Muhammad Salehab and
Kwang S. Kim*b
aDepartment of Chemistry, Pohang University of Science and Technology, Pohang 790-784, Korea
bCentre for Superfunctional Materials, Department of Chemistry, Ulsan National Institute of Science and Technology (UNIST), Ulsan 689-798, Korea. E-mail: kimks@unist.ac.kr; Fax: +82-52-217-5419; Tel: +82-52-217-5410
First published on 30th April 2015
Non-coplanar shaped carbazole based monomers were used to synthesize microporous polycarbazole materials utilizing an inexpensive FeCl3 catalyzed reaction. The reactions proceed through direct oxidative coupling and extensive crosslinking polymerization routes. The obtained porous networks exhibit a maximum Brunauer–Emmett–Teller specific surface area of 946 m2 g−1 with a total pore volume of 0.941 cm3 g−1, and display a high carbon dioxide uptake capacity (up to 13.6 wt%) at 273 K and 1 atm. Selective adsorption of CO2 over N2 calculated using the ideal adsorbed solution theory (IAST) shows that these networks display enhanced selectivity with a maximum value of 155 at 298 K. Remarkably, in contrast to other materials, this value is significantly higher than the selectivity values (102–107) obtained at 273 K. Introduction of the electron rich carbazole structure into the aromatic system and pore geometry contribute to higher adsorption enthalpy which in turn leads to high selective adsorption values. These polymeric networks also show a high working capacity with reasonably high regenerability factors. The combination of a simple inexpensive synthesis approach and high selective adsorption make these materials potential candidates for CO2 storage, selective gas adsorption, and other environmental applications.
The gas storage capacity and separation capability of microporous organic polymers can be controlled by certain factors such as material nature, specific surface area, pore structure, and material's morphology.17 In this concern, non-coplanar monomers are of primary importance as the polymers they form have inherent porous properties. Furthermore, incorporation of bulky organic linkers with high connectivity can effectively prevent network interpenetration which leads to materials with enhanced texture properties. In this way, by designing molecules with certain porous geometries such as microporous organic polymers afford good performance for the gas storage and separation. Carbazole has got much attention and a lot of carbazole based polymeric materials have been synthesized already.18–22 By means of these new strategies, aromatic heterocyclic POFs with CO2-philic moieties and good porosity can be prepared without using harsh conditions. The carbazole based polymers have reasonably good gas uptake but have a drawback of poor gas selectivity. Recently, carbazole group has been reported to enhance the selectivity of traizines.18 In order to enhance the gas selectivity, we introduced carbazole on the selected building blocks with non-coplanar geometry molecule. In this way, by designing molecules with certain porous geometries such as microporous organic polymers afford good performance for the gas storage and separation. To the best of our understanding, no such carbaozle containing materials with remarkably high CO2/N2 selectivity at 298 K have been reported to date. Herein, we present the synthesis of new POFs for energy-efficient CO2 capture with favorable CO2 uptake and sufficiently high CO2/N2 selectivity.
Gas adsorption measurements were performed using a Belsorp mini II (Micromeritics, Japan) device. Before each measurement, a weighed sample was heat treated at 150 °C under vacuum for 16 hours. Brunauer–Emmett–Teller (BET) surface area was investigated using N2 adsorption–desorption isotherms measured at 77 K. Pore size distributions were calculated using non local density functional theory. CO2, CH4 and N2 storage adsorption isotherms were measured at 273 and 298 K and up to 1 atmosphere.
The structures of the polymers were characterized at the molecular level by 13C solid state NMR. The 13C CP-MAS NMR spectrum of the obtained polymer with the assignments of the resonances is shown in Fig. S3.† The representative spectra of C-SP, CL-SP, C-NP and CL-NP exhibit four signals for aromatic carbons at 100–150 ppm. The peak at about 139.9 ppm corresponds to the substituted phenyl carbons binding with carbazole nitrogen atom. The high-intensity peak for the other substituted phenyl carbons is located at about 125 ppm. The signal peak at about 120.8 and 109.8 ppm are attributed to the unsubstituted phenyl carbons. Additionally, for the CL-SP and CL-NP materials the presence of the methylene carbon signal at about 40 ppm confirms that the linkers have participated in the formation of the cross-linking porous polymers.
