Troger's base functionalized covalent triazine frameworks for CO2 capture

Liming Taoa, Fang Niub, Jingang Liuc, Tingmei Wanga and Qihua Wang*a
aState Key Laboratory of Solid Lubrication, Lanzhou Institute of Chemical Physics, Chinese Academy of Sciences, Lanzhou, Gansu 730000, P. R. China. E-mail: qhwang@licp.cas.cn; Fax: +86 0931 4968252; Tel: +86 0931 4968252
bCollege of Chemistry and Chemical Engineering, Lanzhou University, Lanzhou, Gansu 730000, P. R. China. E-mail: niufang@lzu.edu.cn
cSchool of Materials Sciences and Technology, China University of Geosciences, Beijing, 100083, P. R. China. E-mail: liujg@iccas.ac.cn

Received 23rd August 2016 , Accepted 28th September 2016

First published on 29th September 2016


Abstract

A series of covalent triazine frameworks containing Troger's base segments (CTF-TBs) were designed and prepared by dynamic trimerization under ionothermal conditions, based on the novel dicyano monomer of dicyano Troger's base (DCTB). The CTF-TBs were all amorphous and exhibited surface areas between 294 to 732 m2 g−1. The effect of reaction conditions including ratio of catalyst/monomer and temperature on the pore properties was discussed. Meanwhile, the adsorption capacity of CO2 for all the CTF-TBs was studied and the CO2 uptake capacity was up to 16.84 wt% at 273 K and 1.10 bar for CTF-TB-3.


Introduction

Microporous organic polymers (MOPs) are promising candidates for carbon capture and storage (CCS).1–3 Recently, due to the urgent demand for highly efficient CO2 adsorption materials with low density and high thermal stability, many CO2-philic functional groups, such as azo,4–6 Troger's base (TB),7–12 benzimidazole,13–15 amino,16 carbazole,17 and porphyrin,18 have been incorporated into MOPs. Most of the CO2-philic groups contain heteroatoms (e.g. nitrogen, phosphorus, and sulfur), which can form Lewis acid–base interactions between the heteroatoms and CO2 molecules. TB segments with double tertiary N atoms have also been designed as active sites for catalysis in organic synthesis,12,19,20 selective adsorption of CO2,7,8 and gas separation.10,21–25 In general, there are two pathways reported for the introduction of TB segments. One is the polymerization of monomers containing TBs. For example, conjugated microporous polymers and microporous polyimides have been prepared using the monomers of diiodotroger's base and diaminotroger's base, respectively.19,26 While in the other pathway, TB segments are formed directly during the polymerizations, for instance, polymers with intrinsic microporous (PIM-EA-TB) prepared by N. B. McKeown et al.25 and 3D microporous polymers (TB-MOP) prepared by S. Dai and coworkers.7 In these MOPs, TB segments have enhanced their CO2 uptakes effectively, proved the TBs are promising candidates in design of MOPs for CO2 adsorption.

On the other hand, covalent triazine frameworks (CTFs) have high nitrogen content and high surface areas, and their preparation procedure is adapted to a large scale.27–32 Nowadays, A. Thomas et al. developed a novel two-step procedure to prepare CTFs, in which an amorphous, non-porous polymer network was converted into an ordered, microporous covalent organic framework under ionothermal conditions.33 This method makes the CTF synthesis easier, safer and scalable. Although with high nitrogen loading of about 21.8 wt%, the CO2 uptake of CTF-1 derived from 1,4-dicyanobenzene (DCB) is not so high. Further modification of CTFs by other CO2-philic groups, e.g. F atoms,34 pyridine groups,35 and benzimidazole groups,36 have been proved their effectiveness in enhancement of CO2 adsorption.

In the present work, we combined CTF and TB together to prepare MOPs with high CO2 adsorptions. CTFs with TB segments are designed and prepared by the ionothermal polymerization procedure. We expect that the CO2-philic TB segments can enhance the binding affinity between the CTF-TBs and CO2, and thus increase the CO2 adsorption. The synthesis and properties of the CTF-TBs, the effect of reaction conditions on the porosity, CO2 adsorption ability and the selectivity of CO2/N2, have been studied in detail.

Experimental

Materials

4-Cyanoaniline, trifluoroacetic acid (TFA) and trifluorosulfonic acid (TFSA) were purchased from J&K Scientific Ltd. (Beijing) and used without further purification. Paraformaldehyde was purchased from Sinopharm Chemical Reagent Co., Ltd. and used directly. DCTB was synthesized in our laboratory according to the literature.26,37 Zinc chloride (ZnCl2, anhydrous, 98%) was purchased from J&K Scientific Ltd. (Beijing), stored in a glove box and used as received.

