DOI:
10.1039/C6RA00735J
(Paper)
RSC Adv., 2016,
6, 31083-31091
PEG–imidazolium-functionalized 6FDA–durene polyimide as a novel polymeric membrane for enhanced CO2 separation†
Received
9th January 2016
, Accepted 17th March 2016
First published on 21st March 2016
Abstract
A series of poly(ethylene glycol)–imidazolium-functionalized 6FDA–durene polyimides (PEG–Im-PIs) with a range of PEG chain lengths was developed. The structures and properties of this series, as well as the gas separation properties of the corresponding polymer membranes, were studied. A PEG group and an imidazolium-based ionic group, both of which acted as CO2-solubilizing groups, were incorporated onto a rigid polyimide backbone to yield a very high CO2 solubility and, hence, excellent CO2/CH4 (49.2) and CO2/N2 (40.1) permselectivities while maintaining a good permeability of polyimide.
Introduction
The development of CO2 capture processes is a key technical, economical, and environmental challenge because a range of applications rely on the separation of CO2 from other light gases, such as N2 and CH4. CO2 and N2 are flue gases produced after the combustion of fossil fuels and air (post-combustion process). Natural gas production also involves a purification step because mined natural gases (primarily CH4) contain around 10% CO2. This quantity of CO2 considerably reduces the value of the mined natural gas because CO2 is not combustible. Furthermore, CO2 is highly corrosive, and its removal is necessary for natural gas pipeline transport.
A variety of techniques have been developed to perform these CO2/light gas separation steps, including pressure swing adsorption, cryogenic distillation, gas–liquid absorption, and membrane-based processes. Among these techniques, polymer membrane-based gas separation approaches have gained a great deal of attention because they are flexible, easy to scale up, and compete successfully with conventional separation processes, such as absorption (amine scrubbing).1–4 In general, polymer materials are mechanically stable, offer good separation capabilities, and can be fabricated in a range of configurations, such as hollow fibers or flat sheets.5 The commercially available, including polyimide (PI), poly(dimethylsiloxane) (PDMS), or cellulose acetate (CA), unfortunately do not meet several industrial requirements in that they do not afford a high CO2 flux (i.e., their CO2 permeability is low) or a high purity (i.e., the CO2/light gas selectivity is poor). In fact, as dense separation materials, polymers typically exhibit a trade-off between permeability and selectivity, represented by the so-called “upper bound” in the Robeson plot.6,7
Polymer membranes that can provide selective CO2 separation processes may be developed through the design of polymers with a high diffusivity–selectivity and/or a high solubility–selectivity. The diffusivity–selectivity reflects the ability of a polymer matrix to select the shape and size of the penetrant molecules.8 Diffusivity–selectivity is mainly governed by the relative mobilities of penetrants and by structural factors, such as the polymer chain stiffness and inter-segmental polymer packing structure. In general, gas permeability can be enhanced by increasing the diffusivity–selectivity for small gas molecule separation applications involving rigid glassy polymer membranes, such as polyimides (PIs), with narrow free volume distributions.
The solubility–selectivity properties, on the other hand, are largely governed by the relative strengths of the polymer–penetrant interactions and the relative condensabilities of the penetrants. The CO2/gas selectivity of a polymer membrane may often be improved by introducing CO2-philic species, such as ionic liquid (ILs),9 amines,10 or poly(ethylene oxide)s (PEO),11,12 all of which increase the solubility–selectivity of the material.
A handful of studies have examined CO2/N2 separation approaches using supported ionic liquid membranes (SILMs),13–15 in which microporous polymers are impregnated with ILs to achieve a high permeability and selectivity. The use of SILMs in practical gas separation processes, however, is limited by stability issues (leaching of ILs from the membranes at pressure differences beyond 0.2 atm).13
Membranes prepared from polymeric ILs, or poly(IL)s, may offer another option for enhancing the separation of CO2 from other gases because poly(IL)s offer a high CO2 sorption capacity and high sorption and desorption rates compared to ILs.16–22 Poly(IL)s are not strictly ionic liquids, but rather are polymers containing various forms of certain ionic salts. Nevertheless, poly(IL)s share similar features with ILs, such as a high CO2 solubility. Although these poly(ionic salts) have proven to be selective in the separation of CO2 from other gases, their utility in practical gas separation applications has been limited by their physical and chemical properties as well as by their low permeabilities (typically below 100 Barrer for CO2). Previously examined poly(IL)s have been prepared from IL-containing reactive monomers, and the polymer structures used to prepare the poly(IL)s have largely been limited to flexible polymer backbones, such as polyacrylates or polystyrenes.
We recently developed a new type of poly(IL)s in which pendant imidazolium or piperazinium-based ionic salts were introduced onto the rigid polyimide (PI) backbone as highly CO2-selective polymer membranes.20,21 Membranes prepared from these newly developed PI-based poly(IL)s displayed high CO2 separation and permeation properties, together with excellent mechanical and thermal stabilities.
