Next Article in Journal
Synthesis and Gas Transport Properties of Poly(2,6-dimethyl-1,4-phenylene oxide)–Silica Nanocomposite Membranes
Next Article in Special Issue
Acidic Gases Separation from Gas Mixtures on the Supported Ionic Liquid Membranes Providing the Facilitated and Solution-Diffusion Transport Mechanisms
Previous Article in Journal
In Situ SAXS Measurement and Molecular Dynamics Simulation of Magnetic Alignment of Hexagonal LLC Nanostructures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Towards Biohydrogen Separation Using Poly(Ionic Liquid)/Ionic Liquid Composite Membranes

by
Andreia S. L. Gouveia
1,2,
Lucas Ventaja
1,
Liliana C. Tomé
2,*,† and
Isabel M. Marrucho
1,2,*
1
Centro de Química Estrutural, Instituto Superior Técnico, Universidade de Lisboa, Avenida Rovisco Pais, 1049-001 Lisboa, Portugal
2
Instituto de Tecnologia Química e Biológica António Xavier, Universidade Nova de Lisboa, Av. da República, 2780-157 Oeiras, Portugal
*
Authors to whom correspondence should be addressed.
Present address: POLYMAT, University of the Basque Country UPV/EHU, Joxe Mari Korta Center, Avda. Tolosa 72, 20018 Donostia-San Sebastian, Spain.
Membranes 2018, 8(4), 124; https://doi.org/10.3390/membranes8040124
Submission received: 19 October 2018 / Revised: 28 November 2018 / Accepted: 28 November 2018 / Published: 2 December 2018

Abstract

:
Considering the high potential of hydrogen (H2) as a clean energy carrier, the implementation of high performance and cost-effective biohydrogen (bioH2) purification techniques is of vital importance, particularly in fuel cell applications. As membrane technology is a potentially energy-saving solution to obtain high-quality biohydrogen, the most promising poly(ionic liquid) (PIL)–ionic liquid (IL) composite membranes that had previously been studied by our group for CO2/N2 separation, containing pyrrolidinium-based PILs with fluorinated or cyano-functionalized anions, were chosen as the starting point to explore the potential of PIL–IL membranes for CO2/H2 separation. The CO2 and H2 permeation properties at the typical conditions of biohydrogen production (T = 308 K and 100 kPa of feed pressure) were measured and discussed. PIL–IL composites prepared with the [C(CN)3] anion showed higher CO2/H2 selectivity than those containing the [NTf2] anion. All the membranes revealed CO2/H2 separation performances above the upper bound for this specific separation, highlighting the composite incorporating 60 wt % of [C2mim][C(CN)3] IL.