Characterization

Fourier transform infrared (FT-IR) spectra were recorded on IFS 66v/S Fourier transform spectrophotometer. The powder wide-angle X-ray diffraction (XRD) was conducted on an X'Pert PRO X-ray diffractometer (PANalytical) with Cu/K-α1 radiation, operated at 60 kV and 55 mA. Thermogravimetric analysis (TGA) was recorded on a Netzsch thermal analysis system (STA 449C), in nitrogen, at a heating rate of 10 °C min−1. Solid-state 13C NMR spectra were performed on a Bruker Avance III Spectrometer operating at a frequency of 100 MHz and using cross-polarization/magic angle spinning (CP/MAS). The magic angle spinning rate was set at 5.0 kHz to minimize spinning sideband overlap. Combustion elemental analysis was conducted on Vario EL III Elemental Analyzer System. X-ray photoelectron spectroscopy (XPS) was performed on ESCALAB 250Xi (ThermoFisher Scientific). SEM images were obtained with a field emission scanning electron microscope (JSM-5600LV, JEOL, Japan). The powder samples were coated with gold by sputtering prior to observation. TEM images were obtained with A JEM-2010 transmission electron microscopy (TEM), operating at an accelerating voltage of 200 kV. Nitrogen sorption experiments and micropore analysis were conducted at 77.3 K using Micromeritics ASAP 2020 HD88. Before sorption measurements, the samples were degassed in vacuum overnight at 150 °C. The surface areas were calculated from multipoint BET plot, and the pore volume was determined by nonlocal density functional theory (NLDFT). Carbon dioxide and nitrogen sorption isotherms at 273 K and 303 K were all obtained with a Micromeritics ASAP 2020 HD88 analyzer at the required temperature. The temperature during adsorption and desorption was kept constant using a circulator. Before the adsorption measurements, the samples were activated in situ by increasing the temperature to 150 °C at a heating rate of 10 °C min−1 under vacuum and the temperature and vacuum was maintained for 5 hours.

Synthesis of 2,8-dicyano-6H,12H-5,11-methanodibenzo[1,5]diazocine (DCTB)

In a 250 mL three-necked flask with a magnetic stir and N2 inlet, 5.91 g (50 mmol) 4-cyanoaniline and 3.15 g (105 mmol, 2.1 equiv.) paraformaldehyde were dissolved in 143 g (1.25 mol, 25 equiv.) trifluoroacetic acid (TFA), under ice-bath. The reaction mixture turned dark-brown gradually and was stirred for three days at nitrogen atmosphere under room temperature. Then the mixture was poured into 200 mL pure water under continuously stirring giving a light yellow precipitate. KOH solution (6 mol L−1) was gradually added to this suspension adjusting pH to 8.5–9. The precipitate was filtered off and washed several times to neutral. The crude product was then purified by column chromatography with hexane/AcOEt to give light yellow powders 3.4 g, yield 50%.

Synthesis of the CTF-TBs

The CTF-TBs were synthesized according to a reported ionothermal procedure.27 The synthesis of CTF-TB-1 was given as an example (Scheme 1). The DCTB monomer (136.2 mg, 0.50 mmol) and ZnCl2 (68.2 mg, 0.50 mmol) were first mixed well in agate mortar in glove box and then transferred into a pyrex ampoule (3 × 18 cm, about 25 mL). The ampoule was evacuated to vacuum, sealed and heated to 400 °C in a muffle furnace and kept there for 40 h. The ampoule was then cooled down to room temperature and opened carefully. The black reaction mixture was subsequently grounded in an agate mortar and then washed thoroughly with diluted HCl to remove most of the ZnCl2. Then the powder was extracted in Soxhlet extractor by deioned water and THF for 24 h each and drying in vacuum at 150 °C. The yield was about 82%. The other CTF-TBs were synthesized in a similar procedure, while the reaction conditions, including the molar ratio of monomer to ZnCl2 and reaction temperatures, were set according to Table 1.
image file: c6ra21196h-s1.tif
Scheme 1 Schematic representation of the synthesis of CTF-TBs.
Table 1 Overview of the reaction conditions
Polymer code DCTB (mmol) ZnCl2 (mmol) Molar ratio of DCTB/ZnCl2 Reaction conditions Yield (%)
CTF-TB-1 0.50 0.50 1[thin space (1/6-em)]:[thin space (1/6-em)]1 400 °C/40 h 82
CTF-TB-2 0.50 1.0 1[thin space (1/6-em)]:[thin space (1/6-em)]2 400 °C/40 h 80
CTF-TB-3 0.50 2.5 1[thin space (1/6-em)]:[thin space (1/6-em)]5 400 °C/40 h 80
CTF-TB-4 0.50 5.0 1[thin space (1/6-em)]:[thin space (1/6-em)]10 400 °C/40 h 84
CTF-TB-5 0.50 10.0 1[thin space (1/6-em)]:[thin space (1/6-em)]20 400 °C/40 h 84
CTF-TB-6 0.50 2.5 1[thin space (1/6-em)]:[thin space (1/6-em)]5 350 °C/40 h 84
CTF-TB-7 0.50 2.5 1[thin space (1/6-em)]:[thin space (1/6-em)]5 450 °C/40 h 74
CTF-TB-8 0.50 2.5 1[thin space (1/6-em)]:[thin space (1/6-em)]5 500 °C/40 h 71
CTF-TB-9 0.50 2.5 1[thin space (1/6-em)]:[thin space (1/6-em)]5 550 °C/40 h 65


In comparison, the super acid TFSA catalyzed polymerization of DCTB in solution was also performed as following. In a 50 mL round bottom flask equipped with a magnetic stir, 20 mL of anhydrous DCM and DCTB (272.3 mg, 1.0 mmol) were added. Under ice bath, TFSA (1.5 g, 10 mmol) was dropped into the mixture slowly. The color turned to deep red finally and some brownish powders precipitated. The mixture was stirred at room temperature for 3 days and it was poured into methanol. The brown powders were collected by filtration and washed several times to neutral and drying in vacuum at 150 °C. The sample was denoted as CTF-TB-S.