PEO (or PEG: poly(ethylene glycol)) membrane materials can provide excellent CO2 separation and permeability properties because PEO (or PEG, a low molecular weight version of PEO) chains bearing polar ether oxygen units display a high affinity toward CO2 molecules to produce a high solubility–selectivity.23 Pure PEO (or PEG), however, tends to form a partially crystalline structure with a low permeability for gas molecules. A variety of approaches have been developed for combining the diffusivity–selective properties of glassy polymers and the solubility–selective properties of PEO in an effort to improve the CO2-selective gas separation performance. Such polymers include PEO-containing polyimides (PEO-PI),24–26 PEO-containing polyamides (PEO-PA),27 PEO-containing polyurethane or polyurea (PEO-PU),28 and PEO-containing polysulfone (PEO-PS).29 Most of these PEO-containing polymer membranes unfortunately have not provided permeabilities comparable to that of rigid polymers, such as PI (with a permeability exceeding 300 Barrer for CO2), except for a few examples.30,31
The present work combines the benefits of the above-mentioned PEG and poly(ionic salt)s, which offer a high CO2 solubility, with the highly permeable properties of rigid polymers by incorporating both PEG and IL onto PI, yielding a novel material for gas separation. The membranes obtained from the PEG–imidazolium-functionalized PIs (PEG–Im-PIs) displayed excellent thermal and mechanical stabilities and, most importantly, extremely high CO2/CH4 and CO2/N2 permselectivities, together with a very high CO2 permeability of 484.8 Barrer. We also investigated the effects of the PEG chains in the PEG–imidazolium groups on the structures and properties of the polymers, as well as the gas separation properties of the corresponding polymer membranes.
Experimental
Materials
4,4′-(Hexafluoroisopropylidene) diphthalic anhydride (6FDA), 2,3,5,6-tetramethyl benzene-1,4-diamine (durene) and N-bromosuccinimide (NBS) were purchased from Tokyo Chemical Industry (TCI) Co., Ltd. (Tokyo, Japan) and used as obtained. Triethyl amine, acetic anhydride, di(ethylene glycol) monomethyl ether, tetra(ethylene glycol) monomethyl ether and poly(ethylene glycol) monomethyl ether (Mn 550 g mol−1) were obtained from Sigma Aldrich. Poly(ethylene glycol) monomethyl ether (Mn 350 g mol−1) were obtained from Alfa-Aesar (a Johnson Matthey Co.). 6FDA and durene were dried under vacuum at 60 °C for 24 h prior to the polymerization. All other chemicals, unless otherwise noted, were obtained from commercial sources and used as received. Synthesis of the PEG–imidazoles having various PEG chains is described, together with their 1H NMR spectra (from Fig. S1 to S4), in the ESI.†
Characterization
1H NMR spectra were obtained on an Agilent 400-MR (400 MHz) instrument using d6-DMSO or CDCl3 as a reference or internal deuterium lock. FT-IR spectra of the materials were recorded as KBr pellets using Nicolet MAGNA 560-FTIR spectrometer in the range of 4000–400 cm−1. Molar masses were determined by Gel Permeation Chromatography (GPC) using two PL gel 30 cm × 5 μm mixed C columns at 30 °C running in THF and calibrated against polystyrene (Mn = 600 to 106 g mol−1) standards using a Knauer refractive index detector. The glass transition temperature (Tg) of each polymer was measured using a Perkin-Elmer Pyris-1 DSC from 20 °C to 300 °C with a scan rate of 10 °C min−1 under nitrogen.
Synthesis of PEG–imidazolium polyimides with various PEG chains (1)
Synthesis of 6FDA–durene polyimide (2). A general two-step synthesis such as polyamic acid formation followed by imidization was performed for the synthesis of 6FDA–durene polyimide. Into a 500 mL two-necked flask equipped with a magnetic stirrer, nitrogen inlet, and a condenser, 6FDA (13.52 g, 30.44 mmol), durene (5 g, 30.44 mmol), and DMAc (90 mL) were added. Then the reaction mixture was cooled to −5 °C (ice-bath), and allowed to stir for 12 h in order to form the corresponding polyamic acid. After this time, triethyl amine (8.91 mL, 63.9 mmol) and acetic anhydride (6.04 mL, 63.9 mmol) were added to the reaction mixture and then the temperature was raised to 110 °C (oil-bath) under vigorous stirring for 3 h to induce a complete imidization of polyamic acid to form polyimide. The viscous mixture was then cooled to r.t. and dissolved in DMAc (10 mL), followed by pouring into methanol (400 cm3). White polymer beads were collected by filtration, washed with deionized water several times and dried at 80 °C under vacuum for 48 h to give the 6FDA–durene polyimide 2 (16.3 g, 94%); δH (400 MHz, CDCl3) 8.1–8.0 (2H, br signal, ArH), 8.0–7.9 (4H, br signal, ArH), and 2.14 (12H, s, CH3); (ATR-FTIR)/cm−1 2925, 1786, 1712, 1370, 1250, 1187, 1112 and 980; GPC (DMF, RI)/Da Mn 123.82 kg mol−1, Mw 174.66 kg mol−1 and Mw/Mn 1.41.
