Shuangshuang Houa,
Shaolei Wanga,
Xuejun Longb and
Bien Tan*a
aKey Laboratory for Large-Format Battery Materials and System, Ministry of Education, Hubei Key Laboratory of Material Chemistry and Service Failure, School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology, Wuhan, 430074, P. R. China. E-mail: bien.tan@mail.hust.edu.cn; Fax: +86 27 87558172; Tel: +86 27 87558172
bEngineering Research Center for Clean Production of Textile Printing and Dyeing, Ministry of Education, Wuhan Textile University, Wuhan 430073, P. R. China
First published on 13th March 2018
In order to achieve efficient CO2 capture, four novel microporous organic polymers, based on distinct polycyclic aromatic hydrocarbons such as fluoranthene, binaphthalene, naphthalene and phenanthrene, were successfully prepared by the solvent knitting method. N2 sorption isotherms indicate that these polymers are predominately microporous with ultrahigh BET surface area i.e., 1788 m2 g−1 for fluoranthene-based Polymer 1, 1702 m2 g−1 for binaphthalene-based Polymer 2 and objective CO2 uptake capacity of 24.79 wt% and 20.19 wt% (273.15 K/1.00 bar) respectively. While compared with the former two polymers, though 1227 m2 g−1 and 978 m2 g−1 are moderate in surface area, however the naphthalene-based Polymer 3 and phenanthrene-based Polymer 4 still exhibit CO2 adsorption of up to 17.44 wt% and 18.15 wt% respectively under the similar conditions. Moreover, the H2 storage and CH4 adsorption in these polymers can be 2.20 wt% (77.3 K/1.13 bar) and 2.79 wt% (273.15 K/1.00 bar). More significantly, the electron-rich PAHs are proved to be new building blocks that provide a wealth of chances to produce hypercrosslinked polymers with efficient gas adsorption capacity, which are greatly influenced by the porous nature of polymers. Given the merits including mild reaction conditions, low cost, high surface area, impressive gas absorption performance, high thermal stability, these polymers are considered to be promising candidates for CO2 capture and energy storage under more practical conditions.
Emerged in the past few decades, hypercrosslinked polymers (HCPs) with many interesting intrinsic features such as large specific surface area, low skeleton density and narrow pore size distribution have exhibited great potential for CO2 uptake.12 Bearing the motivations of diverse arts, many relative contributions upon the capture of CO2 in HCPs have been successively unfolded. For example, Cooper et al. reported a high surface area network of 1015 m2 g−1 with CO2 capture capacity up to 3.96 mmol g−1 at 273.15 K/1.00 bar.13 Our group demonstrated a cobalt coordinated polymer with a high BET surface area 1360 m2 g−1 possessing CO2 adsorption of 21.39 wt% at 273.15 K/1.00 bar.14 Jiang and coworkers prepared a nitrogen-rich material (FCDTPA) with surface area of 871 m2 g−1 and CO2 uptake ability up to 2.83 mmol g−1 at 273.15 K/1.00 bar.15
Usually, a careful selection of building blocks is a necessary step before producing ideal HCPs. Aromatic building blocks such as benzene, biphenyl and 1,3,5-triphenylbenzene have been commonly used to produce HCPs.16,17,28 Due to the absence of polar groups or heteroatoms, many reported HCPs, prepared from aromatic building blocks, lack strong CO2 binding sites and unfortunately tend to show a low or moderate CO2 adsorption.16,18,19 Heterocyclic compounds that consist of nitrogen, oxygen and sulfur have also been widely investigated in this regard.12,20–22 Attributed to the lone pair electrons of heteroatoms, which are playing a crucial role in offering interaction sites through dipole–dipole interactions,23 heterocyclic network could exhibit enhanced CO2 adsorption, and this value is similar to that of polymers generated from benzene, though the surface area is much lower than that of the later.16 Moreover, simulations based on employing density functional theory (DFT) calculations have further indicated that with the largest binding energy towards CO2, more negative charge distribution over moieties of a precursor could help in improving the interactions with CO2 molecules.24 So it is not hard to speculate that, apart from aromatic building blocks and heterocyclic compounds, the building units rich in electron are still probably better choices to produce HCPs with high CO2 capture capacity.
