Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Triptycene based 1,2,3-triazole linked network polymers (TNPs): small gas storage and selective CO2 capture

Snehasish Mondal and Neeladri Das *
Department of Chemistry, Indian Institute of Technology Patna, Patna 800 013, Bihar, India. E-mail: neeladri@iitp.ac.in; neeladri2002@yahoo.co.in; Tel: +91 9631624708

Received 1st September 2015 , Accepted 13th October 2015

First published on 13th October 2015


Abstract

Herein, facile synthesis and characterization of four triazole linked network polymers (TNPs) in high yields is described. These nitrogen rich polymers are derived using a “click” reaction between 2,6,14-triazido triptycene and various di- or triethynyl comonomers. The TNPs are microporous and exhibit high surface area (SABET up to 1348 m2 g−1). Due to the incorporation of the 1,2,3-triazole motif (a CO2-philic moiety), the TNPs record moderate to high CO2 uptake (up to 4.45 mmol g−1 at 273 K and 1 bar). The TNPs also show very good CO2/N2 (up to 48) and CO2/CH4 (8–9) selectivity. While the highest storage capacity has been registered by TNP4 (CO2 4.45 mmol g−1 at 273 K and 1 bar, CH4 23.8 mg g−1, H2 1.8 wt%), the highest CO2/N2 and CO2/CH4 selectivity is shown by TNP3 which contains additional nitrogen rich building blocks in the form of heteroaromatic pyrazine rings. These results suggest that TNPs are porous materials with potential practical application in gas storage and separation.


Introduction

Design and syntheses of porous materials have attracted considerable research attention over the past decade.1–4 In recent years, research output in this direction has witnessed considerable growth.5–13 Depending on the pore size, porous materials may be classified as ultra-microporous14–16 microporous,12,13 mesoporous17–19 or macroporous.20,21 Among these, microporous materials have emerged as potential candidates for applications that include, but are not limited to, gas storage and separation. Microporous materials, such as zeolites22,23 and metal–organic frameworks (MOFs),1–5 contain metal atoms. Alternatively, microporous materials can also be designed in the form of metal-free organic polymers that are highly crosslinked. Acronyms such as COFs, CMPs, CTFs, EOFs, HCPs, MOPs, OCFs, PAFs, PIMs, POFs, POPs, PPFs, PPNs, etc. have been coined to describe metal-free porous organic polymers.6–13 In general, MOFs have limited industrial applications as porous materials because of the presence of relatively labile coordination bonds that lower their thermal stability.24 On the other hand, in microporous organic polymers (MOPs), light weight elements (such as H, C, N, and O) are interlinked via relatively stronger covalent bonds. Therefore these metal-free materials have lower mass densities and higher stabilities (chemical and thermal). In addition, MOPs can be easily tailored with respect to properties such as guest selectivity and pore properties, due to the availability of a plethora of multifunctional organic molecules that can serve as monomers towards the syntheses of MOPs.

In contemporary research dealing with the development of microporous materials, the design of MOPs for selective CO2 capture and sequestration has received special attention.24–46 This has stemmed from environmental and energy concerns related to CO2, which is a major greenhouse gas. It is well known that thermal power plants emitting flue gas (containing approximately 15% CO2) are a major contributor to anthropogenic CO2. Therefore, in order to contain global climate warming arising out of anthropogenic CO2 emissions, there is a pressing demand to develop novel materials that can efficiently and selectively adsorb large quantities of CO2 emitted from fossil-fuel fired power plants (classified as fixed-point CO2 emission sources). Also it is desirable to separate CO2 (present as an impurity) from natural gas that contains 80–95% methane.47,48 This explains the growing research interest to design MOPs having practical applications such as selective capture and sequestration of CO2.

As far as the design of polymers for better CO2 adsorption and gas selectivity is concerned, it has been reported that incorporation of polar groups in the framework of porous organic polymers not only enhances carbon dioxide uptake but also improves its selectivity over methane and nitrogen. This enhancement in CO2 uptake is attributed to van der Waals interactions (quadrupole–dipole interactions and/or hydrogen bonding) between CO2 and the polar functional groups that are present in the backbone of MOPs. Additionally, polymeric networks synthesized from electron rich monomers may also exhibit higher CO2 uptake. This property is attributed to favorable Lewis acid–base interactions between the electron deficient carbon atom in CO2 and the relatively electron rich polymer.49,50

1,2,3-Triazole is a nitrogen rich motif and it can be easily incorporated in the polymer framework by reacting monomers with multiple azide and ethynyl functional groups. Nguyen and co-workers have described the synthesis of porous networks by covalently linking tetrahedral monomers using click chemistry.33 Triazole molecules (1,2,3- and 1,2,4-triazole) may be considered CO2-philic, considering the magnitude of their CO2 binding energies.51 It is therefore anticipated that polymers having these nitrogen rich moieties in their backbone may be potential candidates for efficient CO2 sorption and separation. Triptycene, on the other hand, is a robust and structurally rigid motif that has three dimensional orientation of three arene rings fused to a [2.2.2]bicyclooctatriene moiety. Triptycene based porous polymers have often exhibited good gas storage and separation properties.52–58 This is often attributed to the “internal molecular free volume (IFV)” that arises from the inefficient packing of rigid three dimensional units of triptycene present in the polymer backbone.52,55,57,58

With these premises in mind, we were interested in designing network polymers that would contain both the 1,2,3-triazole motif (as polar groups) and triptycene motif (as a bulky rigid group). It is anticipated that such polymeric networks would be porous and simultaneously exhibit reasonably high uptake of CO2, primarily because of the CO2-philic nature of the polar 1,2,3-triazole moiety and the porosity arising from the inefficient packing of the triptycene units in the polymers.