Results and discussion

Synthesis and characterization of the CTF-TBs

The monomer DCTB was synthesized by the dimerization of 4-cyanoaniline catalyzed by TFA with a good yield. In the FT-IR spectra of DCTB and 4-cyanoaniline (Fig. S1), the absorption bands at 3375 cm−1 and 3488 cm−1 disappeared, indicating the aminos had been consumed. The new absorption bands at around 2813 cm−1 and 2892 cm−1 could be assigned to the methylene groups in TB segments.7 Meanwhile, the characteristic absorption bands of nitrile groups can also be seen at 2226 cm−1. In the TGA curve (Fig. S4) of DCTB under N2, the first/main weight loss at about 350 °C indicating the decomposition of TB segments, in well accordance with the endo thermal peaks shown in the DSC curve at the same temperature range (Fig. S5).

The CTF-TBs were synthesized by ionothermal polymerization in sealed tube. In order to find the relationship between the reaction conditions and the porosity properties, a set of reaction conditions was listed in Table 1. The molar ratios of DCTB to ZnCl2 were set as 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]2, 1[thin space (1/6-em)]:[thin space (1/6-em)]5, 1[thin space (1/6-em)]:[thin space (1/6-em)]10, and 1[thin space (1/6-em)]:[thin space (1/6-em)]20, respectively. The reaction temperature was set as 350 °C, 400 °C, 450 °C, 500 °C and 550 °C. All the reaction time was set as 40 hours. In all cases, after polymerization, the light yellow powdery reactants turned black dust in the sealed glass tube. For CTF-TB-1 to CTF-TB-6, there were no positive pressures when the sealed tubes were opened. In comparison, for CTF-TB-7, CTF-TB-8 and CTF-TB-9 reacted at 450 °C, 500 °C and 550 °C, there was obviously positive pressure, which should be paid more attention. All black dusts were grounded in agate mortar to get fine powders. The black powder were all washed and extracted thoroughly to remove the residual ZnCl2. The final yields were in the range of 65–84%, which were a little lower than the reported results on CTF-1. This can be mainly attributed to the TB segments, which were relatively fragile to thermal treatment, especially at elevate temperatures (Fig. S4). For comparison, the polymerization catalyzed by TFSA in solution was also carried out at room temperature. The formation of CTF-TBs was characterized by FT-IR, PXRD, solid-state 13C NMR spectra, and thermogravimetric analysis (TGA).

Representative FT-IR spectra of polymer (CTF-TB-3, CTF-TB-S) and DCTB were shown in Fig. 1a. After polymerization, the nitrile groups at about 2226 cm−1 were totally disappeared, indicating a high degree of reaction. The characteristic peaks of triazine rings were not very strong, which were located at 1430 cm−1 and 870 cm−1, may be covered by the nearby peaks. At the same time, maybe because the TB segments were fragile to high temperature, the characteristic peaks at around 2800 cm−1 were very weak. The adsorption peaks of aliphatic C–H vibrations were all relatively weak in the other CTF-TBs (Fig. S2 and S3). As the reaction temperature rose up to 550 °C, the adsorption peaks of TB and triazine segments became even weaker (Fig. S3), due to the partially decomposition or fragmentation of both TB segments and triazine groups.38,39 In the solid-state 13C NMR spectra shown in Fig. 1b, the chemical signals located at 107.5 ppm should be assigned to the asymmetric nitrile carbons,40,41 which almost disappeared completely after reaction. The nitriles first trimerized to triazine groups and then partially decomposed under high temperatures. Accordingly, the peaks of TB segments disappeared completely after reaction due to the fragile aliphatic methylene, in well accordance with the results by FT-IR. CTF-TB-7 exhibited remarkable thermal stability up to 800 °C in N2 (Fig. S4). The initial weight loss before 100 °C could be attributed to the absorbed moisture. The slightly weight loss from 300 °C should be attributed to the further degradation of organic frameworks. Therefore, they have similar thermal stability compared with the reported MOPs with TB segments.7,8,24 There was no observable glass transition temperature within the range of 30–500 °C, except the endothermal peak according to the evaporation of adsorbed moisture below 100 °C. The PXRD patterns of CTF-TB-3 and CTF-TB-7 (Fig. S6) indicated no long-range structures. They were all amorphous with broad diffraction peaks at about 24.8°, which was in well accordance with the reported MOPs containing TB segments.8,22,26 The morphology was also characterized by FESEM as shown in Fig. 1c and S7, which revealed aggregated particles. In the HR-TEM images shown in Fig. 1d and S8, the alternating areas of light and dark contrast revealed their disordered porous structural natures.


image file: c6ra21196h-f1.tif
Fig. 1 (a) FT-IR spectra of the monomer DCTB, CTF-TB-S and CTF-TB-3 obtained by KBr pellets, in the range of 4000–400 cm−1; (b) solid-state 13C NMR spectra of DCTB, CTF-TB-3, and CTF-TB-7 (100 MHz); (c) FESEM image of CTF-TB-3; (d) HR-TEM of CTF-TB-3.