Polycyclic aromatic hydrocarbons (PAHs) are a big family of organic compounds with two or more laterally fused benzene rings and produced by combustion processes involving carbonaceous fuels. Owing to the structurally composed π conjugated systems, they are rich in electron and more likely to be promising building blocks for creating novel HCPs.25 However, PAHs have gathered growing concerns, because they are widely distributed environmental containment, do not degrade easily under natural conditions and often have detrimental biological effects, toxicity, mutagenicity and carcinogenicity.26,27 Considering their ubiquitous occurrence in all components of environment, recalcitrance, bioaccumulation potential and carcinogenic activity, controlling the emission of PAHs has become more urgent and far-reaching.26,27
What is more, it should be noted that, for the benefit of fabricating original HCPs with fascinating structures and multifunctional applications, a series of synthetic approaches such as the solvent knitting method,28 the knitting method with FDA as external crosslinker16 and the Scholl coupling reaction17 have been creatively reported. While in these strategies, the latest solvent knitting method, based on the Friedel–Crafts alkylation reaction mechanism, concerning the employment of dichloroalkane as both economical solvent and external crosslinker, was recently put forward by our group. Apart from distinguishing characteristics as simple, one-step, cost-effective, it is also something of an innovation in the constraint of methodology with respect to providing a wealth of splendid opportunities for polymer networks with higher surface area, narrow pore size distribution, good gas adsorption performance and so on, thus being considered to be a reliable polymerization technique.28
We, therefore, set out to use low cost PAHs such as fluoranthene, binaphthalene, naphthalene and phenanthrene aim to knit microporous organic polymers for efficient gas adsorption. The fluoranthene-based polymer of all the four synthesized samples reveals the highest surface area of 1788 m2 g−1 and the best CO2 uptake property of 24.79 wt% (273.15 K/1.00 bar), H2 storage ability of 2.20 wt% (77.3 K/1.13 bar) and CH4 adsorption capacity of 2.79 wt% (273.15 K/1.00 bar). In addition, it is noteworthy that the electron-rich PAHs have been found to be novel building blocks for the production of microporous polymers with high gas adsorption properties. The most plausible reasons for the distinct structural properties and higher gas uptake capacity of these polymers may largely be attributed to their porous nature including the pore size, pore size distribution as well as BET surface area, which is expected to give more insight to produce microporous with desirable properties.
Fig. 1 Cross-polarization (CP) 13C MAS natural abundance NMR spectra of materials from Polymer 1 to Polymer 4. |
The surface morphologies and inherent porous structures of all polymers were examined by electron microscopy. Fig. 2 shows the FE-SEM images of Polymer 1 to Polymer 4. All four polymer samples are irregular blocks without any long-range order, and in comparison, Fig. S1† gives the FE-SEM images of corresponding building units at different magnifications. Fig. 2 also shows the TEM images of Polymer 1 to Polymer 4 at the same magnification. All four polymer samples have hierarchically porous structure with abundant micropores and a small amount of mesopores. The TGA analysis curves in Fig. S10† indicated that all the four polymer networks have similar decomposition behaviors except Polymer 1 in a high temperature region when the temperature is more than 450 °C, which is mainly caused by different crosslinking degree of the networks. The significant weight loss in a high temperature range over 400 °C under a nitrogen atmosphere is the result of the destruction of polymeric networks, which importantly verifies the good chemical and thermal stability of these polymers.
The porosity of polymers were determined by nitrogen adsorption and desorption isotherms at 77.3 K. As shown in Fig. 3a, all the adsorption isotherms exhibited a type I character with a steep nitrogen gas uptake at low relative pressure (P/P0 < 0.001), implying the adsorption into abundant micropores, which is in good agreement with the results of TEM. It is fairly obvious that all the microporous networks are accompanied with primary micropores, and bits of mesopores that are reflected by almost no obvious hysteresis loop in the middle pressure area. In Fig. 3b, the pore size distribution with a large pore volume calculated by the NLDFT model indicated that all the four polymers possess hierarchical pore distribution, and the pore size distribution principally covers a very narrow range no more than 2 nm in microporous religion, including the ultramicropore whose size is less than 0.7 nm in diameter. Table 1 presents more details about these polymers that the total pore volumes vary between 0.48 cm3 g−1 and 0.86 cm3 g−1 with micropore volume from 0.27 cm3 g−1 to 0.53 cm3 g−1 as well as the t-plot microporous area over BET surface area is within the scope of 65.57% to 75.89%. Based on the above analysis, it can be concluded that the solvent knitting method could bridge various building units to generate microporous organic materials under mild conditions.