1. Introduction

Due to its outstanding intrinsic features, hydrogen (H2) is considered to be the energy carrier of the future. Besides being the simplest element in the universe, the H2 molecule has the highest energy content per unit weight of any known fuel. However, H2 is not a primary fuel source, which means that it is not available in nature and thus needs to be produced [1]. Hydrogen has been produced mainly on an industrial scale from fossil fuels, through natural gas reforming or coal gasification, and from water, using electrolysis in which water (H2O) can be split into hydrogen and oxygen (O2) [2]. Although water splitting is an ecologically clean process compared to the previously mentioned H2 production processes, it is a highly energy-demanding technology [3].
Hydrogen production using biological processes has been attracting growing attention since it is more environmentally friendly and less energy intensive than the other described H2 production systems because its conditions are close to room temperature (303–313 K) and ambient pressure (100 kPa) [2]. BioH2 production can be divided into two main categories: light-dependent (direct or indirect biophotolysis and photo fermentation) and -independent (dark fermentation) methodologies [4,5,6]. The dark fermentation process has several advantages compared to the other biological processes namely, its potential for cost-effective hydrogen production, the high rate of cell growth, and the non-requirement of light energy [6]. In spite of the recognized potential of biohydrogen for sustainable development, there are still issues regarding its production and purification [7], such as the elimination of CO2, N2, and other impurities (H2O and H2S), so that an enriched H2 stream can be obtained for efficient energy generation, mostly through fuel cells [8]. Among the several methodologies for separating hydrogen, such as pressure swing adsorption (PSA), cryogenic distillation, and membrane separation, the first two are designed mainly for large-scale hydrogen production and cannot be used without modification for an upgrade of relatively small amounts of biohydrogen [9]. Thus, membrane technology has been reported as an attractive alternative for biohydrogen separation and purification [10] since it can be introduced easily into hydrogen-producing bioreactors, leading to an integrated process of bioH2 production and purification [11,12], not omitting its important engineering and economic advantages. Particularly, polymeric membranes, such as polysulfone (PSF) and polyimide (PI) [13], have been considered a suitable choice for biohydrogen separation as they can not only be used at the bioreactors’ operating conditions but they also have low cost, high energy efficiency, and a smaller ecological footprint than conventional separation processes [14,15,16].
Few articles have been published using membranes for bioH2 separation. Among them, the combination of different polymers and ionic liquids (ILs) to prepare polymer/IL composite membranes is one of the most promising approaches, which takes into account the inherent CO2-philicity of ILs [16]. For instance, Kanehashi et al. [17] prepared different membranes based on a fluorine-containing polyimide and different amounts of [C4mim][NTf2] IL up to 81 wt %, and studied their gas separation performance at 308 K and 100 kPa of feed pressure. The highest CO2/H2 permselectivity (6.6) was obtained for the membrane that contained the maximum amount of IL (81 wt %) [17]. Moreover, Friess et al. [18] studied the gas permeation properties at 298 K and 100 kPa of feed pressure through membranes composed of poly(vinylidene fluoride-co-hexafluoropropylene) and 20 to 80 wt % of [C2mim][NTf2] IL. Again, the largest CO2 permeability (533 Barrer) and CO2/H2 permselectivity (12.1) were obtained when the highest amount of IL (80 wt %) was used [18].
With the aim of designing efficient CO2/H2 separation membranes, this work explores the use of poly(ionic liquid)s (PILs), which are well recognized by their high CO2 affinity and designer nature [19]. Different approaches have been proposed to produce PIL-based CO2 separation membranes, such as neat PIL membranes [20,21,22,23], PIL copolymer membranes [19,24], and the incorporation of ILs alone or together with nanoporous materials as fillers, including zeolites or metal-organic frameworks (MOFs), to form PIL/IL/filler mixed matrix membranes (MMMs) [25,26,27]. Among all these approaches, the blend of PILs and ILs to produce homogeneous PIL–IL composite membranes is the most promising due to their high CO2 separation performance, as well as the good mechanical stability of the membranes [28]. Notwithstanding the potential of PILs for CO2 separation [28,29,30,31,32], only a reduced number of works concerning CO2/H2 separation have been reported in the literature. For instance, Carlisle et al. [33] explored the CO2/H2 separation through imidazolium PIL–IL gel membranes produced by UV polymerization. The time-lag experiments performed at room temperature and 200 kPa of feed pressure showed ideal CO2/H2 selectivities ranging from 6.6 to 12 for membranes prepared with 5 to 100 mol% of a cross-linking monomer and different amounts of free IL and IL monomer. Their best result (CO2 permeability of 540 Barrer and CO2/H2 selectivity of 12) was achieved for the composite containing 100 mol% of cross-linking monomer and 75 wt % of IL [33]. Moreover, Kharul et al. [34] studied the CO2 and H2 separation properties of polybenzimidazole-based PILs. The different synthesized polybenzimidazole-based PILs exhibited very similar CO2 and H2 permeabilities (<30 Barrer) and, consequently, CO2/H2 selectivities approximately equal to 1 [34].
Amongst the PIL–IL membranes developed so far for CO2/N2 separation, the most widely explored are those composed of imidazolium-based PILs with fluorinated or cyano-functionalized anions [35,36,37]. However, our group reported PIL–IL membranes made of pyrrolidinium-based PILs combining the same anions, which are particularly simple to prepare through a metathesis reaction of a commercially available polyelectrolyte. The PIL–IL membranes displayed CO2/N2 separation performances near or even above the Robeson upper bound [38,39,40,41]. In fact, the CO2-phylic behavior of the [NTf2] anion and the CO2 separation efficiency of the [C(CN)3] anion [42] motivated us to explore the most promising pyrrolidinium-based PIL–IL composites based on these two anions for CO2/N2 separation, now for CO2/H2 separation.
In this work, solvent casting method was used to prepare composite membranes composed of two pyrrolidinium-based PILs: poly([Pyr11][NTf2]) and poly([Pyr11][C(CN)3]). The poly([Pyr11][NTf2]) was blended with 40 and 60 wt % of the structurally similar [C4mpyr][NTf2] IL and also with 40 wt % of an imidazolium-based IL ([C2mim][NTf2]), while poly([Pyr11][C(CN)3]) was mixed with 40 and 60 wt % of [C2mim][C(CN)3] IL. The chemical structures of the PILs and ILs used are shown in Figure 1. The CO2 and H2 permeation properties (permeability, diffusivity and solubility) were determined at two different temperatures (T = 293 K and T = 308 K) under a transmembrane pressure differential of 100 kPa. A temperature of 293 K was used first so that the results obtained herein could be compared to those previously reported by our group, while T = 308 K was chosen to reproduce the hydrogen bioproduction conditions [13].

2. Experimental Section

2.1. Materials

Poly(diallyldimethylammonium) chloride solution (average Mw 400,000–500,000, 20 wt % in water), acetone (99.8%), and acetonitrile (99.8%) were purchased from Sigma-Aldrich (St. Louis, MO, USA). Lithium bis(trifluoromethylsulfonyl)imide (LiNTf2, 99 wt % pure) and sodium tricyanomethanide (NaC(CN)3, 98 wt % pure) were supplied by IoLiTec GmbH (Heilbronn, Germany). The PILs used were previously synthesized by anion metathesis reactions, as described in previous studies [39,43]. All the starting materials used for PIL syntheses, as well as the organic solvents, were used as received. The water was double distilled. The ILs, 1-ethyl-3-methylimidazolium tricyanomethanide ([C2mim][C(CN)3], >98 wt % pure), 1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([C2mim][NTf2], >99 wt % pure), and 1-butyl-1-methylpyrrolidinium bis(trifluoromethylsulfonyl)imide ([C4mpyr][NTf2], >99 wt % pure) were provided by IoLiTec GmbH. Carbon dioxide (CO2) and hydrogen (H2) were supplied by Air Liquide with a minimum purity of 99.99%. Gases were used with no further purification.