Herein, we report the synthesis of a series of new triptycene based and 1,2,3-triazole linked network polymers (TNPs) using 2,6,14-triazidotriptycene and various di- and triethynyl substituted comonomers via a Cu(I)OAc catalyzed “click” reaction. Complete structural characterization of these polymers and their performance as materials for gas storage as well as selective gas uptake has been described elaborately. TNPs constructed from monomers with three ethynyl functional groups have a relatively higher surface area than TNPs derived from diethynyl comonomers. However, all TNPs exhibit microporosity (pore diameter < 2 nm). Additionally, these microporous TNPs also demonstrate similar or even better carbon dioxide storage ability and selectivity than a wide range of previously reported MOPs.

Experimental section

Materials and methods

Triptycene, aromatic bromides and Cu(I)OAc were obtained from Sigma Aldrich and were utilized without further purification. 2,6,14-Triazidotriptycene was prepared from triptycene using a literature reported protocol.59 All the aromatic ethynyl compounds were synthesized from the corresponding aromatic bromides using the Sonogashira cross-coupling reaction.60 Chromatographic purifications were performed using silica gel (100–200 mesh). Solid-state 13C cross-polarization magic angle spinning (CP-MAS) NMR spectra were obtained using a BRUKER 300 MHz (H-1 frequency) NMR spectrometer at a mass rate of 8 kHz and CP contact time of 2 ms with a delay time of 3 sec. FTIR analyses were performed using a Shimadzu IR Affinity-1 spectrometer. Elemental analyses were performed using a vario MICRO cube analyser. P-XRD data were recorded using a Rigaku TTRAX III X-ray diffractometer. Porosity analyses were performed using a Quantachrome Autosorb iQ2 Analyzer using UHP grade adsorbates. In a typical experiment, TNP (60–90 mg) was taken in a 9 mm large cell and attached to the degasser unit and degassed at 120 °C for 12–24 h. The samples were refilled with helium and weighed carefully and then the cells were attached to the analyzer unit. The temperature was maintained using a KGW isotherm bath (provided by Quantachrome) filled with liquid N2 (77 K), or a temperature controlled bath (298 K and 273 K).

Synthesis of TNP1

A typical experiment for the syntheses of TNPs is described using TNP1 as a representative example. 2,6,14-Triazidotriptycene (189 mg, 0.5 mmol), 2,6,14-triethynyltriptycene (163 mg, 0.50 mmol) and Cu(I)OAc (4 mg, 0.03 mmol) were taken in a Schlenk flask under an inert atmosphere. Degassed DMF (30 mL) was added to the flask with continuous stirring to dissolve the reactants. The flask (containing the reactant) was subsequently heated at 120 °C with vigorous stirring. After a few minutes of stirring, the reaction mixture assumed the form of a light yellow gelatinous precipitate. The reaction mixture was allowed to stir for additional 24 hours with the resultant formation of a large amount of gelatinous precipitate. The reaction mixture was cooled to room temperature and filtered through a glass frit followed by washing with DMF, DMSO, acetone, THF and dichloromethane. The product thus obtained was dried under reduced pressure at 100 °C and crushed with a mortar and pestle to yield a brownish yellow powder.

Yield: 338 mg, 96%; FT-IR (KBr): 3404, 1660, 1464, 1221, 1035 cm−1. Elemental analysis: calculated for C46H25N9: C 78.51, H 3.58, N 17.91. Found: C 72.57, H 3.67, N 15.13.

Synthesis of TNP2

TNP2 has been prepared following a similar procedure to that described for TNP1 using 2,6,14-triazidotriptycene (189 mg, 0.5 mmol), 4,4′-diethynyl-biphenyl (152 mg, 0.75 mmol) and Cu(I)OAc (4 mg, 0.03 mmol). Unlike TNP1, a large amount of yellow precipitate was observed after 24 hours stirring at 120 °C. After drying at reduced pressure at 120 °C, the final product was obtained as a light yellow powder.

Yield: 300 mg, 88%; FT-IR (KBr): 3400, 1606, 1477, 1226, 1038 cm−1. Elemental analysis: calculated for C44H26N9: C 77.63, H 3.85, N 18.52. Found: C 71.97, H 4.36, N 15.93.

Synthesis of TNP3

This polymer was synthesized using 2,6,14-triazidotriptycene (189 mg, 0.5 mmol), 2,6-diethynylpyrazine (96 mg, 0.75 mmol) and Cu(I)OAc (4 mg, 0.03 mmol). Unlike other TNPs, formation of a light brown fluffy precipitate was observed within a few minutes of stirring at 120 °C. After 24 hours of stirring, formation of a considerable amount of fluffy precipitate was observed. After drying at reduced pressure at 120 °C, the final product was obtained as a light brown fluffy powder.

Yield: 242 mg, 85%; FT-IR (KBr): 1569, 1489, 1238, 1033 cm−1. Elemental analysis: calculated for C32H17N12: C 67.48, H 3.01, N 29.51. Found: C 62.36, H 3.61, N 25.93.

Synthesis of TNP4

This polymer was synthesized using 2,6,14-triazidotriptycene (189 mg, 0.5 mmol), 1,3,5-triethynylbenzene (75 mg, 0.50 mmol) and Cu(I)OAc (4 mg, 0.03 mmol). In this case, a large amount of dark yellow gelatinous precipitation was observed after stirring for 24 hours. After drying at reduced pressure at 120 °C, the final product was obtained as a dark brown solid. The solid product was then crushed in a mortar to yield the final product as a brownish yellow powder.

Yield: 259 mg, 98%; FT-IR (KBr): 3391, 1614, 1486, 1226, 1046 cm−1. Elemental analysis: calculated for C32H17N9: C 72.86, H 3.25, N 23.90. Found: C 65.74, H 4.09, N 20.82.

Results and discussion

Synthesis and characterization of TNPs

The TNPs were synthesized utilizing the well-known Cu(I) catalyzed click reaction that is known to regioselectively yield the 1,4-disubstituted 1,2,3-triazole derivative. It is well documented in the literature that “standard” azide-alkyne click reactions, carried out in water or water/DMF mixture, utilize Cu(I) species generated in situ via reduction of CuSO4 with sodium ascorbate. However, in recent years, copper(I) acetate is being used as the source of Cu(I) instead of the CuSO4/sodium ascorbate mixture, since the latter catalytic system either yields insoluble products or reactions are sluggish in nature.61 Chen and Guan reported click polymerization using CuOAc (source of CuI) in DMF (as a solvent).62 Based on this literature report, we have synthesized a series of 1,4-disubstituted 1,2,3-triazole bridged network polymers (TNPs) using 2,6,14-triazidotriptycene and various monomers containing multiple ethynyl functional groups as outlined in Scheme 1.
image file: c5ta06939d-s1.tif
Scheme 1 Synthesis of TNP1–TNP4.