In a comparison of the FTIR spectra of the polymers, four main peaks are observed as shown in Fig. S4.† The peaks at about 1610 and 1510 cm−1 are attributed to aromatic ring skeletal vibrations, which were consistent with the structure of the corresponding monomers. The weak absorption peaks at about 2900–3000 cm−1 correspond to (C–H band stretching vibrations), while the additional broad peak at about 3450 cm−1 (O–H band stretching vibrations) may be ascribed to the incomplete removal of methanol even after drying.
TGA shows that all polymers decompose between 360–400 °C, indicating impressive thermal stability (Fig. S5†). This higher thermal stability of the polymers is desired in order to withstand application in the harsh power plant conditions.
The porosity of the polymers were evaluated by using nitrogen adsorption–desorption isotherms measured at 77 K (Fig. 1a), in which isotherms show rapid uptake at low relative pressures reflecting a permanent microporous nature of the networks. The isotherms for CL-NP and C-SP can be classified broadly as type I according to the IUPAC classification,24 while C-NP and CL-SP exhibit mixed type I and type II behavior. For all polymers, the isotherms show a continuous increase after the adsorption at low relative pressure (P/P0 < 0.01), indicating formation of monolayer followed by multi layers. The increase in the nitrogen adsorption at a high relative pressure (P/P0 > 0.9) may arise from the interparticulate porosity associated with the meso/macro-structures of the samples.25 Hysteresis is observed apparently for CL-SP and C-NP over the entire pressure range, which might be attributed to the swelling effects.26 The BET surface areas for the polymers calculated over a relative pressure (P/P0) range from 0.01 to 0.1 are 573–946 m2 g−1 (Table 1). C-NP and CL-SP show a higher surface area than CL-NP and C-SP, which is comparable to other recently developed microporous polymers.17,27 C-NP shows the largest micropore volume (0.504 cm3 g−1) of all the polymer networks, followed by CL-SP (0.465 cm3 g−1), C-SP (0.369 cm3 g−1) and CL-NP (0.319 cm3 g−1). The pore size distribution (PSD) curves (Fig. 1b) calculated using non-local density functional theory further suggest that the polymers have a narrow pore size distribution with the pore widths centered around 0.6 nm. These PSD curves are in good agreement with the shape of the nitrogen isotherms (Fig. 1a) strengthening the evidence that all the polymers have the dominant microporous character. Overall the textural properties observed for CL-SP are superior due to the presence of excessive crosslinking compared to the corresponding product C-SP. However, a reverse trend was obtained in the case of CL-NP when compared to C-NP. This observation may be due to the specific geometry of the NP which suggests that besides the selection of synthesis route, the non-coplanar geometry of the monomer is significantly important in obtaining a higher porous character.
![]() | ||
Fig. 1 (a) N2 adsorption–desorption isotherms and (b) pore size distribution for the polymeric networks. |
Network | SAa | Vtb | V0.1c | PDd | CO2e | CH4e | Selectivity (initial slope)f | Selectivity (IAST)g | ||||
---|---|---|---|---|---|---|---|---|---|---|---|---|
273 K (298 K) | Qst | 273 K (298 K) | Qst | CO2![]() ![]() |
CO2![]() ![]() |
CO2![]() ![]() |
CO2![]() ![]() |
|||||
a Specific surface area (m2 g −1) calculated from the nitrogen adsorption isotherm using the BET method.b Total pore volume (cm3 g−1) at P/P0 = 0.995.c Micropore volume (cm3 g−1).d Average pore diameter (nm).e Gas uptake at in (wt%) and the isosteric enthalpies of adsorption (Qst) in kJ mol−1.f Selectivity was calculated from initial slope calculations at 273 K and (298 K).g Selectivity was calculated from IAST for 15/85 gas mixtures for CO2/N2 and 5/95 gas mixtures for CO2/CH4 at 273 K (298 K). | ||||||||||||