Porosity properties and gas uptake capacities of the CTF-TBs

The porous character of the CTF-TBs was studied by nitrogen adsorption–desorption experiments at 77.3 K. As shown in Fig. 2a, d, S9, S10 and S12, except CTF-TB-S and CTF-TB-6, all the isotherms showed a steep increase in adsorbed volume at low relative pressure, suggesting that these CTF-TBs were microporous materials. All the nine CTF-TBs show type I N2 sorption isotherms with Brunauer–Emmett–Teller (BET) surface areas ranging from 294 to 731 m2 g−1 (Table 2). The Langmuir surface areas of CTF-TBs were varied from 431 to 1077 m2 g−1. None of them showed remarkable hysteresis at the low pressure region in the N2 isotherms. Therefore, the dominating pores were micropores, which were in well accordance to the pore size distribution as shown in Fig. 2c, f, S11, and S13. For CTF-TB-S obtained in the solution polymerization, although the reaction extent was high according to the FT-IR analysis, the surface area was very small (Fig. S9). Compared with other CTFs catalyzed by TFSA, CTF-TB-S showed much lower surface area, maybe due to the formation of oligomers with low molecular weights.
image file: c6ra21196h-f2.tif
Fig. 2 Porosity properties of CTF-TBs. (a) Nitrogen adsorption and desorption isotherms at 77.3 K of CTF-TB-3; (b) the influence of molar ratio (DCTB to ZnCl2) on the surface areas; (c) pore size distributions (PSD) for CTF-TB-3 calculated by the NLDFT method; (d) nitrogen adsorption and desorption isotherms at 77.3 K of CTF-TB-7; (e) the influence of reaction temperature on the surface areas; (f) pore size distributions (PSD) for CTF-TB-7 calculated by the NLDFT method.
Table 2 Porosity data for the CTF-TBs from N2 isotherms collected at 77.3 K
Polymer code SABETa [m2 g−1] SALangb [m2 g−1] SAmicroc [m2 g−1] SAextd [m2 g−1] V0.1e [m3 g−1] Vtotf [m3 g−1] V0.1/Vtot
a BET surface area calculated over the pressure range 0.05–0.30 P/P0 at 77.3 K.b Langmuir specific surface area calculated from the nitrogen adsorption isotherm by application of the Langmuir equation.c Micropore surface area calculated from the nitrogen adsorption isotherm using the t-plot method.d The external surface areas calculated using the t-plot method based on the Halsey thickness equation.e V0.1, pore volume at P/P0 = 0.1 at 77.3 K.f Vtot, total pore volume calculated at P/P0 = 0.99 at 77.3 K.g Not detected any reasonable data.
CTF-TB-1 294 431 243 51 0.146 0.155 0.942
CTF-TB-2 446 653 302 144 0.212 0.244 0.869
CTF-TB-3 612 901 338 274 0.283 0.349 0.811
CTF-TB-4 581 858 293 288 0.265 0.329 0.805
CTF-TB-5 495 721 257 238 0.225 0.295 0.763
CTF-TB-6 10 17 17 g 0.007 0.009 0.778
CTF-TB-7 731 1077 488 244 0.350 0.415 0.843
CTF-TB-8 689 1014 486 203 0.332 0.373 0.890
CTF-TB-9 625 919 424 202 0.298 0.343 0.868
CTF-TB-S 11 16 4 g g g g


The influence of the amount of ZnCl2, as well as reaction temperatures on the properties of the materials was investigated. First, the influence of the DCTB/ZnCl2 ratio was investigated under standard conditions at 400 °C. It was found that the amount of ZnCl2 had obvious effect on the pore properties of CTF-TBs, because it played the role of both solvent and catalyst in the reaction. When the ratios increased from 1[thin space (1/6-em)]:[thin space (1/6-em)]1 to 1[thin space (1/6-em)]:[thin space (1/6-em)]20, the BET surface areas first increased and then decreased gradually, as shown in Fig. 2b, which was in good agreement with that of dicyanobisphenyl (DCBP) and dicyanobenzimidazole (DCBI) under similar conditions.36,42 For CTF-TB-3, where the ratio of DCTB to ZnCl2 was 1[thin space (1/6-em)]:[thin space (1/6-em)]5, the BET surface area was 612 m2 g−1. This value was much lower than that of CTF-1 derived from 1,4-dicyanobenzene (DCB, 1123 m2 g−1) and dicyanobenzimidazole (DCBI, 1025 m2 g−1),27,36 under similar conditions. The main reason should be the serious decomposition of the unstable TBs under high temperatures. Further increase of ZnCl2 amount lead to decrease in BET surface areas may be due to the dilutive effect of excess amount of solvents.