In a typical reaction for the synthesis of TNPs, 2,6,14-triazidotriptycene,59 the corresponding ethynyl monomer (di- or triethynyl) and 6 mol% (with respect to 2,6,14-triazidotriptycene) Cu(I)OAc were dissolved in 30 mL of DMF and refluxed for 24 hours under a N2 atmosphere. In the case of TNP1 and TNP4 that utilized triethynyl monomers, gelatinous precipitation was observed after completion of the reaction. In this case, upon drying the precipitate under vacuum, the final product was obtained as a brownish yellow solid. FESEM images of the product obtained for TNP1 and TNP4 revealed an aggregated sheet-like structure. On the other hand, in the case of reactions leading to the formation of TNP2 and TNP3 (using diethynyl monomers in conjugation with 2,6,14-triazidotriptycene), the corresponding products precipitated as fluffy solids which were distinctly different from the gelatinous precipitates observed in the case of TNP1 and TNP4. As a consequence, for TNP2 and TNP3, the FESEM images revealed the presence of aggregated particles in their morphologies which was distinctly different from that observed in the case of TNP1 and TNP4 (Fig. 1). Wide-angle X-ray diffraction (WAXD) analyses of TNPs suggested that these polymeric materials were amorphous in nature due to the featureless broad nature of the diffraction pattern in each case (Fig. 1). This amorphous nature of the TNPs was due to the rapid and irreversible formation of the triazole linkage during polymerization.12b


image file: c5ta06939d-f1.tif
Fig. 1 FESEM micrograph of TNP1–TNP4 (left); WAXD pattern of TNP1–TNP4 recorded at ambient temperature at a scan rate of 3° min−1 (right).

The TNPs thus obtained were structurally characterized using solid state 13C (CP-MAS) NMR, and Fourier transform infrared (FTIR) spectroscopy. The 13C CP-MAS spectrum of TNP4 has been shown as a representative example (Fig. 2). The peak centered at 53 ppm corresponds to the bridgehead carbon of triptycene. The appearance of this peak is consistent with that of literature reported triptycene based monomer or polymers54–57,59 thereby confirming the successful incorporation of the triptycene motif in the polymer backbone. The peaks due to the phenyl ring carbons of the triptycene appear in the range of 120–150 ppm. The resonance signal centered at 168 ppm is assigned to a carbon present in the 1,2,3-triazole motif, thereby confirming the formation of 1,2,3-triazole linkages via the click reaction. The 13C CP-MAS spectra (Fig. S1) of the other TNPs match well with those of TNP4, thereby supporting the formation of the desired polymers as shown in Scheme 1.


image file: c5ta06939d-f2.tif
Fig. 2 13C CP-MAS NMR spectrum of TNP4, asterisks represent the spinning sideband.

Polymerization has also been confirmed by FTIR spectroscopic data analysis (Fig. S2). The FTIR spectrum of TNP2 is shown as a representative example in Fig. 3. The characteristic band of the azide functional group present in the triptycene monomer (2096 cm−1) is absent in TNP2. Simultaneously, a band due to the C–H stretching vibration of terminal alkynes (3280 cm−1), observed in the corresponding diethynyl comonomer, is missing in the FTIR spectrum of TNP2. This strongly suggests the click polymerization reaction between monomers containing terminal azide and terminal alkyne groups.


image file: c5ta06939d-f3.tif
Fig. 3 FT-IR spectrum of 2,6,14-triazidotriptycene (top), TNP2 (middle), and 4,4′-diethynylbiphenyl (bottom).

Porosity measurements and gas storage studies

Porous properties of the TNPs were investigated by subjecting these polymers to N2 adsorption–desorption measurements at 77 K. The isotherms are characterized by a steep increase in slope in the low relative pressure range (P/P0 = 0–0.1) and this implies uptake of large quantities of N2 gas. Additionally, the isotherms are reversible in nature. These characteristic features of the obtained adsorption–desorption isotherms (Fig. 4) suggest that these are type-I isotherms and the corresponding TNP polymers reported herein are microporous materials, as per IUPAC classifications.63 A gradual increase in N2 uptake in the relatively high pressure range (P/P0 = 0.1–0.9) was also observed in these isotherms. In the case of all TNPs, the adsorption and desorption branches of the hysteresis loop remain almost parallel over an appreciable range of P/P0, thereby suggesting that these materials (TNP1–4) show type H4 adsorption–desorption hysteresis. According to IUPAC, “the type H4 loop appears to be associated with narrow slit-like pores, but in this case the type I isotherm character is indicative of microporosity”.63
image file: c5ta06939d-f4.tif
Fig. 4 N2 adsorption isotherm of TNP1–TNP4 at 77 K (a); pore size distribution of TNP1–TNP4 (b).

An additional feature observed in the isotherms of TNP2 and TNP3 is their steep rise while the relative pressure (P/P0) is greater than 0.9, which is an indication of the presence of larger pores (macropores) and interparticle voids.24,64 Thus isotherms of TNP2 and TNP3 are distinctly different from those of TNP1 and TNP4. It may be recalled that in the FESEM micrographs, TNP2 and TNP3 exhibited particle-like nature having interparticle porosity, while TNP1 and TNP4 exhibit more aggregated plate-like morphologies suggesting that these have insignificant interparticle porosity. Thus conclusions drawn from the analysis of FESEM micrographs and adsorption–desorption isotherms substantiate each other.