Samples | SBETa (m2 g−1) | SLb (m2 g−1) | MAc (m2 g−1) | PVd (cm3 g−1) | MPVe (cm3 g−1) | MAf (%) |
---|---|---|---|---|---|---|
a Apparent surface area calculated from nitrogen adsorption isotherms at 77.3 K using the BET equation.b Surface area calculated from nitrogen adsorption isotherms at 77.3 K using the Langmuir equation.c t-Plot micropore area.d Pore volume calculated from the nitrogen isotherms at P/P0 = 0.995 and 77.3 K.e t-Plot micropore volume calculated from the nitrogen isotherms at P/P0 = 0.050.f t-Plot microporous area/BET surface area × 100%. | ||||||
Polymer 1 | 1788 | 2102 | 1357 | 0.82 | 0.53 | 75.89 |
Polymer 2 | 1702 | 2063 | 1116 | 0.86 | 0.44 | 65.57 |
Polymer 3 | 1227 | 1490 | 807 | 0.63 | 0.32 | 65.77 |
Polymer 4 | 978 | 1173 | 684 | 0.48 | 0.27 | 69.94 |
The interesting physical properties of these knitted polymers, as listed in Table 1, including their high surface area and microporous nature prompted us to pay more attention to their gas uptake performance. Taking together the CO2 isotherms in Fig. 4a and the gas uptake capacities in Table 2, it is obvious Polymer 1 has an ultrahigh BET surface area of 1788 m2 g−1 and the highest CO2 adsorption amount of 24.79 wt% (273.15 K/1.00 bar) compared to the Polymer 2, Polymer 3 and Polymer 4 whose BET surface areas are 1702 m2 g−1, 1227 m2 g−1 and 978 m2 g−1 with corresponding CO2 uptake capacity of 20.19 wt%, 17.44 wt% and 18.15 wt% (273.15 K/1.00 bar) respectively. It may be due to their much higher microporosity and ultramicropore whose diameter is comparable to the kinetic diameter of CO2 thus increasing the interaction between CO2 molecules and the pore walls.29 Although lower than the reported imine-linked porous polymer network PPF-1 (26.7 wt%, SBET = 1740 m2 g−1),30 the CO2 uptake of Polymer 1 is comparable with some reported porous polymers with high CO2 uptake performance such as carbazolic porous organic framework Cz-POF-3 (4.77 mmol g−1, SBET = 1927 m2 g−1),31 benzimidazole-linked polymer BILP-4 (5.3 mmol g−1, SBET = 1235 m2 g−1),32 azo-linked porous organic polymer ALP-1 (5.37 mmol g−1, SBET = 1235 m2 g−1),33 fluorinated covalent triazine-based framework FCTF-1-600 (5.53 mmol g−1, SBET = 1535 m2 g−1),34 and bipyridine-based nitrogen rich covalent triazine framework bipy-CTF600 (5.58 mmol g−1, SBET = 2479 m2 g−1).35 The CO2 capture capacity of Polymer 1 is higher than many CMPs produced by Sonogashira–Hagihara coupling reactions such as the pyrene-based porous aromatic framework of PAF-26 (1.16 mmol g−1, SBET = 702 m2 g−1),36 the amide-functionalized CMP of CMP-1-AMD1 (1.51 mmol g−1, SBET = 316 m2 g−1),37 the hexabenzocoronene-based porous organic polymers of HBC-POP-1 (2.05 mmol g−1, SBET = 668 m2 g−1),38 the tri(4-ethynylphenyl)amine-based porous aromatic framework of PAF-34 (2.50 mmol g−1, SBET = 953 m2 g−1)39 and the post-metalation of the porous aromatic framework of PAF-26-COOMg (2.85 mmol g−1, SBET = 572 m2 g−1)40 at 273.15 K/1.00 bar. Moreover, the capture of CO2 in Polymer 1 is also much higher than some types of porous materials with much higher BET surface area under the similar conditions such as the tetraphenylmethane-based HCP (1.66 mmol g−1, SBET = 1679 m2 g−1),41 COF-102 (1.56 mmol g−1, SBET = 3620 m2 g−1)42 and PAF-1 (2.1 mmol g−1, SBET = 5460 m2 g−1).43 In order to determine the binding affinity between the polymers and CO2 molecules, the isosteric heat (Qst) of the four obtained polymers toward CO2 was calculated from the adsorption isotherms at 273.15 K and 298.15 K using the Clausius–Clapeyron equation (Fig. S11†),44 which were found to be of 25.65–24.41 kJ mol−1, 26.40–24.82 kJ mol−1, 26.93–25.30 kJ mol−1 and 28.09–26.38 kJ mol−1 for Polymer 1, Polymer 2, Polymer 3 and Polymer 4 respectively.