2.2. Preparation of PIL–IL Membranes

Several free-standing membranes composed of the synthesized PILs and specific quantities of different ILs containing the same anions were produced by solvent casting. The first step was the preparation of 6 (w/v)% and 12 (w/v)% solutions of poly([Pyr11][C(CN)3]) and poly([Pyr11][NTf2]), respectively, in the most suitable solvents and the addition of the respective IL amounts (Table 1). The solutions were mixed using a magnetic stirrer until complete dissolution of the PIL and IL components and then poured into Petri dishes for slow evaporation of the solvent at room temperature. With the aim of obtaining homogeneous membranes, the solvent evaporation took place slowly, for 2/3 days, depending on the solvent used (Table 1), and in a saturated solvent environment. The thicknesses of the prepared membranes (70–210 μm) were measured using a digital micrometer (Mitutoyo, model MDE-25PJ, Kanagawa, Japan). Average thickness was calculated from six measurements taken at different locations of each PIL–IL membrane. All the PIL–IL membranes studied were considered to have good stability since they were malleable and flexible enough to be used in the gas permeation measurements, even for the composites with 60 wt % of IL. Moreover, the evaluation of the mechanical stability of the PIL–IL composite membranes having the [C(CN)3] anion was recently reported by Tomé et al. [44] (Young’s modulus (PIL–40IL) ~14 MPa; Young’s modulus (PIL–60IL) ~5).

2.3. Gas Permeation Experiments

A time lag equipment described in detail elsewhere [38] was used to measure and determine the ideal CO2 and H2 permeabilities and diffusivities through the prepared PIL–IL composites. Initially, each membrane was degassed under vacuum inside the permeation cell for at least 12 h before testing. The gas permeation experiments were performed at 293 K and 308 K with an upstream pressure of 100 kPa (feed) and vacuum (<0.1 kPa) as the initial downstream pressure (permeate). Three separate CO2 and H2 experiments on a single composite membrane were measured. Between each run, the permeation cell and lines were evacuated until the pressure was below 0.1 kPa.
The gas transport through the PIL–IL membranes was assumed to occur according to the solution-diffusion mass transfer mechanism [45]. Thus, the permeability (P) is related to diffusivity (D) and solubility (S) as follows:
P = D × S
The permeate flux of each studied gas (Ji) was experimentally determined using Equation (2) and assuming an ideal gas behavior and a homogeneous membrane [46]:
J i = V p Δ p d A t R T
where V p is the permeate volume, Δ p d is the variation of downstream pressure, A is the effective membrane surface area, t is the experimental time, R is the gas constant, and T is the temperature. Equation (3) was then used to calculate the ideal gas permeability (Pi) from the pressure driving force ( Δ p i ) and membrane thickness ( ):
P i = J i Δ p i /
Gas diffusivity (Di) was determined according to Equation (4). The time-lag parameter (θ) was deduced by extrapolating the slope of the linear portion of the pd vs. t curve back to the time axis, where the intercept is equal to θ [47]:
D i = 2 6 θ
After defining both Pi and Di, the gas solubility (Si) was also calculated using Equation (1).
The ideal permeability selectivity (or permselectivity), α i / j , which can also be expressed as the product of the diffusivity selectivity and the solubility selectivity, was obtained by dividing the permeability of the more permeable species i to the permeability of the less permeable species j, as expressed in Equation (5):
α i / j = P i P j = ( D i D j ) × ( S i S j )

3. Results and Discussion

3.1. CO2 and H2 Permeation Properties

3.1.1. Gas Permeability (P)

The CO2 and H2 permeabilities through the PIL–IL composite membranes that were studied are presented in Figure 2. The CO2 permeability was always higher than that of H2 and both permeabilities increased with increasing temperature, although the increment was not the same between the studied composites, varying from 15 to 50% for CO2 permeability values and from 39 to 77% for H2 permeabilities. The CO2 permeabilities at 293 K for all the membranes discussed here are in good agreement with those already reported [38,39,41], which emphasizes the high reproducibility of the method used. As expected, the incorporation of high amounts of IL led to enhanced CO2 and H2 permeabilities. Additionally, at 308 K, the temperature of bioH2 purification, the PIL NTf2–40 [C2mim][NTf2] membrane presented similar CO2 and H2 permeability values to those of PIL NTf2–60 [C4mpyr][NTf2]. This means that despite the higher structural compatibility of [C4mpyr][NTf2] with the pyrrolidinium-based PIL, the imidazolium-based cation of the IL is determinant in promoting increased gas permeabilities. However, and as already reported by our group [41], the use of [C2mim][NTf2] instead of [C4mpyr][NTf2] only allowed the incorporation of free IL up to 40 wt % so that a mechanically stable and homogeneous membrane could be obtained. Considering the PIL–IL membranes that comprise the [C(CN)3] anion in both PIL and IL, the PIL C(CN)3–60 [C2mim][C(CN)3] composite showed the highest CO2 permeability (505 Barrer) at 308 K but not the highest H2 permeability (40.3 Barrer), which was obtained for the PIL NTf2–60 [C4mpyr][NTf2] composite membrane (46.0 Barrer). Moreover, CO2 permeability increased about 42% while H2 permeability increased approximately 57% when the IL content in the PIL C(CN)3–[C2mim][C(CN)3] composite varied from 40 to 60 wt %. A significant difference in CO2 permeability (76%) was obtained for the PIL–IL membranes that contained [C4mpyr][NTf2] IL when the IL amount increased from 40 to 60 wt %, while the increment in H2 permeability was only around 34%.