DFT model based analysis results of pore size distributions (PSDs) of TNPs (calculated from N2 adsorption data at 77 K) are depicted in Fig. 4b. In the case of TNP1 and TNP4 wherein both comonomers are trifunctional, the pores are smaller than 2 nm. For these polymers, a narrow peak is observed in the ultramicroporous region (pore width < 0.7 nm) at 0.59 nm and 0.57 nm respectively, thereby suggesting the presence of ultramicropores in these polymers. The PSD profiles of TNP2 and TNP3 also confirm the presence of micropores. Here a major peak is centered at 1.7 nm and 1.6 nm respectively for TNP2 and TNP3. The total pore volume of TNPs (calculated at P/P0 = 0.99) is in the range of 0.53 to 0.66 cm3 g−1 (Table 1). In general, PSD is relatively narrow in the case of polymers (TNP1 and TNP4) that utilize both trifunctional monomers relative to polymers (TNP2 and TNP3) where one of the comonomers is bifunctional. These results clearly show that the pore size distribution of TNPs can be controlled by varying the number of reactive sites in the monomers. The microporous nature of TNPs is attributed to the presence of the rigid three dimensional triptycene motifs in the polymer backbone.

Table 1 Reaction yield and pore properties of TNP1–TNP4
Polymer SABETa (m2 g−1) SALangb (m2 g−1) V total (cm3 g−1) Yield (%)
a Surface area calculated based on the BET model from the nitrogen adsorption isotherm (P/P0 = 0.01–0.1). b Surface area calculated based on the Langmuir model from the nitrogen adsorption isotherms (P/P0 = 0.05–0.35). c The total pore volume calculated at P/P0 = 0.99.
TNP1 1090 1475 0.58 96
TNP2 460 756 0.53 88
TNP3 745 1006 0.62 85
TNP4 1348 1723 0.66 98


The surface areas (SAs) of TNPs have been obtained using the Brunauer–Emmett–Teller (BET) model within the pressure range P/P0 = 0.01 to 0.1 (Fig. S3). The SABET obtained for TNP1, TNP2, TNP3 and TNP4 are 1090 m2 g−1, 460 m2 g−1, 745 m2 g−1 and 1348 m2 g−1 respectively. The corresponding Langmuir surface areas are 1475 m2 g−1, 756 m2 g−1, 1006 m2 g−1 and 1723 m2 g−1 respectively (Table 1 and Fig. S4). Polymers TNP1 and TNP4 synthesized from triethynyl comonomers exhibit higher surface areas relative to the other two polymers (TNP2 and TNP3) where a diethynyl comonomer has been utilized for polymerization. This difference in the surface area of network polymers (TNP1 and TNP4) resulting from triethynyl monomers relative to polymers (TNP2 and TNP3) derived from disubstituted monomers is attributed to the higher extent of cross-linking as well as difference in the geometry of the comonomers. Irrespective of the nature of ethnyl monomer used in conjugation with 2,6,14-triazidotriptycene, the resulting TNPs are significantly porous and their corresponding surface areas (SABET) are comparable to or better than those of a wide variety of microporous polymers reported in the literature such as mPAF (948–1314 m2 g−1),27 benzimidazole-linked polymers (BILPs, 599–1306 m2 g−1),25,65,66 imine-linked PPFs (419–1740 m2 g−1),24 nitrogen-doped microporous carbons (263–744 m2 g−1),29 imine-linked POFs (466–1521 m2 g−1),31 nanoporous azo-linked polymers (ALPs, 862–1235 m2 g−1),32 “click-based” POPs (1090–1440 m2 g−1)33 and triptycene-based microporous polymers such as benzimidazole networks (600 m2 g−1),56 porous polymers derived from Tröger's base (TB-MOP, 694 m2 g−1),58 and microporous polyimides (STPIs, 4–541 m2 g−1).54 The surface areas and pore properties are summarized in Table 1.

Considering the microporous nature of TNPs and the presence of nitrogen-rich 1,2,3-triazole moieties in the polymer framework, we were interested to investigate the potential of TNPs as porous materials for gas storage applications. It has been reported in the literature that nitrogen enriched microporous organic polymers perform well as materials for storage of small gas molecules in general and particularly CO2. Therefore to assess the performance of TNPs in selective gas storage materials, adsorption isotherms for CH4, CO2, and H2 were collected under different conditions.

The CO2 sorption isotherms of TNPs were collected at 273 and 298 K and pressures up to 1 bar. The corresponding CO2 isotherms register a steep rise in the initial low pressure region and are completely reversible in the entire region due to the absence of any significant adsorption–desorption hysteresis [Fig. 5a (273 K) and Fig. S5 (298 K)]. The reversibility suggests that interactions between TNPs and CO2 are weak to the extent of regeneration of the polymers without application of heat.32 As observed in the case of N2 sorption analysis, TNPs (1 and 4) derived from triethynyl comonomers show higher uptake of CO2 relative to TNPs (2 and 3) derived from diethynyl comonomers. At 273 K and 1.0 bar, TNP4 demonstrates the highest CO2 uptake of 196 mg g−1 (4.45 mmol g−1), while TNP2 shows the lowest value of CO2 uptake at 70 mg g−1 (1.59 mmol g−1). Importantly, the magnitude of CO2 uptake by these TNPs is comparable to or better than those of various literature reported COFs. For example, the uptake of TNP4 (4.45 mmol g−1 at 273 K/1 bar) exceeds this value for –OH functionalized POFs (4.2 mmol g−1),67 imine linked ILP (1.97 mmol g−1),29 BLP1 (4.27 mmol g−1),65 and tetraphenyladamantane based PAN (3.36–4.0 mmol g−1).37 In general, the CO2 uptake values (1.59–4.45 mmol g−1 at 273 K/1 bar) of TNPs are comparable with those of various microporous organic polymers such as triazine based TBILP (2.66–5.18 mmol g−1),42 microporous covalent triazine polymers (MCTP, 3.65–4.64 mmol g−1),40 and carbazolic porous organic frameworks (Cz-POFs, 1.75–4.77 mmol g−1).38 Furthermore, the CO2 uptake capacities of these TNPs are comparable to those of MOFs such as Zn2(C2O4) (C2N4H3)2·(H2O)0.5 (4.35 mmol g−1)68 and amine-functionalized MOFs such as bio-MOF-11 (6.0 mmol g−1).69 The observed substantial uptake of CO2 by these triptycene-based TNPs can be considered as a consequence of high internal molecular free volume (IMFV) due to the triptycene units and high nitrogen content due to the triazole motifs. Recently, theoretical studies have shown that 1,2,3-triazole molecules have a large CO2-binding energy (BE) (20.0 kJ mol−1) possibly due to the high affinity of triazole rings for CO2via strong electrostatic interactions.51