Bromination of 6FDA–durene polyimide (2) to give the bromobenzylated PI (3)
The polyimide 2 (13 g, 22.7 mmol) and a catalytic amount of benzoyl peroxide (BPO) was dissolved with tetrachloroethane (60 mL) in a 250 mL two-necked flask equipped with a magnetic stirrer, nitrogen inlet, and a condenser. This was heated to 85 °C for a complete dissolution before adding N-bromosuccinimide (2.43 g, 13.6 mmol) and allowed for 12 h stirring at this temperature. The resultant red colored polymer solution was cooled to r.t. and precipitated into methanol (400 mL). The yellow-colored polymer beads were collected by filtration and washed with deionized water, and dried under vacuum at 80 °C for 48 h to give the brominated polyimide 3 (13.4 g, 90%); δH (400 MHz, CDCl3) 7.9–7.6 (10H, br signal, ArH), 7.4–7.3 (4H, br signal, ArH), 7.3–7.2 (2H, br signal, ArH), 7.2–7.1 (4H, br signal, ArH), 7.15 (2H, s, ArH), 7.0–6.7 (16H, br signal, ArH), 4.51 (2H, s, ArCH2), 2.17 (4H, s, ArH) and 1.70 (6H, s, CH3); (KBr)/cm−1 3043, 2967, 1747, 1641, 1496, 1236, 1145, 1024, 939, 841 and 620.
Incorporation of the PEG–imidazoles to give the PEG–Im-functionalized PIs (1)
To a solution of the brominated PI 3 (5 g, 7.67 mmol) in DMF (25 cm3), the corresponding PEG–imidazoles in DMF was added dropwise. The reaction mixture was heated to 90 °C for 24 h with vigorous stirring under nitrogen. After this time, the brown solution was cooled to r.t. and precipitated into ethyl acetate (200 cm3). The brown yellow-colored polymer powders were collected by filtration, washed 3 times with ether and dried at 80 °C under vacuum to give the desired PEG–imidazolium-functionalized PIs 1;
[C2PEG–ImPI][Br] (1a). (4.6 g, 83%); δH (400 MHz, d6-DMSO) 9.0–8.9 (1H, br signal, ArH), 8.3–7.8 (6H, br signal, ArH), 7.7–7.6 (1H, br signal, ArH), 7.3–7.2 (1H, br signal, ArH), 5.7–5.4 (2H, br signal, ArCH2), 4.4–4.2 (2H, br signal, NCH2), 3.7–3.6 (2H, br signal, NCH2CH2O), 3.4–2.9 (4H, br signal, NCH2CH2OCH2CH2O), 2.5–2.4 (3H, br signal, N(CH2CH2O)2CH3) and 2.1–1.9 (9H, br signal, CH3); (KBr)/cm−1 2925, 1795, 1710, 1560, 1485, 1450, 1360, 1270, 1190, 1100, 986 and 840.
[C4PEG–ImPI][Br] (1b). (4.1 g, 81%); δH (400 MHz, d6-DMSO) 9.0–8.9 (1H, br signal, ArH), 8.3–7.8 (6H, br signal, ArH), 7.8–7.7 (1H, br signal, ArH), 7.3–7.2 (1H, br signal, ArH), 5.7–5.5 (2H, br signal, ArCH2), 4.4–4.3 (2H, br signal, NCH2), 3.8–3.6 (2H, br signal, NCH2CH2O), 3.5–3.3 (10H, br signal, NCH2CH2(OCH2CH2O)2CH2), 3.2–3.1 (2H, br signal, N(CH2CH2O)3CH2CH2O), 2.5–2.4 (3H, br signal, N(CH2CH2O)4CH3) and 2.1–1.9 (9H, br signal, CH3); (KBr)/cm−1 2910, 1795, 1710, 1560, 1485, 1450, 1360, 1270, 1190, 1100, 986 and 841.
[C8PEG–ImPI][Br] (1c). (4.2 g, 86%); δH (400 MHz, d6-DMSO) 9.0–8.8 (1H, br signal, ArH), 8.3–7.8 (6H, br signal, ArH), 7.7–7.6 (1H, br signal, ArH), 7.3–7.2 (1H, br signal, ArH), 5.6–5.4 (2H, br signal, ArCH2), 4.3–4.2 (2H, br signal, NCH2), 3.7–3.6 (2H, br signal, NCH2CH2O), 3.5–3.3 (26H, br signal, NCH2CH2(OCH2CH2O)6CH2), 3.2–3.1 (2H, br signal, N(CH2CH2O)7CH2CH2O), 2.5–2.4 (3H, br signal, N(CH2CH2O)8CH3) and 2.1–1.9 (9H, br signal, CH3); (KBr)/cm−1 2920, 1795, 1710, 1560, 1480, 1440, 1360, 1270, 1190, 1100, 986 and 840.
[C12PEG–ImPI][Br] (1d). (4.1 g, 87%); δH (400 MHz, d6-DMSO) 9.0–8.8 (1H, br signal, ArH), 8.3–7.8 (6H, br signal, ArH), 7.7–7.6 (1H, br signal, ArH), 7.3–7.2 (1H, br signal, ArH), 5.7–5.5 (2H, br signal, ArCH2), 4.4–4.2 (2H, br signal, NCH2), 3.8–3.6 (2H, br signal, NCH2CH2O), 3.5–3.3 (42H, br signal, NCH2CH2(OCH2CH2O)10CH2), 3.2–3.1 (2H, br signal, N(CH2CH2O)11CH2CH2O), 2.5–2.4 (3H, br signal, N(CH2CH2O)12CH3) and 2.1–1.8 (9H, br signal, CH3); (KBr)/cm−1 2923, 1795, 1710, 1555, 1480, 1445, 1360, 1270, 1190, 1100, 986 and 850.