3.1.2. Gas Diffusivity (D)

The experimental gas diffusivity results at 293 K and 308 K through the prepared membranes are listed in Table 2. A high difference (one or, in some cases, two orders of magnitude) between CO2 and H2 diffusivity values, which corresponds to CO2/H2 diffusivity selectivities around 0.1, was observed. This difference in gas diffusivities was mainly due to the smaller size of H2 (2.89 Å) compared to CO2 kinetic diameter (3.30 Å) [48]. Moreover, both CO2 and H2 diffusivity increased with increasing temperature and with increasing IL content in the PIL–IL composite (Table 2). The same behavior was also found for CO2 and H2 permeabilities (Figure 2). From Table 2, it can also be seen that CO2 and H2 diffusivities through the prepared membranes can be ordered as follows: PIL NTf2–40 [C4mpyr][NTf2] < PIL NTf2–60 [C4mpyr][NTf2] < PIL NTf2–40 [C2mim][NTf2] < PIL C(CN)3–40 [C2mim][C(CN)3] < PIL C(CN)3–60 [C2mim][C(CN)3], which means that the presence of the [C(CN)3] anion in the composites, either in the PIL or IL’s structure, leads to higher CO2 and H2 diffusivities compared to the [NTf2] anion. The same trend was also observed for N2 diffusivities [38,39,41]. Thus, and as expected, it can be concluded that gas diffusivities follow the order of the kinetic diameters CO2 < N2 < H2. It can also be noted that the presence of imidazolium-based cation ([C2mim]+) in the ILs led to superior gas diffusivities compared to the pyrrolidinium-based cation ([C4mpyr]+).
Another interesting point is the comparison between gas permeability and diffusivity behaviors. Regardless of the anion, although the composite that comprised 60 wt % of IL had the highest H2 diffusivities (>1200 m2 s−1 at 308 K), it did not present the highest H2 permeabilities (Figure 2). An equivalent behavior was also obtained for the PIL NTf2–40 [C4mpyr][NTf2] composite membrane, which displayed the lowest H2 diffusivities (546 m2 s−1 at 308 K) but not the lowest H2 permeabilities. In the case of CO2, it can be seen from Table 2 and Figure 2 that CO2 permeability followed the same trend as CO2 diffusivity, with the exception of the PIL NTf2–60 [C4mpyr][NTf2] and PIL C(CN)3–40 [C2mim][C(CN)3] membranes. This behavior led us to conclude that the very high difference (three or, in some cases, four orders of magnitude) among H2 diffusivities is somehow compensated by a reverse behavior in H2 solubilities (as will be discussed in the next section), which has a significant impact on the H2 permeability results.

3.1.3. Gas Solubility (S)

Gas solubility (S) values were calculated using Equation (1) at temperatures of 293 K and 308 K and are displayed in Figure 3. The CO2 solubility decreased with increasing temperature while H2 solubility increased with increasing temperature for all the PIL–IL membranes studied. Analogous reverse H2 solubility behavior with temperature was also found and discussed by Raeissi et al. [49] in imidazolium-based ILs, such as [C4mim][NTf2], which means that hydrogen dissolves better at higher than at lower temperatures. This trend seems to be the general rule for H2 solubility in ILs [49,50,51,52,53] and has been attributed to the extreme lightness and small intermolecular forces of hydrogen molecules [49].
It can also be observed that, as expected, both CO2 and H2 solubilities were enhanced with the incorporation of high amounts of IL in the composite. For instance, when the amount of [C2mim][C(CN)3] increased from 40 to 60 wt %, the CO2 and H2 solubilities at 293 K increased almost 50% and 15%, respectively, while at 308 K the increment in CO2 and H2 solubilities was around 39% and 5%, respectively. Similar behavior was found for the PIL–IL composites comprising the [C4mpyr][NTf2] IL. Moreover, the large differences between CO2 and H2 solubilities can be explained by the high CO2 critical temperature (CO2, 31 °C; H2, −240 °C), corresponding to the superior condensability of CO2 (Tε/k = 195.2 K) compared to H2 (Tε/k = 59.7 K) [48,54]. The fact that H2 can almost be considered an ideal gas due to its small kinetic diameter and non-interacting nature, while CO2 displays a higher kinetic diameter and a quadrupole moment, also plays a role in the difference in solubilities of the two gases. For the PIL–IL composites studied in this work at 308 K, the CO2 solubility ranged from 14 to 28.5 (×10−6) m3(STP)·m−3·Pa−1 whereas the H2 solubility values were two orders of magnitude lower, varying from 0.17 to 0.51 (×10−6) m3(STP)·m−3·Pa−1. Among all the tested membranes, the PIL NTf2–60 [C4mpyr][NTf2] composite presented the highest CO2 and H2 solubilities at both 293 and 308 K. Regarding the influence of the anion’s structure and considering the same amount of free IL in the composite, it can be concluded that the presence of the [NTf2] anion in the PIL–IL membranes leads to higher CO2 and H2 solubilities compared to those membranes comprising the [C(CN)3] anion. As mentioned before, this behavior masks the higher H2 diffusivities of composites with a cyano-functionalized anion, somehow explaining the low influence of H2 diffusivity on the H2 permeability results.