image file: c5ta06939d-f5.tif
Fig. 5 CO2 uptake isotherm of TNP1–TNP4 at 273 K (a), isosteric heat of adsorption {Qst} of TNPs for CO2 (b), methane uptake isotherm of TNP1–TNP4 at 273 K (c) and Qst of TNPs for CH4 (d). Adsorption (filled) and desorption (empty).

To further understand the interaction of CO2 with TNPs, the isosteric heats of adsorption (Qst) for CO2 in TNPs were estimated from the CO2 adsorption isotherms at 273 and 298 K. The Qst values for CO2 at zero coverage (at the onset of adsorption) are in the range of 34.8–38.5 kJ mol−1. Strong interactions between CO2 and polar 1,2,3-triazole motifs as well as the narrow distribution in the microporous region may be the reason for the high Qst value observed in the low pressure region. The Qst values (34.8–38.5 kJ mol−1) of TNPs are higher than those of a variety of previously reported porous organic polymers such as nanoporous azo-linked polymers (ALPs: 27.9–29.6 kJ mol−1),32 imine-linked porous polymer frameworks (PPFs: 21.8–29.2 kJ mol−1),24 benzimidazole-linked polymers (BLPs: 26.7–28.0. kJ mol−1),25,65,66 nitrogen-rich networks (PECONFs: 26–34 kJ mol−1)64 nitrogen-rich diaminotriazine-based polymers (APOPs, 26.6–33.3 kJ mol−1),70 and carbazolic porous organic frameworks (Cz-POFs, 24.8–27.8 kJ mol−1).38 However these values are slightly less than those of microporous covalent triazine polymers (MCTPs, 30.6–40 KJ mol−1).40

In addition to CO2, TNPs were also investigated as materials for the sorption of other gases such as H2 and CH4 (Table 2). CH4 gas isotherms of TNPs were collected at 273 K (Fig. 5c) and 298 K (Fig. S5) up to 1 bar pressure. At 273 K, TNP4 registers the highest methane uptake (23.8 mg g−1) while TNP2 records the lowest (10.0 mg g−1). Irrespective of the extent of methane sorption, all the isotherms are completely reversible. We have also calculated the isosteric heat of adsorption (Qst) for methane in TNPs and values at zero coverage are found to be in the range of 13.2–20.0 kJ mol−1. These magnitudes of Qst are comparable with those of various hetero-functionalized porous organic polymers.24,25,27,32

Table 2 H2, CO2, CH4, and N2 uptakes, isosteric heats of adsorption (Qst) for TNPs, and selectivity for CO2/N2 and CO2/CH4
Polymer H2 at 1 bar (77 K) (mg g−1) CO2 at 1 bar (mg g−1) CH4 at 1 bar (mg g−1) N2 at 1 bar (mg g−1) Selectivity
273 K 298 K Q st 273 K 298 K Q st 273 K 298 K CO2/N2 273 K(298 K) CO2/CH4 273 K(298 K)
TNP1 14 178 99 37.0 17.3 11.5 13.2 3.4 2.1 36(23) 8(5)
TNP2 8 70 43 38.5 10.0 5.2 20 2.3 1.1 40(23) 8(5)
TNP3 11 116 81 34.8 19.8 10.9 16.2 3.8 2.1 48(31) 9(6)
TNP4 18 196 127 36.5 23.8 14.2 20 7.3 3.2 31(27) 8(6)


Similarly, H2 adsorption isotherms were collected at 77 K for TNPs (Fig. 6). The uptake capacity of H2 increases in the order of TNP2 < TNP3 < TNP1 < TNP4 which is consistent with the increasing order of surface area (SABET). The H2 uptakes of TNPs are in the range of 0.8–1.8 wt% at 77 K and 1 bar, which is comparable with those of various organic porous polymers, such as nanoporous azo-linked polymers (ALPs, 1.39–2.19 wt%)32 and benzimidazole-linked polymers (BILP1, 1.9 wt%),65 under identical experimental conditions.


image file: c5ta06939d-f6.tif
Fig. 6 H2 uptake isotherms of TNP1–TNP4 at 77 K. Adsorption (filled) and desorption (empty).

Gas selectivity

After evaluation of the porosity and gas-uptake capacities of these TNPs, we were interested to study their ability to selectively capture CO2 over CH4 and N2. It is well known that in order to reduce environmental pollution due to CO2 (a greenhouse gas) present in flue gas (N2/CO2: 85[thin space (1/6-em)]:[thin space (1/6-em)]15), it is necessary to selectively capture CO2 over N2. Also it is well accepted that CO2 is a contaminant in natural gas (CH4/CO2: 95[thin space (1/6-em)]:[thin space (1/6-em)]5) and it is necessary to selectively capture CO2 over methane in order to improve the quality of natural gas as a fuel and also control corrosion in pipelines due to CO2.