Samples | CO2 uptakea (wt%) | CO2 uptakeb (wt%) | CH4 uptakec (wt%) | H2 uptaked (wt%) |
---|---|---|---|---|
a CO2 uptake determined volumetrically using a Micromeritics ASAP 2020 M analyzer at 1.00 bar and 273.15 K.b CO2 uptake determined volumetrically using a Micromeritics ASAP 2020 M analyzer at 1.00 bar and 298.15 K.c CH4 uptake determined volumetrically using a Micromeritics ASAP 2020 M analyzer at 1.00 bar and 273.15 K.d H2 uptake determined volumetrically using a Micromeritics ASAP 2020 M analyzer at 1.13 bar and 77.3 K. | ||||
Polymer 1 | 24.79 | 14.79 | 2.79 | 2.20 |
Polymer 2 | 20.19 | 11.69 | 2.00 | 1.82 |
Polymer 3 | 17.44 | 10.66 | 2.08 | 1.59 |
Polymer 4 | 18.15 | 11.23 | 2.00 | 1.57 |
As the increasing demand for energy and environmental concerns has made H2 storage much appealing in porous materials research, investigation of H2 uptake performance of polymers has become more and more attractive. As shown in Fig. 4c, the hydrogen adsorption and desorption measurements at low pressure were carried out at 77.3 K, and all the isotherms of polymers for H2 adsorption are fully reversible. Intriguingly, Polymer 1 has the highest H2 uptake of 2.20 wt% (77.3 K/1.13 bar) compared to the Polymer 2, Polymer 3 and Polymer 4 whose H2 uptake capacities are 1.82 wt%, 1.59 wt% and 1.57 wt% respectively under the similar conditions. Though lower than that of the CPOP-1 (2.80 wt%),45 zeolite-like Carbon (2.60 wt%),46 MIL-101(2.50 wt%)47 and carbon AX-21 (2.40 wt%),48 the H2 adsorption ability of Polymer 1 is still comparable to SPOP-3 (2.22 wt%)49 and PAF-3 (about 2.07 wt%)43 at 77.3 K/1.00 bar, and also much higher than that of many reported materials such as benzene-based CMP of CMP-0 (1.4 wt%),45 the nanoporous organic framework of NPOF-2 (1.45 wt%)50 as well as the tetraphenylethylene-based copolymer networks (0.95–1.76 wt%)51 under similar conditions.
Touted as an ideal energy resource, methane storage may be a serious issue in future energy schemes, so the methane storage capacity of the polymers is also worth exploring. As shown in Fig. 4d, Polymer 1 has a better methane adsorption of 2.79 wt% (273.15 K/1.00 bar) than that of Polymer 2, Polymer 3 and Polymer 4 whose methane storage abilities are 2.00 wt%, 2.08 wt% and 2.00 wt% (273.15 K/1.00 bar) respectively. It is easy to note that the methane storage of Polymer 1 is, though lower than the carbonized FCDTPA-K-700 (2.36 mmol g−1 at 273.15 K/1.13 bar),15 the activated carbon of K-PAF-1-600 (2.4 mmol g−1 at 273.15 K/1.00 bar) that stemmed from PAF-1,52 while still much higher than that of the reported FCDTPA (0.89 mmol g−1 at 273.15 K/1.13 bar),15 and the solution-processable hypercrosslinked polymers (0.08–0.14 wt% at 273.15 K/1.13 bar).53
Keeping in view different structure and gas capture ability of these polymers, the correlation of structural diversity and gas adsorption performance has come up as an interesting phenomenon. Various efforts have been made to seek for the best interpretation for such abnormal phenomenon, and in the most of cases it has been ascribed to the porous nature of such materials including the pore size and pore size distribution as well as surface area of polymers. Universally it is acknowledged that factors like slight variations in the chemical structure, aggregation structure, and porous properties including the specific surface area, micropore volume, pore shape, pore size and pore size distribution largely affect the CO2 adsorption capacity of microporous organic polymers.54 However, more convincing explanation can be directly derived from Fig. S12.† As evident, based on some selected microporous hypercrosslinked polymers,16,28 the relationship of BET surface area and CO2 uptake at 273.15 K/1.00 bar is vividly presented in a good fitting equation of y = 0.00676x + 3.8732 with correlation coefficient value R2 = 0.99728. According to the experimental results, the obtained polymers with corresponding surface area and CO2 uptake at 273.15 K/1.00 bar have been totally scattered at different places above the equation. So it is not hard to understand that, aside from the BET surface area and other possible factors like the chemical structure of monomers, aggregation structure, micropore volume, pore shape, the pore size and pore size distribution are extremely important for CO2 adsorption. The previous reports have indicated that the pore size less than 1 nm is useful to adsorb CO2 molecules,55 especially the ultramicropore with diameter no more than 0.7 nm, makes marvelous contributions to CO2 uptake, for the small pore size could favorably increase the interaction between the CO2 molecules and the pore walls of polymers.29 Notably, the PAHs with available π conjugated systems have obvious superiority in electron distribution, which have emerged as a superior class of versatile building blocks for establishing microporous organic polymers with high surface area and enhancing gas uptake ability. Furthermore, this finding not only largely gives some enlightenment on the rational design and synthesis of desirable polymers, but may also positively facilitate the considerable development of HCPs in many interdisciplinary domains in the future.
Footnote |
† Electronic supplementary information (ESI) available: SEM images and FT-IR spectra of building blocks. FT-IR spectra, TGA curves and Qst of polymers. The fitting relationship of CO2 uptake at 273.15 K/1.00 bar and BET surface area of polymers. See DOI: 10.1039/c8ra01332b |
This journal is © The Royal Society of Chemistry 2018 |