3.2. CO2/H2 Separation Performance

The CO2 and H2 permeabilities and the ideal CO2/H2 permselectivities determined at 293 K and 308 K are summarized in Table 3. Amongst the PIL−IL membranes studied, those bearing the [C(CN)3] anion revealed slightly higher CO2/H2 permselectivities than those containing the [NTf2] anion. This behavior was also observed in our previous works concerning the use of PIL–IL membranes for CO2/N2 and CO2/CH4 separation [38,39,41]. Moreover, from Table 3, it can be seen that CO2/H2 permselectivities decreased with increasing temperature. This result can be explained by the variations in CO2/H2 solubility selectivity with temperature, which leads to a decrease in CO2/H2 permselectivity as the temperature increases [55]. In fact, CO2/H2 solubility selectivity through the studied composite membranes decreased from 78–145 at 293 K to 42–84 at 308 K. It can also be emphasized that CO2/H2 permselectivity seems to be controlled by a solubility mechanism, since CO2/H2 diffusivity selectivity (D CO2/H2) values were approximately equal to 0.1 at both temperatures while solubility selectivity (S CO2/H2) values ranged from 78–145 at 293 K and 42–84 at 308 K.
The gas separation performance of the studied PIL–IL membranes is shown in Figure 4, where the CO2/H2 permselectivity is plotted against the permeability of the more permeable gas species (CO2). This graph displays a tradeoff (black line) between gas permeability and permselectivity. These upper bound limits for several gas pairs were first developed by Robeson [56] who correlated data obtained from measurements carried out with polymeric membranes at low temperatures (298–308 K). Later, Rowe et al. [55] studied the influence of temperature on the tradeoff between gas permeability and permselectivity for different gas pairs. Thus, the upper bound at 300 K developed by Rowe et al. [55] for the CO2/H2 gas pair is represented in Figure 4 and was used to evaluate the performance of the studied PIL–IL membranes for biohydrogen purification (T = 308 K and 100 kPa).
Figure 4 clearly shows that all the PIL–IL membranes that were studied displayed CO2/H2 separation performances above the upper bound. The best CO2/H2 separation performance was obtained for the membrane composed of poly([Pyr11][C(CN)3]) and 60 wt % of [C2mim][C(CN)3] IL, which is in agreement with what has been observed in our recent works regarding the use of PIL–IL composites for CO2/N2 separation [39]. Literature data points for other reported PIL–IL membranes are also plotted in Figure 4 for comparison. The gas permeation measurements reported by Carlisle et al. [33] were performed at room temperature with a transmembrane pressure differential of 200 kPa. Also, their PIL–IL membranes were not prepared by the solvent casting method but through UV polymerization by mixing different percentages of imidazolium-based IL monomers, a cross-linking monomer, and free IL [33]. From Figure 4, it can be seen that the PIL–IL membranes reported in the literature also present CO2/H2 separation performances above the upper bound for the CO2/H2 gas pair at 300 K, but the PIL C(CN)3–60 [C2mim][C(CN)3] membrane studied in this work still revealed superior CO2/H2 separation performance.

4. Conclusions

In this work, dense composite membranes made of pyrrolidinium-based PILs with [C(CN)3] or [NTf2] anions and different amounts of free IL ([C2mim][C(CN)3], [C4mpyr][NTf2] or [C2mim][NTf2]) incorporated were prepared by the solvent casting method. Their CO2 and H2 permeation properties (permeability, diffusivity, and solubility) were determined at biohydrogen production conditions (T = 308 K and 100 kPa of feed pressure) and discussed. The CO2 and H2 permeation properties were measured at T = 293 K and the effect of temperature on gas separation performance was evaluated.
The PIL–IL membranes containing the [NTf2] anion presented the highest H2 permeability and solubility values, while the PIL–IL composites having the [C(CN)3] anion showed the highest H2 diffusivities and CO2/H2 permselectivities. As previously reported, increments in gas permeabilities, diffusivities, and solubilities were observed with increasing temperature and amounts of IL, with the exception of H2 solubility that showed the opposite behavior with temperature compared to what occurred in CO2 solubility. Overall, all the PIL–IL membranes studied revealed similar or superior CO2/H2 separation performance compared to the few PIL–IL composites reported so far in the literature. Particularly, at 308 K, the best result was obtained through the PIL C(CN)3–60 IL C(CN)3 composite membrane (CO2 permeability of 505 Barrer and CO2/H2 selectivity of 12.5), which, as shown in our previous work, also presented remarkable results for CO2/N2 separation.

Author Contributions

Conceptualization, L.C.T. and I.M.M.; Methodology, A.S.L.G.; Gas Separation Measurements and Membrane Preparation, L.V. and A.S.L.G.; Writing-Original Draft Preparation, A.S.L.G.; Writing-Review & Editing, A.S.L.G., L.C.T. and I.M.M.

Funding

This work was supported by FCT through the project PTDC/CTM-POL/2676/2014 and R&D units UID/QUI/00100/2013 (CQE) and UID/Multi/04551/2013 (GreenIT). This project has received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie grant agreement No 745734.

Acknowledgments

Andreia S. L. Gouveia and Liliana C. Tomé are grateful to FCT (Fundação para a Ciência e a Tecnologia) for their Doctoral (SFRH/BD/116600/2016) and Post-doctoral (SFRH/BPD/101793/2014) research grants, respectively.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ΔpdVariation of downstream pressure
ΔpiPressure driving force
AEffective membrane surface area
bioH2Biohydrogen
CO2Carbon dioxide
DDiffusivity
H2Hydrogen
H2OWater
H2SHydrogen sulfide
ILsIonic liquids
JiSteady-state gas flux
Membrane thickness
N2Nitrogen
O2Oxygen
PPermeability
PILsPoly(ionic liquid)s
RIdeal gas law constant
SSolubility
tTime
TTemperature
VpPermeate volume
αi/jPermselectivity
θTime-lag parameter
Cations
[C2mim]+1-ethyl-3-methylimidazolium
[C4mpyr]+1-butyl-3-methylpyrrolidinium
Anions
[NTf2]Bis(tri-fluoromethylsulfonyl)imide
[C(CN)3]Tricyanomethanide