In order to access the performance of TNPs in CO2/CH4 and CO2/N2 gas selectivities, it is required to record the single component adsorption isotherms of these gases at 273 K (Fig. 7) and 298 K (Fig. S6) up to 1 bar pressure. For a given TNP, an initial steeper rise in the CO2 adsorption isotherm in comparison to that in the case of the corresponding N2 or CH4 isotherm was anticipated, considering the higher Qst values of TNPs for CO2 in zero surface coverage (Table 2) due to the favorable interactions between CO2 and N2 centers present in the 1,2,3-triazole moieties. It must be mentioned here that in flue gas, CO2 partial pressure is typically 0.15 bar at 273 K.25 Thus for a given porous material, it is worth considering its relative uptake of gases (CO2vs. N2 or CH4) at 0.15 bar while evaluating its performance in selective gas adsorption. In the case of TNPs, at 0.15 bar the adsorption isotherms indicate significantly higher CO2 uptake relative to N2 or CH4. Considering this fact, these TNPs can be considered as useful materials for selective CO2 gas adsorption applications under these conditions.


image file: c5ta06939d-f7.tif
Fig. 7 Gas uptake capacities for TNP1–TNP4 at 273 K. CO2 (green squares), CH4 (blue stars), and N2 (red circles).

Subsequently, CO2/N2 and CO2/CH4 gas selectivity of these TNPs was estimated from the initial slope ratios using Henry's law constants for single-component adsorption isotherms (Fig. S7 and S8). This method has been routinely used to study gas selectivity properties of a wide variety of organic and inorganic hybrid materials.25,32

Therefore, the N2/CO2 selectivity of TNPs has been calculated from the isotherms recorded at 273 K and 298 K in the pressure range below 0.15 bar, and the results are summarized in Table 2. Overall, the selectivity at 273 K (31–48) is higher than the selectivity at 298 K (23–31) for all TNPs. Amongst these four network polymers, TNP3 shows the highest selectivity for CO2 over N2 (48 at 273 K). This may be attributed to the presence of additional pyrazine motifs in TNP3 that are absent in the other three TNPs. This polymer (TNP3) offers more nitrogen sites for interaction with CO2 in comparison to the other TNPs. The data also suggest a considerable trade-off between gas storage capacity and gas selectivity amongst the four TNPs reported herein. For example, TNP4 having the highest surface area (SABET = 1348 m2 g−1) exhibits less selectivity for CO2 over N2 (31 at 273 K) in comparison to TNP1 (SABET = 1090 m2 g−1, CO2/N2 = 36 at 273 K) and TNP2 (SABET = 460 m2 g−1, CO2/N2 = 40 at 273 K). In general, the TNP with a relatively higher surface area exhibits relatively lower CO2/N2 selectivity. TNP3 is an exception since it has higher nitrogen content than other TNPs. Thus our results corroborate with the general perception that a higher porosity level is accompanied by a compromise in CO2/N2 selectivity.25,32 Nevertheless, the CO2/N2 selectivities of these 1,2,3-triazole linked network polymers (TNPs) are better than those of a wide range of microporous materials reported in the literature.32,35–38,41,70

Additionally, the CO2/CH4 selectivity of TNPs was also investigated to ascertain their applicability in the purification of natural gas (CH4/CO2: 95[thin space (1/6-em)]:[thin space (1/6-em)]5) by capture of CO2. Methane adsorption isotherms were collected at 273 K and 298 K and the selectivity was determined using the initial slope calculation, and the results obtained are depicted in Table 2. CO2/CH4 selectivities at 273 K and 1 bar are in the range of 8–9. At a higher temperature (298 K), the CO2/CH4 selectivity was found to decrease to 5–6. In general, the TNPs exhibit significantly lower CO2/CH4 selectivity relative to CO2/N2 selectivity. It is reported that due to the higher polarizability of methane relative to N2, the adsorption potential of methane is much higher than that of N2 and this may be the explanation for the observed lower CO2/CH4 selectivities in comparison to CO2/N2 selectivities.32

Conclusions

In conclusion, facile synthesis and characterization of a series of triptycene based and 1,2,3-triazole laced alternating copolymers has been described. The resulting 1,2,3-triazole linked network polymers (TNPs) are microporous materials exhibiting BET surface area up to 1348 m2 g−1. We have also shown that incorporation of nitrogen-rich pyrazine moieties results in enhancement of CO2/N2 gas selectivity relative to TNPs that do not possess this hetero-aromatic structural motif in the polymer backbone. In general, we have shown that integration of triptycene and 1,2,3-triazole motifs in the backbone of organic network polymers results in the yield of materials that demonstrate excellent performance in the storage of small gases at low pressure. Considering the microporosity, large gas uptake and high selectivities of CO2 over N2 and CH4, TNPs are potential candidates for practical applications such as CO2 capture as well as purification of flue and natural gas. Currently, we are extending our research in the direction of a facile synthesis of novel microporous organic polymers with applications in the clean energy sector by utilizing new multifunctional monomers for tailoring the gas sorption properties.

Acknowledgements

N. D. thanks the CSIR, Govt of India, New Delhi [CSIR No. 02(0126)/13/EMR-II] for financial support. S. M. thanks UGC, New Delhi for a Research Fellowship. The authors acknowledge the Indian Institute of Technology Patna for providing the infrastructure required for this research. The authors also acknowledge Central NMR Facility, CSIR-NCL for acquiring 13C CP-MAS data.