Membrane preparation
All the PEG–Im-functionalized PIs (1) and pristine PI (6FDA–durene, 2) membranes were prepared in a DMF solution of the corresponding polymers using the solution-casting method. The corresponding polymers 1 (or 3) (1.0 g) were dissolved in 5.0 cm3 of dry DMF and stirred at r.t. overnight. The solutions were poured onto glass plates after a thorough filtration through a plug of cotton. The plates were then placed in an oven, covered with aluminum foils having small holes and allowed to slow solvent evaporation at 70 °C for 48 h and further dried at 90 °C for 16 h in a vacuum oven. After becoming completely dried, the resulting membranes were cooled to r.t. and peeled off from the glass plate, and then being dried at the ambient temperature. The membrane thickness was controlled to be 70 to 90 μm.
Membrane characterization
The densities of the membranes (g cm−3) were determined experimentally using a top-loading electronic Mettler Toledo balance (XP205, Mettler-Toledo, Switzerland) coupled with a density kit based on Archimedes' principle. The samples were weighed in air and a known-density liquid, high purity water. The measurement was performed at room temperature by the buoyancy method and the density was calculated as follows
where, W0 and W1 are the membrane weights in air and water respectively. The water sorption of the PEG–ImPI membranes was not considered due to their extremely low water uptake property.
The X-ray diffraction patterns of the membranes were measured using a Rigaku DMAX-2200H diffractometer by employing a scanning rate of 4° min−1 in a 2θ range from 5° to 30° with a Cu Kα1 X-ray (λ = 0.1540598). The d-spacings were calculated using the Bragg's law (d = λ/2
sin
θ).
Tensile strength and elongation at break of the membranes were measured on a Shimadzu EZ-TEST E2-L instrument benchtop tensile tester using a crosshead speed of 1 mm min−1 at 25 °C under 50% relative humidity. Engineering stress was calculated from the initial cross sectional area of the sample and Young's modulus (E) was determined from the initial slope of the stress–strain curve. The membrane samples were cut into a rectangular shape with 80 mm × 8 mm (total) and 80 mm × 4 mm (test area), and five specimens were used for the measurements.
Gas permeation procedure
The pure gas permeation measurements were performed using a high-vacuum time lag measurement unit based on constant-volume/variable-pressure method. All the experiments were performed at a feed pressure of 2 atm and a feed temperature of 35 °C. Before the gas permeation measurements, both the feed and the permeate sides were thoroughly evacuated to below 10−5 Torr until the readout showed zero values to remove any residual gases. The downstream volume was calibrated using a Kapton membrane and was found to be 50 cm3. The upstream and downstream pressures were measured using a Baraton transducer (MKS; model no. 626B02TBE) with a full scale of 10
000 and 2 Torr, respectively. The pressure rise versus time transient of the permeate side, equipped with a pressure transducer, was recorded and passed to a desktop computer through a shield data cable. The permeability coefficient was determined from the linear slope of the downstream pressure rise versus time plot (dp/dt) according to the following equation: |
 | (1) |
where P is the permeability expressed in Barrer (1 Barrer = 10−10 cm3 (STP) cm cm−2 s−1 cm−1 Hg−1); V (cm3) is the downstream volume; l (cm) is the membrane thickness; A (cm2) is the effective area of the membrane; T (K) is the temperature of measurement; p0 (Torr) is the pressure of the feed gas in the upstream chamber and dp/dt is the rate of the pressure rise under the steady state. The gas permeation tests were repeated more than three times for all the gases, and the standard deviation from the mean values of permeabilities was within ca. ±3%. Sample to sample reproducibility was high and within ±3%. The effective membrane areas were 15.9 cm2. The ideal permselectivity, αA/B, of the membrane for a pair of gases (A and B) is defined as the ratio of the individual gas permeability coefficients as |
 | (2) |
The diffusivity and solubility were obtained from the time-lag (θ) value according to the following equations:
|
 | (3) |
|
 | (4) |
where
D (cm
2 S
−1) is the diffusivity coefficient,
l is the membrane thickness (cm) and
θ is the time lag (s), obtained from the intercept of the linear steady state part of downstream pressure rise
versus time plot. Solubility,
S, was calculated from
eqn (4) with permeability and diffusivity obtained from
eqn (1) and
(3).
Results and discussion
Synthesis of the PEG–imidazolium-functionalized polyimides having a variety of PEG chains
The synthesis of PEG–imidazolium-functionalized polyimides (PEG–Im-PI)s 1 is summarized in Scheme 1. The 6FDA–durene polyimide 2 was first synthesized by a polycondensation reaction between durene and 6FDA, following the literature procedures.32,33 This step was followed by a selective bromination at the benzylic position and the incorporation of a PEG–imidazolium group.
 |
| Scheme 1 Schematic representation of the preparation of the PEG–imidazolium-functionalized PI membranes with various PEG chain lengths. | |
The polyimide 2 was found to have a very high molecular weight (Mw = 174.66 kDa, as confirmed by GPC), and similarly high molecular weights were also reported.32,33 The selective bromination of the ArCH3 unit of PI 2 was conducted in tetrachloroethane solvent using 0.6 equiv. NBS to produce the bromobenzylated PI 3. Selective bromination was confirmed using a comparative 1H NMR spectroscopic analysis of 2 and 3 (Fig. S5 in the ESI†). Although the intensity of the benzylic proton (Ha) decreased and a new bromobenzyl proton peak (Hb) appeared at 4.5 ppm, no significant changes were observed among the other aromatic peaks, indicating the selective bromination of the benzyl group in 2. The degree of bromination in 3 was estimated to be 95%, based on the ratio of the integral of the bromobenzyl proton (Hb) in 3 relative to that of the benzylic proton (Ha) in 2.