References

  1. Mazloomi, K.; Gomes, C. Hydrogen as an energy carrier: Prospects and challenges. Renew. Sustain. Energy Rev. 2012, 16, 3024–3033. [Google Scholar] [CrossRef]
  2. Das, D.; Veziroǧlu, T.N. Hydrogen production by biological processes: A survey of literature. Int. J. Hydrog. Energy 2001, 26, 13–28. [Google Scholar] [CrossRef]
  3. Kalamaras, C.M.; Efstathiou, A.M. Hydrogen Production Technologies: Current State and Future Developments. Conf. Pap. Energy 2013, 2013, 9. [Google Scholar] [CrossRef]
  4. Das, D.; Veziroglu, T.N. Advances in biological hydrogen production processes. Int. J. Hydrog. Energy 2008, 33, 6046–6057. [Google Scholar] [CrossRef]
  5. Manish, S.; Banerjee, R. Comparison of biohydrogen production processes. Int. J. Hydrog. Energy 2008, 33, 279–286. [Google Scholar] [CrossRef]
  6. Singh, L.; Wahid, Z.A. Methods for enhancing bio-hydrogen production from biological process: A review. J. Ind. Eng. Chem. 2015, 21, 70–80. [Google Scholar] [CrossRef]
  7. Alves, H.J.; Bley Junior, C.; Niklevicz, R.R.; Frigo, E.P.; Frigo, M.S.; Coimbra-Araújo, C.H. Overview of hydrogen production technologies from biogas and the applications in fuel cells. Int. J. Hydrog. Energy 2013, 38, 5215–5225. [Google Scholar] [CrossRef]
  8. Merkel, T.C.; Zhou, M.; Baker, R.W. Carbon dioxide capture with membranes at an IGCC power plant. J. Memb. Sci. 2012, 389, 441–450. [Google Scholar] [CrossRef]
  9. Dunikov, D.; Borzenko, V.; Blinov, D.; Kazakov, A.; Lin, C.Y.; Wu, S.Y.; Chu, C.Y. Biohydrogen purification using metal hydride technologies. Int. J. Hydrog. Energy 2016, 41, 21787–21794. [Google Scholar] [CrossRef]
  10. Sanders, D.F.; Smith, Z.P.; Guo, R.; Robeson, L.M.; McGrath, J.E.; Paul, D.R.; Freeman, B.D. Energy-efficient polymeric gas separation membranes for a sustainable future: A review. Polymer 2013, 54, 4729–4761. [Google Scholar] [CrossRef]
  11. Bakonyi, P.; Nemestóthy, N.; Ramirez, J.; Ruiz-Filippi, G.; Bélafi-Bakó, K. Escherichia coli (XL1-BLUE) for continuous fermentation of bioH2 and its separation by polyimide membrane. Int. J. Hydrog. Energy 2012, 37, 5623–5630. [Google Scholar] [CrossRef]
  12. Bakonyi, P.; Kumar, G.; Nemestóthy, N.; Lin, C.Y.; Bélafi-Bakó, K. Biohydrogen purification using a commercial polyimide membrane module: Studying the effects of some process variables. Int. J. Hydrog. Energy 2013, 38, 15092–15099. [Google Scholar] [CrossRef] [Green Version]
  13. Bakonyi, P.; Nemestóthy, N.; Bélafi-Bakó, K. Biohydrogen purification by membranes: An overview on the operational conditions affecting the performance of non-porous, polymeric and ionic liquid based gas separation membranes. Int. J. Hydrog. Energy 2013, 38, 9673–9687. [Google Scholar] [CrossRef] [Green Version]
  14. Basu, S.; Khan, A.L.; Cano-Odena, A.; Liu, C.; Vankelecom, I.F.J. Membrane-based technologies for biogas separations. Chem. Soc. Rev. 2010, 39, 750–768. [Google Scholar] [CrossRef] [PubMed]
  15. Mohamad, I.N.; Rohani, R.; Mastar, M.S.; Nor, M.T.M.; Md. Jahim, J. Permeation properties of polymeric membranes for biohydrogen purification. Int. J. Hydrog. Energy 2016, 41, 4474–4488. [Google Scholar] [CrossRef]
  16. Pientka, Z.; Peter, J.; Zitka, J.; Bakonyi, P. Application of Polymeric Membranes in Biohydrogen Purification and Storage. Curr. Biochem. Eng. 2014, 1, 99–105. [Google Scholar] [CrossRef]
  17. Kanehashi, S.; Kishida, M.; Kidesaki, T.; Shindo, R.; Sato, S.; Miyakoshi, T.; Nagai, K. CO2 separation properties of a glassy aromatic polyimide composite membranes containing high-content 1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ionic liquid. J. Memb. Sci. 2013, 430, 211–222. [Google Scholar] [CrossRef]
  18. Friess, K.; Jansen, J.C.; Bazzarelli, F.; Izák, P.; Jarmarová, V.; Kačírková, M.; Schauer, J.; Clarizia, G.; Bernardo, P. High ionic liquid content polymeric gel membranes: Correlation of membrane structure with gas and vapour transport properties. J. Memb. Sci. 2012, 415–416, 801–809. [Google Scholar] [CrossRef]
  19. Yuan, J.; Mecerreyes, D.; Antonietti, M. Poly(ionic liquid)s: An update. Prog. Polym. Sci. 2013, 38, 1009–1036. [Google Scholar] [CrossRef]