It was revealed that higher reaction temperatures generate mesoporosity in addition to microporosity.27 Therefore, a series of temperatures from 350 °C to 550 °C was checked. When reaction temperatures increased, the BET surface areas first increased gradually and then decreased slightly as shown in Fig. 2e. For CTF-TB-7 obtained at 450 °C, the BET surface area was higher than that of CTF-TB-3 (731 m2 g−1 vs. 612 m2 g−1). However, compared with the reported CTFs with high surface areas, for example, CTF-0 derived from 1,3,5-tricyanobenzene (TCB, 2011 m2 g−1) and CTF-1 derived from DCB (1750 m2 g−1) under similar conditions,28,43 the present value was much lower. Furthermore, higher temperatures lead to some side reactions such as carbonizations, decompositions and ring fragmentations, as evidenced by FT-IR and solid-state 13C NMR analysis. In the present system, both the triazine rings and the TBs could decompose under higher temperatures to a relatively high extent, leading to partially collapse of pores. The obviously different tendency of BET surface areas vs. reaction temperatures can be mainly ascribed to the relatively lower thermal stability of TB segments, compared with benzene rings in CTF-1. According to results obtained by A. Thomas, the mesoporous pores mainly derived from a retrotrimerization/recombination pathway.28 If the TB segments were destroyed by the high temperatures, they cannot reorganize to form porous structures, thus leading to decrease in surface areas.

The gas storage properties of the CTF-TBs were tested. There are two kinds of CO2-philic groups in the CTF-TBs, e.g., the TB segments which contained tertiary amine and the triazine groups. Therefore, it can be predicted that the CTF-TBs would have high CO2 adsorption ability. We tested the CO2 adsorption properties at 273 K and a pressure up to 1.10 bar (Fig. 3a and b). The adsorption data were summarized in Table 3, from which we can see that all CTF-TBs showed moderate to high capacity of CO2 adsorption, consistent with our expecting. The CO2 adsorptions for all CTF-TBs ranged from 10.67 wt% to 16.84 wt% at 273 K and 1.10 bar. However, the CO2 adsorptions were not in good agreement with their BET surface areas, which was quite different from the results obtained in the CTF-BIs. CTF-TB-3 had the highest CO2 adsorption amount at 273 K and 1.10 bar (about 16.84 wt%). This adsorption amount was in the range of the highest values observed for CTFs under similar conditions. For example, CTF-1 (10.87 wt%, 746 m2 g−1) and CTF-1-600 (16.81 wt%, 1553 m2 g−1) based on DCB,34 CTF-P6M (18.35 wt%, 947 m2 g−1) prepared by microwave method,29 MCTF@500 (13.90 wt%, 640 m2 g−1) containing metalloporphyrin segments,44 PCTF-1 (14.21 wt%, 2235 m2 g−1) derived from tetranitrile containing tetraphenyl ethylene segments,45 and TPI-1 (10.78 wt%, 809 m2 g−1) containing imide groups,46 etc. Compared with the above mentioned CTFs, the present values showed comparable or better adsorption ability with relatively lower BET surface areas (CTF-TB-3, 16.2 wt% at 1.0 bar, 612 m2 g−1), which can be ascribed to the introduction of CO2-philic TB segments into the frameworks. However, compared with the recently reported FCTF-1-600 (24.33 wt%, 1535 m2 g−1),34 bipy-CTF600 (24.55 wt%, 2479 m2 g−1),35 and CTF-BI-11 (21.7 wt%, 1549 m2 g−1),36 the present result was still much lower. On the other hand, compared with the reported MOPs containing TBs, the present CTF-TBs showed comparable adsorption ability with relatively lower BET surface area under similar conditions. For instance, TB-MOP (17.8 wt%, 694 m2 g−1) with triptycene segments,7 and TB-MOP (16.89 wt%, 802 m2 g−1) derived from tris(4-aminophenyl)amine.12 However, compared with TB-COP-1 (22.8 wt%, 1340 m2 g−1) derived from tetraanilyladamantane,8 the present CTF-TBs didn't show any advantage. The main reason would be ascribed to the partially decomposition and fragmentations of TB segments, which lead to sharply decrease of both surface areas and TB concentrations in the final CTF-TBs.


image file: c6ra21196h-f3.tif
Fig. 3 (a) CO2 adsorption isotherms at 273 K of CTF-TBs; (b) CO2 adsorption isotherms at 303 K of CTF-TBs; (c) the ratios of four N-configurations (N(x)) in CTF-TBs; (d) heat of adsorption for CTF-TB-3 and CTF-TB-7.
Table 3 CO2 and N2 capture capacities, CO2/N2 selectivities for the CTF-TBs
Polymer codea SABET (m2 g−1) CO2 adsorption (cm3 g−1) N2 adsorption (cm3 g−1) CO2/N2c
273 Kb 303 K 273 K 303 K 273 K 303 K
a The CO2 adsorptions of CTF-TB-6 and CTF-TB-S were not measured due to the very low BET surface areas.b CO2 adsorption of CTF-TBs at 273 K and 1.10 bar.c Selectivity estimated using the ratios of the Henry law constant calculated from the initial slopes of the single-component gas adsorption isotherms at low pressure coverage (<100 mbar).
CTF-TB-1 294 54.94 33.57 4.07 1.25 44.3 50.0
CTF-TB-2 446 65.92 43.46 5.34 2.56 41.0 52.2
CTF-TB-3 612 85.77 55.13 6.69 2.11 43.4 48.1
CTF-TB-4 581 61.37 37.65 4.77 2.49 34.4 39.0
CTF-TB-5 495 54.35 33.56 2.74 2.10 53.3 39.5
CTF-TB-7 732 78.59 50.62 6.96 3.72 31.0 35.4
CTF-TB-8 689 66.26 39.96 4.34 2.70 42.5 32.5
CTF-TB-9 626 66.47 41.01 5.66 1.77 34.1 41.0