Finally, PEG–imidazolium functionalization was carried out by treating a DMF solution of 3 with PEG–imidazole groups having a variety of ethylene glycol chain lengths, with C2, C4, C8, and C12, to give the corresponding PEG–Im-functionalized PIs (1) with bromide anions [C2PEG–Im-PI][Br], [C4PEG–Im-PI][Br], [C8PEG–Im-PI][Br], and [C12PEG–Im-PI][Br] (1a, 1b, 1c, and 1d), respectively.
As shown in Fig. 1, the 1H NMR spectra of the PEG–imidazolium-functionalized PIs display peaks characteristic of imidazolium protons (Hd) at 9.0 ppm and benzylic protons (Hc) of PEG–imidazoles at 5.6 ppm, indicating the successful incorporation of the PEG–imidazolium groups. The degree of functionalization was found to be almost 100% based on a calculation of the integral ratio of the Hb protons in 3 to the Hc protons in 1. FT-IR spectroscopy was performed and further confirmed the structures of the PEG–Im-functionalized PIs based on the peaks present at 1460 cm−1, 1480 cm−1, and 1560 cm−1, which corresponded to the vibrational modes of the benzyl imidazolium cations (Fig. S6 in ESI†).34
 |
| Fig. 1 1H NMR spectra of [C2PEG–Im-PI][Br] (a), [C4PEG–Im-PI][Br] (b), [C8PEG–Im-PI][Br] (c), and [C12PEG–Im-PI][Br] (d). | |
Preparation of the PEG–Im-functionalized PI membranes
The PEG–Im-functionalized PIs (1) displayed a high solubility in common organic solvents, such as CHCl3, DCM, DMF, DMSO, and DMAc. The corresponding membranes were prepared by casting a DMF solution of polymer 1, followed by vacuum drying to give dense, transparent, and flexible membranes (Fig. 2).
 |
| Fig. 2 Photographs of the PEG–Im-PI membranes. | |
Physical properties
DSC analysis was carried out to investigate the changes that occurred at the glass transition temperature (Tg) as a function of the PEG chain length in the PEG–imidazolium groups of the PEG–Im-PIs 1. All four PEG–Im-PIs ([C2PEG–Im-PI][Br], [C4PEG–Im-PI][Br], [C8PEG–Im-PI][Br], and [C12PEG–Im-PI][Br]) exhibited a high Tg (>300 °C, Table 1 and Fig. 3), even in the presence of the pendant PEG–imidazolium groups. These Tg values were much higher than those obtained from poly(IL)s or other typical glassy polymers, demonstrating that these newly developed materials (PEG–Im-PIs 1) were glassy in nature. The Tg values increased with the PEG chain length in the PEG–imidazolium cation, possibly due to the increased crystallization of the PEG component.
Table 1 Physical parameters that characterized the PEG–Im-PI membranes
Membrane |
Tg (°C) |
d-Spacing (Å) |
Density |
6FDA–durene |
424 (ref. 14) |
6.7 |
1.33 |
[C2PEG–Im-PI][Br] |
320 |
6.4 |
1.37 |
[C4PEG–Im-PI][Br] |
325 |
6.3 |
1.38 |
[C8PEG–Im-PI][Br] |
335 |
6.1 |
1.41 |
[C12PEG–Im-PI][Br] |
340 |
5.9 |
1.45 |
 |
| Fig. 3 DSC curves of the PEG–Im-PI membranes. | |
Wide-angle X-ray scattering (WAXS) results obtained from the PEG–Im-PIs revealed that the intersegmental (d-) spacings between the polymer chains in the membranes decreased as the PEG chain length increased (Table 1 and Fig. 4). These decreases in the d-spacings were associated with an increase in the densities of the PEG–Im-PIs. [C12PEG–Im-PI][Br], characterized by the longest PEG chain length, yielded the densest film among the four polymers tested (Table 1). The high density of the PEG–Im-PIs with longer PEG chain lengths were expected to enhance the resistance of the PEG–Im-PI membrane substructure to gas transport by reducing the diffusivity coefficients. A reduced permeability was expected to accompany an increase in the PEG chain length (vide infra).
 |
| Fig. 4 Wide-angle X-ray diffraction plots obtained from the PEG–Im-PI membranes. | |
Mechanical properties
The mechanical properties of the four PEG–Im-PI membranes at 50% RH showed excellent tensile strengths up to 73.2 MPa, with Young's moduli as high as 2.1 GPa (Fig. S7 in ESI† and Table 2). The mechanical strengths of the PEG–imidazolium-functionalized PIs indicated that these membranes were appropriate for gas separation measurements.