  20. Jeffrey Horne, W.; Andrews, M.A.; Shannon, M.S.; Terrill, K.L.; Moon, J.D.; Hayward, S.S.; Bara, J.E. Effect of branched and cycloalkyl functionalities on CO2 separation performance of poly(IL) membranes. Sep. Purif. Technol. 2015, 155, 89–95. [Google Scholar] [CrossRef]
  21. Bara, J.E.; Gabriel, C.J.; Hatakeyama, E.S.; Carlisle, T.K.; Lessmann, S.; Noble, R.D.; Gin, D.L. Improving CO2 selectivity in polymerized room-temperature ionic liquid gas separation membranes through incorporation of polar substituents. J. Memb. Sci. 2008, 321, 3–7. [Google Scholar] [CrossRef]
  22. Bara, J.E.; Hatakeyama, E.S.; Gin, D.L.; Noble, R.D. Improving CO2 permeability in polymerized room-temperature ionic liquid gas separation membranes through the formation of a solid composite with a room-temperature ionic liquid. Polym. Adv. Technol. 2008, 19, 1415–1420. [Google Scholar] [CrossRef]
  23. Bara, J.E.; Lessmann, S.; Gabriel, C.J.; Hatakeyama, E.S.; Noble, R.D.; Gin, D.L. Synthesis and Performance of Polymerizable Room-Temperature Ionic Liquids as Gas Separation Membranes. Ind. Eng. Chem. Res. 2007, 46, 5397–5404. [Google Scholar] [CrossRef]
  24. Hu, X.; Tang, J.; Blasig, A.; Shen, Y.; Radosz, M. CO2 permeability, diffusivity and solubility in polyethylene glycol-grafted polyionic membranes and their CO2 selectivity relative to methane and nitrogen. J. Memb. Sci. 2006, 281, 130–138. [Google Scholar] [CrossRef]
  25. Hao, L.; Li, P.; Yang, T.; Chung, T.-S. Room temperature ionic liquid/ZIF-8 mixed-matrix membranes for natural gas sweetening and post-combustion CO2 capture. J. Memb. Sci. 2013, 436, 221–231. [Google Scholar] [CrossRef]
  26. Hudiono, Y.C.; Carlisle, T.K.; LaFrate, A.L.; Gin, D.L.; Noble, R.D. Novel mixed matrix membranes based on polymerizable room-temperature ionic liquids and SAPO-34 particles to improve CO2 separation. J. Memb. Sci. 2011, 370, 141–148. [Google Scholar] [CrossRef]
  27. Hudiono, Y.C.; Carlisle, T.K.; Bara, J.E.; Zhang, Y.; Gin, D.L.; Noble, R.D. A three-component mixed-matrix membrane with enhanced CO2 separation properties based on zeolites and ionic liquid materials. J. Memb. Sci. 2010, 350, 117–123. [Google Scholar] [CrossRef]
  28. Tomé, L.C.; Marrucho, I.M. Ionic liquid-based materials: A platform to design engineered CO2 separation membranes. Chem. Soc. Rev. 2016, 45, 2785–2824. [Google Scholar] [CrossRef] [PubMed]
  29. Sadeghpour, M.; Yusoff, R.; Aroua Mohamed, K. Polymeric ionic liquids (PILs) for CO2 capture. Rev. Chem. Eng. 2017, 33, 183–220. [Google Scholar] [CrossRef]
  30. Qian, W.; Texter, J.; Yan, F. Frontiers in poly(ionic liquid)s: Syntheses and applications. Chem. Soc. Rev. 2017, 46, 1124–1159. [Google Scholar] [CrossRef]
  31. Ajjan, F.N.; Ambrogi, M.; Tiruye, G.A.; Cordella, D.; Fernandes, A.M.; Grygiel, K.; Isik, M.; Patil, N.; Porcarelli, L.; Rocasalbas, G.; et al. Innovative polyelectrolytes/poly(ionic liquid)s for energy and the environment. Polym. Int. 2017, 66, 1119–1128. [Google Scholar] [CrossRef] [Green Version]
  32. Zulfiqar, S.; Sarwar, M.I.; Mecerreyes, D. Polymeric ionic liquids for CO2 capture and separation: Potential, progress and challenges. Polym. Chem. 2015, 6, 6435–6451. [Google Scholar] [CrossRef]
  33. Carlisle, T.K.; Nicodemus, G.D.; Gin, D.L.; Noble, R.D. CO2/light gas separation performance of cross-linked poly(vinylimidazolium) gel membranes as a function of ionic liquid loading and cross-linker content. J. Memb. Sci. 2012, 397–398, 24–37. [Google Scholar] [CrossRef]
  34. Shaligram, S.V.; Wadgaonkar, P.P.; Kharul, U.K. Polybenzimidazole-based polymeric ionic liquids (PILs): Effects of ‘substitution asymmetry’ on CO2 permeation properties. J. Memb. Sci. 2015, 493, 403–413. [Google Scholar] [CrossRef]
  35. Carlisle, T.K.; Wiesenauer, E.F.; Nicodemus, G.D.; Gin, D.L.; Noble, R.D. Ideal CO2/Light Gas Separation Performance of Poly(vinylimidazolium) Membranes and Poly(vinylimidazolium)-Ionic Liquid Composite Films. Ind. Eng. Chem. Res. 2012, 52, 1023–1032. [Google Scholar] [CrossRef]