C-NP | 946 | 0.941 | 0.504 | 3.98 | 13.6 (8.7) | 28.8 | 1.86 (1.09) | 20.3 | 29 (23) | 9.8 (10) | 102 (102) | 14.7 (14.2) |
CL-NP | 573 | 0.433 | 0.319 | 3.03 | 9.5 (5.5) | 27.2 | 1.18 (0.64) | 19.9 | 33 (36) | 10.2 (11.6) | 103 (121) | 15.1 (16.4) |
C-SP | 660 | 0.523 | 0.369 | 3.17 | 10.7 (5.8) | 26.2 | 1.26 (0.69) | 20.1 | 31 (39) | 10.3 (8.4) | 117 (127) | 16.0 (16.7) |
CL-SP | 865 | 0.735 | 0.465 | 3.40 | 13.1 (8.0) | 28.4 | 1.50 (0.94) | 19.2 | 36 (41) | 10.6 (11.3) | 107 (155) | 16.3 (18.3) |
Low-pressure gas sorption measurements for CO2 were collected in order to investigate the impact of the narrow pores on gas storage and the preferential binding of CO2 over CH4 and N2 at 273 and 298 K. Remarkably, C-NP and CL-SP show CO2 uptakes of 13.6 and 13.1 wt% at 273 K, respectively (Fig. 2a and Table 1). The CO2 adsorption capacity of these polymers is comparable to the many previously reported POFs.6,12,17,27–30 This further corroborates that both knitting and oxidative synthesis approaches are simple routes for the large scale manufacturing of polymer based CO2 adsorbents. On the basis of the CO2 adsorption isotherms, we found that the CO2 uptake capacity increases with an increase in the specific surface area. To investigate the impact of microporosity and the chemical nature of the pore surfaces of the as synthesized polymers on the CO2 uptake, we measured CO2 adsorption isotherms of polymers at 298 K. Utilizing this data we were able to calculate isosteric heats of adsorption by applying the Clausius–Clapeyron equation. At zero coverage, the isosteric heat of adsorption (Qst) calculated for these polymeric materials lie in 26–29 kJ mol−1, as shown in Fig. 2b. These values are in comparison to those reported for many other POFs.12,27–30 Higher Qst values for polymers indicate strong interactions of the polarizable CO2 molecules through dipole–quadrupole interactions with the N sites of framework due to the inherent microporosity of polymers. Additionally, the presence of narrow pores provides favorable multiple wall interactions that raise the adsorption potential of CO2. The change in Qst with an increase in CO2 loading may be attributed to the differences in textural and chemical properties of the polymers. Although nitrogen sites of arylamines are less reactive than alkyl amines (i.e. porous polymer network containing frameworks),16 they show a reasonably high affinity for CO2 which is evident from the observed large Qst values. We can say that high Qst values are due to the integrated basic effects of arylamines and pore geometry. The readily reversible adsorption–desorption process and moderate Qst values (below the chemisorptive process limit) indicate that CO2 interactions with pore walls are weak enough to allow for material regeneration without applying heat.31
![]() | ||
Fig. 2 (a) CO2 adsorption (closed)–desorption (open) isotherms at 273 K and (b) isosteric heat of adsorption of CO2 for the polymers. |
CH4 gas is being considered as an alternative fuel for automotive applications because of its abundance and cleaner burning nature. We also studied CH4 uptake in order to evaluate the impact of narrow pores on the binding affinity of CH4. Following a similar trend to that observed for CO2 adsorption, C-NP (1.10 wt%) and CL-SP (0.92 wt%) show higher CH4 uptake than CL-NP (0.66 wt%) and C-SP (0.84 wt%) networks at 273 K. The CH4 Qst values (Fig. S6†) calculated for these polymers remain as 19–21 kJ mol−1 which agree with the values obtained for most reported polymers.12,31 Additionally, the lower Qst values of CH4 for these polymers suggest that there are reduced interactions between CH4 and the pore walls due to its non-polar nature.