Table 2 Tensile properties of the PEG–Im-PI membranes
Membrane |
Maximum tensile strength, MPa |
Elongation at break, % |
Young's modulus, GPa |
6FDA–durene |
76.7 |
10.4 |
1.7 |
[C2PEG–Im-PI][Br] |
73.2 |
7.9 |
2.1 |
[C4PEG–Im-PI][Br] |
71.7 |
7.5 |
2.0 |
[C8PEG–Im-PI][Br] |
66.6 |
5.5 |
1.9 |
[C12PEG–Im-PI][Br] |
57.1 |
5.2 |
1.8 |
Gas separation properties
The pure gas permeabilities and permselectivities of the PEG–Im-PIs 1 were measured at 2 atm and 30 °C using the constant-volume/variable-pressure method, and the results are summarized in Table 3. The results were compared with those obtained from the pristine PI 1 (6FDA–durene PI). The permeabilities of the PEG–Im-PI membranes to all gases decreased dramatically as the PEG chain length increased (Table 3). This trend corresponded to a significant decrease in the diffusivity coefficients (Fig. 5a and Table 4) caused by a decrease in the interchain spacing (lower d-spacings) with increasing PEG chain length (Table 1, vide supra). In other words, the presence of longer PEG chains in the PEG–Im-PIs further reduced the free volume of the polymers, possibly due to an increase in the crystallinity of the PEG (higher Tg values were observed for the PIs with longer PEG chains), which decreased the gas permeability of these materials. In fact, the diffusivity coefficients of the PEG–imidazolium-functionalized PI were much lower than the value obtained from the pristine 6FDA–durene PI 1. For example, the CO2 diffusivities of the PEG–Im-PIs decreased by factors of 2, 3, 6, and 10 for [C2PEG–Im-PI][Br], [C4PEG–Im-PI][Br], [C8PEG–Im-PI][Br], [C12PEG–Im-PI][Br], respectively, compared to the 6FDA–durene PI 1 (Table 4).
Table 3 Gas permeability (P) and permselectivity (α) of the pristine PI and PEG–Im-PI membranes at 2 atm and 35 °Ca
Membrane |
PCO2 |
PN2 |
PCH4 |
αCO2/N2 |
αCO2/CH4 |
P in Barrers, where 1 Barrer = 10−10 [cm3 (STP) cm] (cm−2 s−1 cm−1 Hg−1). |
6FDA–durene |
495 |
41.1 |
37.3 |
12.1 |
13.2 |
[C2PEG–Im-PI][Br] |
484.8 |
18.1 |
13.2 |
26.7 |
36.7 |
[C4PEG–Im-PI][Br] |
284.3 |
8.7 |
6.3 |
32.4 |
45.1 |
[C8PEG–Im-PI][Br] |
107.9 |
3.0 |
2.2 |
35.1 |
48.4 |
[C12PEG–Im-PI][Br] |
54.1 |
1.3 |
1.1 |
40.1 |
49.2 |
 |
| Fig. 5 Diffusivity coefficients (a) and solubility coefficients (b) of the PEG–Im-PI membranes as a function of the PEG chain length. | |
Table 4 Gas diffusivity coefficientsa and solubility coefficientsb at 2 atm and 35 °C
Membrane |
DCO2 |
DN2 |
DCH4 |
SCO2 |
SN2 |
SCH4 |
SCO2/N2 |
SCO2/CH4 |
Diffusivity coefficient (10−8 cm2 s−1). Solubility coefficient (cm3 (STP) cm−3 cm−1 Hg−1). |
6FDA–durene |
29.1 |
18.1 |
6.57 |
0.17 |
0.022 |
0.057 |
7.49 |
2.98 |
[C2PEG–Im-PI][Br] |
15.9 |
9.03 |
3.31 |
0.31 |
0.020 |
0.039 |
15.5 |
7.79 |
[C4PEG–Im-PI][Br] |
9.87 |
5.11 |
1.72 |
0.29 |
0.017 |
0.036 |
17.0 |
7.90 |
[C8PEG–Im-PI][Br] |
5.14 |
3.24 |
1.25 |
0.21 |
0.009 |
0.018 |
23.3 |
11.6 |
[C12PEG–Im-PI][Br] |
2.80 |
1.84 |
0.84 |
0.19 |
0.007 |
0.013 |
27.5 |
14.8 |
On the other hand, all PEG–Im-PIs showed much higher solubilities toward CO2 than to other gases (CH4 and N2, Fig. 5b and Table 4). These behaviors were attributed to the effects of both the polar PEG and imidazolium-based IL groups. The extremely high CO2 solubility of the PEG–imidazolium-functionalized PI yielded a high CO2 permeability to these newly developed PEG–Im-PI membranes 1 (Table 3). The solubility (in addition to the diffusivity) is another contributor to the permeability. In fact, a CO2 permeability as high as 484.8 Barrer was obtained for [C2PEG–Im-PI][Br] due to its extremely high CO2 solubility (0.31, Table 4). This permeability was much higher than the corresponding values obtained from other PEO-based copolymers or poly(IL)s and approached the value measured from pristine 6FDA–durene PI 1 (495 Barrer, Table 4). The solubility toward CO2 was the highest for [C2PEG–Im-PI][Br]. This value decreased rapidly as the PEG chain length increased. This trend was thought to originate from a decrease in the free volume within the polymer matrix as the PEG chain length increased because the free volume can affect the diffusivity as well as the solubility.35 Similarly, a gradual decrease in the solubility toward other gases (N2 and CH4) was observed as the PEG chain length increased (Table 4 and Fig. 5b).