As mentioned above, the CO2 adsorption ability did not show a positive proportional relationship with the BET surface areas. CTF-TB-7 had a higher BET surface area than that of CTF-TB-3 (732 m2 g−1 vs. 612 m2 g−1), but had a lower CO2 adsorption amount (15.43 wt% vs. 16.84 wt%). The main reason should be again ascribed to the partially decomposition and fragmentations, as mentioned above. From the elemental analysis results (Table S1), as the reaction temperature increased, the loadings of N decreased gradually from 11.82 wt% (350 °C) to 8.52 wt% (550 °C). It should be evidence of the ring fragmentation. X-ray photoelectron spectroscopy (XPS) analysis of CTF-TBs were performed in order to further explore the origin of the effect of reaction temperatures on the CO2 adsorption ability. The N 1s XPS results of DCTB and CTF-TBs (Fig. S14) reveal that about three to four distinct N configurations exist in the skeletons of CTF-TBs: pyridinic N (397.98–398.40 eV, N(1)), pyrrolic N (399.60–399.99 eV, N(2)), quaternary N (400.86–401.19 eV, N(3)), and N-oxide (402.65–403.94 eV, N(4)).47,48 For CTF-TB-6 (350 °C) and CTF-TB-3 (400 °C), there were three distinct N configurations while for CTF-TB-7 (450 °C), CTF-TB-8 (500 °C), and CTF-TB-9 (550 °C), there were four, also indicated the decomposition or fragmentation reactions may occur under high temperature conditions. The atomic ratios of the four N-configurations in CTF-TBs were further evaluated from the deconvolution of N 1s peaks, and the results are plotted as a function of synthesis temperature in Fig. 3c. In general, the ratio of N(1) and N(2) tend to decrease when the synthesis temperature increases; while the ratios of N(3) and N(4) increase with increasing synthesis temperature. The high CO2 uptakes by nitrogen containing polymers were found to arise from strong interactions of the polarizable CO2 molecules and N atoms with Lewis basic sites. The interactions between CO2 and quaternary N and N-oxide were much weaker than those of pyridinic and pyrrolic N. Although CTF-TB-7 had a much higher BET surface area, the content of quaternary N and N-oxide were also much higher than that of CTF-TB-3 (24.5% vs. 9.2%, Fig. S14), thus leading to a lower CO2 uptakes.

To determine the binding affinity of CTF-TBs for CO2, we calculated the isosteric heat of adsorption (Qst) for CO2 using the Clausius–Clapeyron equation, as shown in Fig. 3d and S15. The Qst values of CTF-TBs were in the range of 27.0–30.4 kJ mol−1, which were well below the values expected for a chemisorption process (>40 kJ mol−1). Therefore, the high CO2 affinity of the CTF-TBs was mainly attributed to the inherent microporosity and the enhanced dipole–quadrupole interactions between the quadrupole moment of the CO2 molecules and the nitrogen-rich polar binding sites, both from TB and triazine rings.

CO2 and N2 sorption properties were measured by volumetric methods at the same conditions of 273 K and 303 K, respectively. The selectivity of CO2 over N2 was estimated according to a reported method by using the Henry law constants (Fig. S16). For all CTF-TBs, the calculated CO2/N2 adsorption selectivities were in the range of 31.0–53.3 at 273 K, and 32.5–52.2 at 303 K, respectively (Table 3 and Fig. S17). The selectivities were comparable to most of reported CTFs, for instance, TPC-1 with F atoms (selectivity for CO2/N2 at 273 K: 38),49 Cz-POFs with carbazoles (selectivity for CO2/N2 at 273 K: 19–37),50 fl-CTFs (selectivity for CO2/N2 at 273 K: 13–35),40 which were calculated by the same method. However, compared with the highly selective MOPs for CO2 adsorptions, e.g. the azo-COPs with multiple azo groups (selectivity for CO2/N2: 73–124 at 273 K, 113–142 at 298 K),4 these values herein were much lower.

Conclusions

In conclusion, a novel dicyano monomer containing Troger's base segments (DCTB) was designed and synthesized. A series of CTFs with TB segments (CTF-TBs) were prepared by the dynamic trimerization under ionothermal condition. The CTF-TBs exhibited BET surface areas up to 732 m2 g−1. The reaction conditions including molar ratio of ZnCl2 to DCTB and temperature had significant impact on the BET surface areas. The capacity of CO2 adsorption for all the CTF-TBs was studied and the CO2 uptake capacity was up to 16.84 wt% at 273 K and 1.10 bar for CTF-TB-3, obtained at the ratio of DCTB to ZnCl2 of 1[thin space (1/6-em)]:[thin space (1/6-em)]5 at 400 °C for 40 hours. This made these materials potential candidates for applications in CO2 capture and storage technology.

Acknowledgements

The authors thank the Fundamental Research Funds for the Central Universities (No. lzujbky-2014-72) and National Natural Science Foundation of China (No. 51103168, 21201093) for financial support.

