra, E. J.; Raymundo-Piñero, E.; Lee, Y. H.; Béguin, F. Carbon 2009, 47, 2984. (50) Zhang, L.; Yang, X.; Zhang, F.; Long, G.; Zhang, T.; Leng, K.; Zhang, Y.; Huang, Y.; Ma, Y.; Zhang, M.; Chen, Y. J. Am. Chem. Soc. 2013, 135, 5921.

Table of Contents artwork

5

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

90x51mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 6 of 9

Page 7 of 9 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

Journal of the American Chemical Society

83x47mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

85x43mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 8 of 9

Page 9 of 9 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

Journal of the American Chemical Society

A C3-symmetric benzotrithiophene tricarbaldehyde (BTT) and a series of BTT-based covalent organic frameworks (BTT-COFs) are constructed for the first time. Meanwhile, structure transformations of the COFs under high-temperature ionothermal condition are studied systematically. The BTT-COFs and ionothermaldealing method have shown great potential for photocatalysis and electrochemistry. 47x41mm (300 x 300 DPI)

ACS Paragon Plus Environment