The ability of the PEG–Im-PIs 1 to separate other nonpolar gases (N2 and CH4), appeared to be normal, unlike the CO2 separation trends, and these materials yielded excellent CO2/N2 and CO2/CH4 permselectivities. As a result, [C2PEG–Im-PI][Br] showed high selectivities of 26.7 and 36.7 for CO2/N2 and CO2/CH4 while maintaining a high CO2 permeability of 484.8 Barrer. The high permeability and high selectivity of [C2PEG–Im-PI][Br] was ascribed to a combination of the enhanced solubility of CO2 in this membrane and the relatively small loss in CO2 diffusivity compared to the pristine PI (Fig. 5a and b), which in turn is caused by our structure being unique in having a pendant PEG and IL, both as a CO2-philic group, onto the diffusivity selective rigid PI backbone.
The CO2/N2 and CO2/CH4 permselectivities increased as the PEG chain length increased due to an enhancement in the solubility selectivities (Table 4). The solubility selectivity increased whereas the diffusivity selectivity remained almost constant as the PEG chain length increased (Fig. 6a and b).
 |
| Fig. 6 Diffusivity selectivity (a) and solubility selectivity (b) of the PEG–Im-PI membranes as a function of the PEG chain length. | |
Permeability vs. selectivity among the PEG–ImPIs
The permeability–selectivity tradeoff results obtained for the CO2/CH4 (Fig. 7a) and CO2/N2 (Fig. 7b) separations in the membranes prepared from PEG–Im-PI with different PEG chains were compared by plotting a Robeson plot for each membrane.6,7 Data obtained from the SILMs (supported ionic liquid membranes)14 and poly(IL)s16,17,36 are included for comparison.
 |
| Fig. 7 “Robeson upper bond 2008” plot for comparing the CO2/CH4 (left) and CO2/N2 (right) separation performances of the PEG–Im-PIs ([C2PEG–Im-PI][Br] (a), [C4PEG–Im-PI][Br] (b), [C8PEG–Im-PI][Br] (c), and [C12PEG–Im-PI][Br] (d)) with other previously reported SILMs and poly(IL)s. Data were taken from [6, 7, 14, 16, 17, 32]. | |
All PEG–Im-PI membranes exhibited outstanding CO2/CH4 performances with curves positioned above the Robeson upper bound of 1991. Moreover, the [C2PEG–Im-PI][Br] and [C4PEG–Im-PI][Br] membrane curves appeared above even the 2008 upper bound. Although all PEG–Im-PI membranes fell below the upper bound line for CO2/N2, their performances were intermediate between those of the SILMs and the poly(IL)s, and they outperformed the poly(IL)s. These results indicated that the PEG–imidazolium-functionalized PI polymers displayed remarkable capacities for enhanced selective separation of CO2.
Conclusions
We prepared a series of novel PEG–imidazolium-functionalized polyimides and successfully demonstrated the utility of the corresponding membranes for CO2 gas separation. Imidazolium-based IL and polar PEG groups, which are highly CO2-solubilizing groups, have previously been incorporated into highly permeable rigid polymers, such as PI, as a strategy for increasing the CO2 selectivity of polymer-membrane-based gas separation membranes. This is the first example of the incorporation of both a PEG and an ionic group onto the rigid PI polymer unit to serve as a CO2-solubilizing group. We investigated the effects of the PEG chain length within the PEG–imidazolium groups on the structures and properties of the polymers, as well as the gas separation properties of the corresponding polymer membranes. The PEG–imidazolium-functionalized PI (PEG–Im-PI) membranes showed higher solubilities toward CO2 than to other gases (CH4 and N2), leading to a high CO2 selectivity while maintaining a good permeability. Excellent mechanical and thermal properties were measured for these newly developed polymers. Overall, these materials were found to be useful for preparing excellent low-cost highly energy efficient polymer membranes for CO2 separation.
Acknowledgements
This work was supported by the Korea Carbon Capture and Sequestration R&D Center under the Korea CCS2020 Program of the Ministry of Education and Science and Technology, Republic of Korea.
Notes and references
- W. J. Schell, J. Membr. Sci., 1985, 22, 217–224 CrossRef CAS.
- W. J. Koros and G. K. Fleming, J. Membr. Sci., 1993, 83, 1–80 CrossRef CAS.
- P. Bernado, E. Drioli and G. Golemme, Ind. Eng. Chem. Res., 2009, 48, 4638–4663 CrossRef.
- N. Du, H. B. Park, M. M. Dal-Cin and M. D. Guiver, Energy Environ. Sci., 2012, 5, 7306–7322 CAS.
- R. W. Baker, Membrane Technology and Applications, J. Wiley, Chichester, New York, 2nd edn, 2004 Search PubMed.
- L. M. Robeson, J. Membr. Sci., 1991, 62, 165–185 CrossRef CAS.
- L. M. Robeson, J. Membr. Sci., 2008, 320, 390–400 CrossRef CAS.