Notes and references

  1. D. M. D'Alessandro, B. Smit and J. R. Long, Angew. Chem., Int. Ed., 2010, 49, 6058–6082 CrossRef PubMed .
  2. M. E. Boot-Handford, J. C. Abanades, E. J. Anthony, M. J. Blunt, S. Brandani, N. Mac Dowell, J. R. Fernandez, M. C. Ferrari, R. Gross, J. P. Hallett, R. S. Haszeldine, P. Heptonstall, A. Lyngfelt, Z. Makuch, E. Mangano, R. T. Porter, M. Pourkashanian, G. T. Rochelle, N. Shah, J. G. Yao and P. S. Fennell, Energy Environ. Sci., 2014, 7, 130–189 CAS .
  3. C. Xu and N. Hedin, Mater. Today, 2014, 17, 397–403 CrossRef CAS .
  4. H. A. Patel, J. S. Hyun, J. Park, D. P. Chen, Y. Jung, C. T. Yavuz and A. Coskun, Nat. Commun., 2013, 4, 1357–1364 CrossRef PubMed .
  5. L. M. Tao, F. Niu, D. Zhang, J. G. Liu, T. M. Wang and Q. H. Wang, RSC Adv., 2015, 5, 96871–96878 RSC .
  6. J. Lu and J. Zhang, J. Mater. Chem. A, 2014, 2, 13831–13834 CAS .
  7. X. Zhu, C. L. Do-Thanh, C. R. Murdock, K. M. Nelson, C. Tian, S. Brown, S. M. Mahurin, D. M. Jenkins, J. Hu, B. Zhao, H. Liu and S. Dai, ACS Macro Lett., 2013, 2, 660–663 CrossRef CAS .
  8. J. Byun, S. H. Je, H. A. Patel, A. Coskun and C. T. Yavuz, J. Mater. Chem. A, 2014, 2, 12507–12512 CAS .
  9. M. Carta, M. Croad, J. C. Jansen, P. Bernardo, G. Clarizia and N. B. Mckeown, Polym. Chem., 2014, 5, 5255–5261 RSC .
  10. M. Carta, M. Croad, R. Malpass-Evans, J. C. Jansen, P. Bernardo, G. Clarizia, K. Friess, M. Lanc and N. B. McKeown, Adv. Mater., 2014, 26, 3526–3531 CrossRef CAS PubMed .
  11. M. Carta, R. Malpass-Evans, M. Croad, Y. Rogan, M. Lee, I. Rose and N. B. McKeown, Polym. Chem., 2014, 5, 5267–5272 RSC .
  12. Z. Z. Yang, H. Y. Zhang, B. Yu, Y. F. Zhao, G. P. Ji and Z. M. Liu, Chem. Commun., 2015, 51, 1271–1274 RSC .
  13. S. Altarawneh, S. Behera, P. Jena and H. M. El-Kaderi, Chem. Commun., 2014, 50, 3571–3574 RSC .
  14. A. K. Sekizkardes, S. Altarawneh, Z. Kahveci, T. İslamoğlu and H. M. El-Kaderi, Macromolecules, 2014, 47, 8328–8334 CrossRef CAS .
  15. A. K. Sekizkardes, T. Islamoglu, Z. Kahveci and H. M. El-Kaderi, J. Mater. Chem. A, 2014, 2, 12492–12500 CAS .
  16. V. Guillerm, L. J. Weselinski, M. Alkordi, M. I. H. Mohideen, Y. Belmabkhout, A. J. Cairns and M. Eddaoudi, Chem. Commun., 2014, 50, 1937–1940 RSC .
  17. Q. Chen, M. Luo, P. Hammershøj, D. Zhou, Y. Han, B. W. Laursen, C. G. Yan and B. H. Han, J. Am. Chem. Soc., 2012, 134, 6084–6087 CrossRef CAS PubMed .
  18. S. Kumar, M. Y. Wani, C. T. Arranja, J. A. Silva, B. Avula and A. J. F. N. Sobral, J. Mater. Chem. A, 2015, 3, 19615–19637 CAS .
  19. X. Du, Y. Sun, B. Tan, Q. Teng, X. Yao, C. Su and W. Wang, Chem. Commun., 2010, 46, 970–972 RSC .
  20. M. Carta, M. Croad, K. Bugler, K. J. Msayib and N. B. McKeown, Polym. Chem., 2014, 5, 5262–5266 RSC .
  21. B. Ghanem, N. Alaslai, X. H. Miao and I. Pinnau, Polymer, 2016, 96, 13–19 CrossRef CAS .
  22. Y. Zhuang, J. G. Seong, Y. S. Do, W. H. Lee, M. J. Lee, M. D. Guiver and Y. M. Lee, J. Membr. Sci., 2016, 504, 55–65 CrossRef CAS .
  23. Y. Zhuang, J. G. Seong, Y. S. Do, W. H. Lee, M. J. Lee, Z. Cui, A. E. Lozano, M. D. Guiver and Y. M. Lee, Chem. Commun., 2016, 52, 3817–3820 RSC .
  24. I. Rose, M. Carta, R. Malpass-Evans, M. C. Ferrari, P. Bernardo, G. Clarizia, J. C. Jansen and N. B. McKeown, ACS Macro Lett., 2015, 4, 912–915 CrossRef CAS .