Effective CO2 adsorbents for gas separation applications are expected to have high selectivity for CO2 over other gases such as N2 (∼75% in flue gas) and CH4 (∼95% in natural gas) along with high CO2 capacity. Therefore, we carried out CO2/N2 and CO2/CH4 selectivity studies for these polymers to examine their gas separation abilities (Fig. 3 and Table 1). The initial slope ratios estimated from Henry's law constants for single-component adsorption isotherms were used to determine the adsorption selectivity. The maximum CO2/N2 and CO2/CH4 selectivities using the ratio of the initial slopes of adsorption isotherms for the polymers were calculated to be 36 and 10.6 at 273 K and 1 atm, respectively (see Fig. S7†). Importantly, enhanced CO2/N2 and CO2/CH4 selectivity values of 41 and 11.6 at 298 K and 1 atm were obtained for the polymers at 298 K and 1 atm. To predict CO2/CH4 and CO2/N2 binary mixture selectivity, an IAST calculation, coupled with a dual-site Langmuir simulation, was employed on the basis of single-component isotherms. Several studies show that the IAST model can accurately predict gas mixture adsorption of porous materials.16,31,32 Fig. 4 and S8† show the predicated selectivities for CO2/CH4 and CO2/N2 as a function of pressure when the gas phase mole fractions are 5/95 and 15/85, which are the typical feed compositions for natural gas and flue gas, respectively. Almost in all cases, selectivities gradually decrease with an increase in the pressure. It is worth noticing that high CO2/N2 selectivity values of 102–117 are obtained at 273 K and 1 atm (Table 1), which is similar to other reported materials.31 Surprisingly, at 298 K, which is closer to actual power plant conditions, CL-SP, C-SP and C-NP show enhanced CO2/N2 selectivity values of 155, 127 and 121, respectively. The initial trend in the selectivity is commonly observed for POFs. These values are sufficiently greater than any other carbazole porous material reported to date.17–20 These results suggest that CL-SP could be an excellent candidate for the selective sorption of CO2 from flue gas. The initial slope calculations for CO2/N2 selectivity resulted in lower values than those calculated by IAST method. However, the IAST selectivities were in reasonable agreement with the values obtained from Henry's constant ratios when the gas feed was set CO2/N2: 50:
50 showing the ideality of the results. Additionally, we showed that at 273 K, CL-SP (16.3) shows higher CO2/CH4 selectivity than C-SP (16), CL-NP (15.1) and C-NP (14.7). As for CO2/N2 selectivity, these materials also show enhanced CO2/CH4 selectivity values at 298 K, i.e., CL-SP (18.3), C-SP (16.3) and CL-NP (16.4). These values are larger or comparable to other reported materials.33,34 Importantly, it should be noted that the enhancement in the CO2/N2 and CO2/CH4 selectivity is due to the reduced uptake of CH4 and N2 at 298 K in these polymers. This increased selectivity can be ascribed to the enhanced dipole–quadrupole interaction between the large quadrupole moment of CO2 molecules (13.4 × 10−40 cm2) and the polar N-doped sites.10 Importantly, it should be noted that the enhancement in the CO2/N2 selectivity may be due to the integrated effects of aromatic carbazole system coupled with narrow pores in these polymers that attribute to the high interaction strength of CO2 with the frameworks.18
![]() | ||
Fig. 4 IAST selectivities of CO2/N2 (a) and CO2/CH4 (b) for binary gas mixtures of 15/85 and 5/95 molar compositions, respectively. |
The potential of these polymeric materials in CO2 capture from flue gas (CO2/N2: 10/90) was evaluated using vacuum swing adsorption at 298 K as suggested by Snurr.35 The working capacities (ΔN1) were determined by calculating the CO2 adsorption difference between 1.0 and 0.1 bar. CO2 uptake under adsorption conditions (Nads1) were found to be 0.32, 0.27, 0.32 and 0.28 mol kg−1 for C-NP, CL-NP, CL-SP and C-SP, respectively. They exhibit ΔN1 of 0.29, 0.23, 0.29 and 0.25 mol kg−1, respectively which is comparable to that of the Co-carborane MOF-4b36 and [Zn2(tcpb){p-(CF3)NC5H4}2]36 but lower than TBILP-1,37 BILP-13 (ref. 38) and PEI (40 wt%) ⊂ PAF-5 (ref. 39) materials. Regenerability (R) was determined by the percent ratio of ΔN1 to the Nads1. Both C-NP and CL-SP show highest regenerability (R) of 90 which is also comparable to zeolite-5A,36 TBILP-1,37 BILP-13 (ref. 38) and PEI (40 wt%) ⊂ PAF-5 materials.39 The higher regenerability levels are associated with the secondary amines which have less preferable binding N-sites for CO2, therefore their regeneration processes is rather smooth than metal organic frameworks. The appreciable Nads1, ΔN1 and R values made these materials a promising candidate for selective gas adsorption.
Footnote |
† Electronic supplementary information (ESI) available: This section contains eight figures and one table including SEM images, X-ray diffraction patterns, 13C CP-MAS NMR spectra, FTIR spectra, thermogravimetric data, isosteric heat of adsorption for CH4 and gas selectivity data as well. See DOI: 10.1039/c5ra06767g |
This journal is © The Royal Society of Chemistry 2015 |