- S. Matteucci, Y. Yampolskii, B. D. Freeman and I. Pinnau, Materials Science of Membranes for Gas and Vapor Separation, John Wiley & Sons, Chichester, 2006, pp. 1–47 Search PubMed.
- J. E. Bara, D. E. Camper, D. L. Gin and R. D. Noble, Acc. Chem. Res., 2010, 43, 152–159 CrossRef CAS PubMed.
- A. B. Rao and E. S. Rubin, Environ. Sci. Technol., 2002, 36, 4467–4475 CrossRef CAS PubMed.
- H. Lin and B. D. Freeman, J. Membr. Sci., 2004, 239, 105–117 CrossRef CAS.
- M. Yoshino, K. Ito, H. Kita and K. Okamoto, J. Polym. Sci., Part B: Polym. Phys., 2000, 38, 1707–1715 CrossRef CAS.
- J. E. Bara, T. K. Carlisle, C. J. Gabriel, D. Camper, A. Finotello, D. L. Gin and R. D. Noble, Ind. Eng. Chem. Res., 2009, 48, 2739–2751 CrossRef CAS.
- P. Scovazzo, J. Membr. Sci., 2009, 343, 199–211 CrossRef CAS.
- P. Cserjési, N. Nemestóthy and K. Belafi-Bakó, J. Membr. Sci., 2010, 349, 6–11 CrossRef.
- J. E. Bara, S. Lessmann, C. J. Gabriel, E. S. Hatakeyama, R. D. Noble and D. L. Gin, Ind. Eng. Chem. Res., 2007, 46, 5397–5404 CrossRef CAS.
- J. E. Bara, C. J. Gabriel, E. S. Hatakeyama, T. K. Carlisle, S. Lessmann, R. D. Noble and D. L. Gin, J. Membr. Sci., 2008, 321, 3–7 CrossRef CAS.
- J. Yuan, D. Mecerreyes and M. Antonietti, Prog. Polym. Sci., 2013, 38, 1009–1036 CrossRef CAS.
- I. Kammakakam, H. W. Kim, S. Nam, H. B. Park and T.-H. Kim, Polymer, 2013, 54, 3534–3541 CrossRef CAS.
- I. Kammakakam, H. W. Yoon, S. Nam, H. B. Park and T.-H. Kim, J. Membr. Sci., 2015, 487, 90–98 CrossRef CAS.
- I. Kammakakam, S. Nam and T.-H. Kim, RSC Adv., 2015, 5, 69907–69914 RSC.
- A. S. Shaplov, S. M. Morozava, E. I. Lonzinskaya, P. S. Vlasov, A. S. L. Gouveia, L. C. Tome, I. M. Marrucho and Y. S. Vygodskii, Polym. Chem., 2016, 7, 580–591 RSC.
- D. Husken, T. Visser, M. Wessling and R. J. Gaymans, J. Membr. Sci., 2010, 346, 194–201 CrossRef CAS.
- K. Okamoto, M. Fujii, S. Okamyo, H. Suzuki, K. Tanaka and H. Kita, Macromolecules, 1995, 28, 6950–6956 CrossRef CAS.
- D. M. Munoz, E. M. Maya, J. de Abajo, J. G. de la Campa and A. E. Lozano, J. Membr. Sci., 2008, 323, 53–59 CrossRef CAS.
- H. Suzuki, K. Tanaka, H. Kita, K. Okamoto, H. Hoshino, T. Yoshinaga and Y. Kusuki, J. Membr. Sci., 1998, 146, 31–37 CrossRef CAS.
- V. I. Bondar, B. D. Freeman and I. Pinnau, J. Polym. Sci., Part B: Polym. Phys., 2000, 38, 2051–2062 CrossRef CAS.
- M. M. Talakesh, M. Sadeghi, M. P. Chenar and A. Khosravi, J. Membr. Sci., 2012, 415–416, 469–477 CrossRef CAS.
- H. W. Kim and H. B. Park, J. Membr. Sci., 2011, 372, 116–124 CrossRef CAS.
- H. Lin, E. Van Wager, J. S. Swinnea, B. D. Freeman, S. J. Pas, A. J. Hill, S. Kalakkunnath and D. S. Kalika, J. Membr. Sci., 2006, 276, 145–161 CrossRef CAS.
- T. Sakaguchi, K. Kameoka and T. Hashimoto, J. Appl. Polym. Sci., 2009, 113, 3504–3509 CrossRef CAS.
- W. Qiu, L. Xu, C.-C. Chen, D. R. Paul and W. J. Koros, Polymer, 2013, 54, 6226–6235 CrossRef CAS.
- W. Qiu, C.-C. Chen, D. R. Paul and W. J. Koros, Macromolecules, 2011, 44, 6046–6056 CrossRef CAS.
- B. Qiu, B. Lin, L. Qiu and F. Yan, J. Mater. Chem., 2012, 22, 1040–1045 RSC.
- L. M. Robeson, Z. P. Smith, B. D. Freeman and D. R. Paul, J. Membr. Sci., 2014, 453, 71–83 CrossRef CAS.
- P. Li, D. R. Paul and T. S. Chung, Green Chem., 2012, 14, 1052–1063 RSC.
Footnote |
† Electronic supplementary information (ESI) available: Details of the characterizations and PEG–imidazoles preparation. See DOI: 10.1039/c6ra00735j |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.