References

  1. K. Sumida, D. L. Rogow, J. A. Mason, T. M. McDonald, E. D. Bloch, Z. R. Herm, T. Bae and J. R. Long, Chem. Rev., 2012, 112, 724–781 CrossRef CAS PubMed.
  2. B. Chen, S. Xiang and G. Qian, Acc. Chem. Res., 2010, 43, 1115–1124 CrossRef CAS PubMed.
  3. J.-R. Li, Y.-G. Ma, M. C. McCarthy, J. Sculley, J.-M. Yu, H.-K. Jeong, P. B. Balbuena and H.-C. Zhou, Coord. Chem. Rev., 2011, 255, 1791–1823 CrossRef CAS.
  4. W. Zhou, Chem. Rec., 2010, 10, 200–204 CrossRef CAS PubMed.
  5. Z. Zhang, Z.-Z. Yao, S. Xiang and B. Chen, Energy Environ. Sci., 2014, 7, 2868–2899 CAS.
  6. W. Shen and W. Fan, J. Mater. Chem. A, 2013, 1, 999–1013 CAS.
  7. T. A. Makal, J.-R. Li, W. Lu and H.-C. Zhou, Chem. Soc. Rev., 2012, 41, 7761–7779 RSC.
  8. Z. Xiang and D. Cao, J. Mater. Chem. A, 2013, 1, 2691–2718 CAS.
  9. S.-Y. Ding and W. Wang, Chem. Soc. Rev., 2013, 42, 548–568 RSC.
  10. K. Sakaushi and M. Antonietti, Acc. Chem. Res., 2015, 48, 1591–1600 CrossRef CAS PubMed.
  11. M. L. Foo, R. Matsuda and S. Kitagawa, Chem. Mater., 2014, 26, 310–322 CrossRef CAS.
  12. (a) Y. Xu, S. Jin, H. Xu, A. Nagai and D. Jiang, Chem. Soc. Rev., 2013, 42, 8012–8031 RSC; (b) X. Feng, X. Dinga and D. Jiang, Chem. Soc. Rev., 2012, 41, 6010–6022 RSC.
  13. Z. Chang, D.-S. Zhang, Q. Chen and X.-H. Bu, Phys. Chem. Chem. Phys., 2013, 15, 5430–5442 RSC.
  14. S.-Y. Lee and S.-J. Park, J. Anal. Appl. Pyrolysis, 2014, 106, 147–151 CrossRef CAS.
  15. R. Swaidan, M. Al-Saeedi, B. Ghanem, E. Litwiller and I. Pinnau, Macromolecules, 2014, 47, 5104–5114 CrossRef CAS.
  16. B. S. Ghanem, R. Swaidan, E. Litwiller and I. Pinnau, Adv. Mater., 2014, 26, 3688–3692 CrossRef CAS PubMed.
  17. A. Corma, Chem. Rev., 1997, 97, 2373–2419 CrossRef CAS PubMed.
  18. R. Ryoo, S. H. Joo, M. Kruk and M. Jaroniec, Adv. Mater., 2001, 13, 677–681 CrossRef CAS.
  19. F. Hoffmann, M. Cornelius, J. Morell and M. Froeba, Angew. Chem., Int. Ed., 2006, 45, 3216–3251 CrossRef CAS PubMed.
  20. A. R. Studart, U. T. Gonzenbach, E. Tervoort and L. J. Gauckler, J. Am. Ceram. Soc., 2006, 89, 1771–1789 CrossRef CAS.
  21. O. Okay, Prog. Polym. Sci., 2000, 25, 711–779 CrossRef CAS.
  22. M. E. Davis and R. F. Lobo, Chem. Mater., 1992, 4, 756–768 CrossRef CAS.
  23. J. Caro, M. Noack, P. Kolsch and R. Schafer, Microporous Mesoporous Mater., 2000, 38, 3–24 CrossRef CAS.
  24. Y. Zhu, H. Long and W. Zhang, Chem. Mater., 2013, 25, 1630–1635 CrossRef CAS.
  25. M. G. Rabbani and H. M. El-Kaderi, Chem. Mater., 2012, 24, 1511–1517 CrossRef CAS.
  26. W. Lu, D. Yuan, J. Sculley, D. Zhao, R. Krishna and H.-C. Zhou, J. Am. Chem. Soc., 2011, 133, 18126–18129 CrossRef CAS PubMed.
  27. M. Errahali, G. Gatti, L. Tei, G. Paul, G. A. Rolla, L. Canti, A. Fraccarollo, M. Cossi, A. Comotti, P. Sozzani and L. Marchese, J. Phys. Chem. C, 2014, 118, 28699–28710 CAS.
  28. M. Saleh, S. B. Baek, H. M. Lee and K. S. Kim, J. Phys. Chem. C, 2015, 119, 5395–5402 CAS.
  29. J. Wang, I. Senkovska, M. Oschatz, M. R. Lohe, L. Borchardt, A. Heerwig, Q. Liu and S. Kaskel, ACS Appl. Mater. Interfaces, 2013, 5, 3160–3167 CAS.
  30. V. M. Suresh, S. Bonakala, H. S. Atreya, S. Balasubramanian and T. K. Maji, ACS Appl. Mater. Interfaces, 2014, 6, 4630–4637 CAS.
  31. P. Pandey, A. P. Katsoulidis, I. Eryazici, Y. Wu, M. G. Kanatzidis and S. T. Nguyen, Chem. Mater., 2010, 22, 4974–4979 CrossRef CAS.
  32. P. Arab, M. G. Rabbani, A. K. Sekizkardes, T. İslamoğlu and H. M. El-Kaderi, Chem. Mater., 2014, 26, 1385–1392 CrossRef CAS.
  33. P. Pandey, O. K. Farha, A. M. Spokoyny, C. A. Mirkin, M. G. Kanatzidis, J. T. Hupp and S. T. J. Nguyen, Mater. Chem., 2011, 21, 1700–1703 RSC.
  34. Y.-S. Bae and R. Q. Snurr, Angew. Chem., Int. Ed., 2011, 50, 11586–11596 CrossRef CAS PubMed.
  35. Z. Wang, D. Wang, F. Zhang and J. Jin, ACS Macro Lett., 2014, 3, 597–601 CrossRef CAS.
  36. H. J. Jo, C. Y. Soo, G. Dong, Y. S. Do, H. H. Wang, M. J. Lee, J. R. Quay, M. K. Murphy and Y. M. Lee, Macromolecules, 2015, 48, 2194–2202 CrossRef CAS.
  37. G. Li, B. Zhang, J. Yan and Z. Wang, Macromolecules, 2014, 47, 6664–6670 CrossRef CAS.
  38. X. Zhang, J. Lu and J. Zhang, Chem. Mater., 2014, 26, 4023–4029 CrossRef CAS.