  25. M. Carta, R. Malpass-Evans, M. Croad, Y. Rogan, J. C. Jansen, P. Bernardo, F. Bazzarelli and N. B. McKeown, Science, 2013, 339, 303–307 CrossRef CAS PubMed .
  26. Z. Wang, D. Wang, F. Zhang and J. Jin, ACS Macro Lett., 2014, 3, 597–601 CrossRef CAS .
  27. P. Kuhn, M. Antonietti and A. Thomas, Angew. Chem., Int. Ed., 2008, 47, 3450–3453 CrossRef CAS PubMed .
  28. P. Kuhn, A. Forget, D. Su, A. Thomas and M. Antonietti, J. Am. Chem. Soc., 2008, 130, 13333–13337 CrossRef CAS PubMed .
  29. S. Ren, M. J. Bojdys, R. Dawson, A. Laybourn, Y. Z. Khimyak, D. J. Adams and A. I. Cooper, Adv. Mater., 2012, 24, 2357–2361 CrossRef CAS PubMed .
  30. S. Dey, A. Bhunia, D. Esquivel and C. Janiak, J. Mater. Chem. A, 2016, 4, 6259–6263 CAS .
  31. K. Wang, H. Huang, D. Liu, C. Wang, J. Li and C. Zhong, Environ. Sci. Technol., 2016, 50, 4869–4876 CrossRef CAS PubMed .
  32. X. Wang, C. Zhang, Y. Zhao, S. Ren and J. Jiang, Macromol. Rapid Commun., 2016, 37, 323–329 CrossRef CAS PubMed .
  33. S. Kuecken, J. Schmidt, L. Zhi and A. Thomas, J. Mater. Chem. A, 2015, 3, 24422–24427 CAS .
  34. Y. Zhao, K. X. Yao, B. Teng, T. Zhang and Y. Han, Energy Environ. Sci., 2013, 6, 3684–3692 CAS .
  35. S. Hug, L. Stegbauer, H. Oh, M. Hirscher and B. V. Lotsch, Chem. Mater., 2015, 27, 8001–8010 CrossRef CAS .
  36. L. Tao, F. Niu, C. Wang, J. Liu, T. Wang and Q. Wang, J. Mater. Chem. A, 2016, 4, 11812–11820 CAS .
  37. D. Didier and S. Sergeyev, Tetrahedron, 2007, 63, 3864–3869 CrossRef CAS .
  38. K. Sakaushi, G. Nickerl, F. M. Wisser, D. Nishio-Hamane, E. Hosono, H. Zhou, S. Kaskel and J. Eckert, Angew. Chem., Int. Ed., 2012, 51, 7850–7854 CrossRef CAS PubMed .
  39. L. Tao, F. Niu, D. Zhang, T. Wang and Q. Wang, New J. Chem., 2014, 38, 2774–2777 RSC .
  40. S. Hug, M. B. Mesch, H. Oh, N. Popp, M. Hirscher, J. Senker and B. V. Lotsch, J. Mater. Chem. A, 2014, 2, 5928–5936 CAS .
  41. K. A. See, S. Hug, K. Schwinghammer, M. A. Lumley, Y. Zheng, J. M. Nolt, G. D. Stucky, F. Wudl, B. V. Lotsch and R. Seshadri, Chem. Mater., 2015, 27, 3821–3829 CrossRef CAS .
  42. M. J. Bojdys, J. Jeromenok, A. Thomas and M. Antonietti, Adv. Mater., 2010, 22, 2202–2205 CrossRef CAS PubMed .
  43. P. Katekomol, J. Roeser, M. Bojdys, J. Weber and A. Thomas, Chem. Mater., 2013, 25, 1542–1548 CrossRef CAS .
  44. X. Liu, H. Li, Y. Zhang, B. Xu, A. Sigen, H. Xia and Y. Mu, Polym. Chem., 2013, 4, 2445–2448 RSC .
  45. A. Bhunia, V. Vasylyeva and C. Janiak, Chem. Commun., 2013, 49, 3961–3963 RSC .
  46. M. R. Liebl and J. Senker, Chem. Mater., 2013, 25, 970–980 CrossRef CAS .
  47. L. Hao, B. Luo, X. Li, M. Jin, Y. Fang, Z. Tang, Y. Jia, M. Liang, A. Thomas, J. Yang and L. Zhi, Energy Environ. Sci., 2012, 5, 9747–9751 CAS .
  48. C. E. Chan-Thaw, A. Villa, G. M. Veith, K. Kailasam, L. A. Adamczyk, R. R. Unocic, L. Prati and A. Thomas, Chem.–Asian J., 2012, 7, 387–393 CrossRef CAS PubMed .
  49. X. M. Hu, Q. Chen, Y. C. Zhao, B. W. Laursen and B. H. Han, J. Mater. Chem. A, 2014, 2, 14201–14208 CAS .
  50. X. Zhang, J. Lu and J. Zhang, Chem. Mater., 2014, 26, 4023–4029 CrossRef CAS .

Footnote

Electronic supplementary information (ESI) available: Structure characterization of CTF-TBs; the nitrogen adsorption and desorption isotherms and pore size distributions of all CTF-TBs; the CO2 adsorption isotherms, and the CO2/N2 selectivities of all CTF-TBs. See DOI: 10.1039/c6ra21196h

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.