  39. Y.-Q. Shi, J. Zhu, X.-Q. Liu, J.-C. Geng and L.-B. Sun, ACS Appl. Mater. Interfaces, 2014, 6, 20340–20349 CAS.
  40. P. Puthiaraj, S.-M. Cho, Y.-R. Lee and W.-S. Ahn, J. Mater. Chem. A, 2015, 3, 6792–6797 CAS.
  41. Y. Zhuang, J. G. Seong, Y. S. Do, H. J. Jo, Z. Cui, J. Lee, Y. M. Lee and M. D. Guiver, Macromolecules, 2014, 47, 3254–3262 CrossRef CAS.
  42. A. K. Sekizkardes, S. Altarawneh, Z. Kahveci, T. İslamoğlu and H. M. El-Kaderi, Macromolecules, 2014, 47, 8328–8334 CrossRef CAS.
  43. M. Salehab and K. S. Kim, RSC Adv., 2015, 5, 41745–41750 RSC.
  44. B. Ashourirad, A. K. Sekizkardes, S. Altarawneh and H. M. El-Kaderi, Chem. Mater., 2015, 27, 1349–1358 CrossRef CAS.
  45. M. Yu, X. Wang, X. Yang, Y. Zhao and J.-X. Jiang, Polym. Chem., 2015, 6, 3217–3223 RSC.
  46. R. Dawson, E. Stöckel, J. R. Holst, D. J. Adams and A. I. Cooper, Energy Environ. Sci., 2011, 4, 4239–4245 CAS.
  47. S. Chu, Science, 2009, 325, 1599 CrossRef CAS PubMed.
  48. R. S. Haszeldine, Science, 2009, 325, 1647–1652 CrossRef CAS PubMed.
  49. X. Lu, D. Jin, S. Wei, Z. Wang, C. An and W. Guo, J. Mater. Chem. A, 2015, 3, 12118–12132 CAS.
  50. Y.-S. Bae and R. Q. Snurr, Angew. Chem., Int. Ed., 2011, 50, 11586–11596 CrossRef CAS PubMed.
  51. H. M. Lee, I. S. Youn, M. Saleh, J. W. Lee and K. S. Kim, Phys. Chem. Chem. Phys., 2015, 17, 10925–10933 RSC.
  52. C. Zhang, Y. Liu, B. Li, B. Tan, C.-F. Chen, H.-B. Xu and X.-L. Yang, ACS Macro Lett., 2012, 1, 190–193 CrossRef CAS.
  53. (a) M. Carta, M. Croad, R. Malpass-Evans, J. C. Jansen, P. Bernardo, G. Clarizia, K. Friess, M. Lanč and N. B. McKeown, Adv. Mater., 2014, 26, 3526–3531 CrossRef CAS PubMed; (b) B. S. Ghanem, M. Hashem, K. D. M. Harris, K. J. Msayib, M. Xu, P. M. Budd, N. Chaukura, D. Book, S. Tedds, A. Walton and N. B. McKeown, Macromolecules, 2010, 43, 5287–5294 CrossRef CAS.
  54. C. Zhang, T.-L. Zhai, J.-J. Wang, Z. Wang, J.-M. Liu, B. Tan, X.-L. Yang and H.-B. Xu, Polymer, 2014, 55, 3642–3647 CrossRef CAS.
  55. T. M. Swager, Acc. Chem. Res., 2008, 41, 1181–1189 CrossRef CAS PubMed.
  56. Y. He, X. Zhu, Y. Li, C. Peng, J. Hu and H. Liu, Microporous Mesoporous Mater., 2015, 214, 181–187 CrossRef CAS.
  57. Q. Wen, T.-Y. Zhou, Q.-L. Zhao, J. Fu, Z. Ma and X. Zhao, Macromol. Rapid Commun., 2015, 36, 413–418 CrossRef CAS PubMed.
  58. 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.
  59. S. Chakraborty, S. Mondal, R. Kumari, S. Bhowmick, P. Das and N. Das, Beilstein J. Org. Chem., 2014, 10, 1290–1298 CrossRef PubMed.
  60. (a) M. I. Mangione, R. A. Spanevello, A. Rumbero, D. Heredia, G. Marzari, L. Fernandez, L. Otero and F. Fungo, Macromolecules, 2013, 46, 4754–4763 CrossRef CAS; (b) S. Chakraborty, S. Mondal and N. Das, Inorg. Chim. Acta, 2014, 413, 214–220 CrossRef CAS; (c) S. Bhowmick, S. Chakraborty, A. Das, P. R. Rajamohanan and N. Das, Inorg. Chem., 2015, 54, 2543–2550 CrossRef CAS PubMed; (d) S. Yuan, B. Dorney, D. White, S. Kirklin, P. Zapol, L. Yu and D.-J. Liu, Chem. Commun., 2010, 46, 4547–4549 RSC.
  61. A. Qin, J. W. Y. Lam and B. Z. Tang, Macromolecules, 2010, 43, 8693–8702 CrossRef CAS.
  62. Y. Chen and Z. Guan, J. Am. Chem. Soc., 2010, 132, 4577–4579 CrossRef CAS PubMed.
  63. K. S. W. Sing, Pure Appl. Chem., 1982, 54, 2201–2218 CrossRef.
  64. P. Mohanty, L. D. Kull and K. Landskron, Nat. Commun., 2011, 2, 1–6 Search PubMed.
  65. M. G. Rabbani and H. M. El-Kaderi, Chem. Mater., 2011, 23, 1650–1653 CrossRef CAS.
  66. M. G. Rabbani, T. E. Reich, R. M. Kassab, K. T. Jackson and H. M. El-Kaderi, Chem. Commun., 2012, 48, 1141–1143 RSC.
  67. A. P. Katsoulidis and M. G. Kanatzidis, Chem. Mater., 2011, 23, 1818–1824 CrossRef CAS.
  68. R. Vaidhyanathan, S. S. Iremonger, K. W. Dawson and G. K. H. Shimizu, Chem. Commun., 2009, 45, 5230–5232 RSC.
  69. J. An, S. J. Geib and N. L. Rosi, J. Am. Chem. Soc., 2010, 132, 38–39 CrossRef CAS PubMed.
  70. W.-C. Song, X.-K. Xu, Q. Chen, Z.-Z. Zhuang and X.-H. Bu, Polym. Chem., 2013, 4, 4690–4696 RSC.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ta06939d

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.