  36. Li, P.; Paul, D.R.; Chung, T.-S. High performance membranes based on ionic liquid polymers for CO2 separation from the flue gas. Green Chem. 2012, 14, 1052–1063. [Google Scholar] [CrossRef]
  37. Li, P.; Pramoda, K.P.; Chung, T.-S. CO2 Separation from Flue Gas Using Polyvinyl-(Room Temperature Ionic Liquid)–Room Temperature Ionic Liquid Composite Membranes. Ind. Eng. Chem. Res. 2011, 50, 9344–9353. [Google Scholar] [CrossRef]
  38. Tomé, L.C.; Mecerreyes, D.; Freire, C.S.R.; Rebelo, L.P.N.; Marrucho, I.M. Pyrrolidinium-based polymeric ionic liquid materials: New perspectives for CO2 separation membranes. J. Memb. Sci. 2013, 428, 260–266. [Google Scholar] [CrossRef]
  39. Tomé, L.C.; Isik, M.; Freire, C.S.R.; Mecerreyes, D.; Marrucho, I.M. Novel pyrrolidinium-based polymeric ionic liquids with cyano counter-anions: High performance membrane materials for post-combustion CO2 separation. J. Memb. Sci. 2015, 483, 155–165. [Google Scholar] [CrossRef]
  40. Tomé, L.C.; Gouveia, A.S.L.; Freire, C.S.R.; Mecerreyes, D.; Marrucho, I.M. Polymeric ionic liquid-based membranes: Influence of polycation variation on gas transport and CO2 selectivity properties. J. Memb. Sci. 2015, 486, 40–48. [Google Scholar] [CrossRef]
  41. Teodoro, R.M.; Tomé, L.C.; Mantione, D.; Mecerreyes, D.; Marrucho, I.M. Mixing poly(ionic liquid)s and ionic liquids with different cyano anions: Membrane forming ability and CO2/N2 separation properties. J. Memb. Sci. 2018, 552, 341–348. [Google Scholar]
  42. Tomé, L.C.; Florindo, C.; Freire, C.S.; Rebelo, L.P.; Marrucho, I.M. Playing with ionic liquid mixtures to design engineered CO2 separation membranes. Phys. Chem. Chem. Phys. 2014, 16, 17172–17182. [Google Scholar] [PubMed]
  43. Pont, A.-L.; Marcilla, R.; De Meatza, I.; Grande, H.; Mecerreyes, D. Pyrrolidinium-based polymeric ionic liquids as mechanically and electrochemically stable polymer electrolytes. J. Power Sources 2009, 188, 558–563. [Google Scholar] [CrossRef]
  44. Tomé, L.C.; Guerreiro, D.C.; Teodoro, R.M.; Alves, V.D.; Marrucho, I.M. Effect of polymer molecular weight on the physical properties and CO2/N2 separation of pyrrolidinium-based poly(ionic liquid) membranes. J. Memb. Sci. 2018, 549, 267–274. [Google Scholar] [CrossRef]
  45. Wijmans, J.G.; Baker, R.W. The solution-diffusion model: A review. J. Memb. Sci. 1995, 107, 1–21. [Google Scholar] [CrossRef]
  46. Matteucci, S.; Yampolskii, Y.; Freeman, B.D.; Pinnau, I. Transport of Gases and Vapors in Glassy and Rubbery Polymers. In Materials Science of Membranes for Gas and Vapor Separation; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2006; pp. 1–47. [Google Scholar]
  47. Rutherford, S.W.; Do, D.D. Review of time lag permeation technique as a method for characterisation of porous media and membranes. Adsorption 1997, 3, 283–312. [Google Scholar] [CrossRef]
  48. Wang, S.; Li, X.; Wu, H.; Tian, Z.; Xin, Q.; He, G.; Peng, D.; Chen, S.; Yin, Y.; Jiang, Z.; et al. Advances in high permeability polymer-based membrane materials for CO2 separations. Energy Environ. Sci. 2016, 9, 1863–1890. [Google Scholar] [CrossRef]
  49. Raeissi, S.; Peters, C.J. Understanding temperature dependency of hydrogen solubility in ionic liquids, including experimental data in [bmim][Tf2N]. AIChE J. 2012, 58, 3553–3559. [Google Scholar] [CrossRef]
  50. Kumełan, J.; Pérez-Salado Kamps, Á.; Tuma, D.; Maurer, G. Solubility of H2 in the Ionic Liquid [bmim][PF6]. J. Chem. Eng. Data 2006, 51, 11–14. [Google Scholar] [CrossRef]
  51. Finotello, A.; Bara, J.E.; Camper, D.; Noble, R.D. Room-Temperature Ionic Liquids: Temperature Dependence of Gas Solubility Selectivity. Ind. Eng. Chem. Res. 2008, 47, 3453–3459. [Google Scholar] [CrossRef]
  52. Kumełan, J.; Tuma, D.; Pérez-Salado Kamps, Á.; Maurer, G. Solubility of the Single Gases Carbon Dioxide and Hydrogen in the Ionic Liquid [bmpy][Tf2N]. J. Chem. Eng. Data 2010, 55, 165–172. [Google Scholar] [CrossRef]
  53. Raeissi, S.; Florusse, L.J.; Peters, C.J. Hydrogen Solubilities in the IUPAC Ionic Liquid 1-Hexyl-3-methylimidazolium Bis(Trifluoromethylsulfonyl)Imide. J. Chem. Eng. Data 2011, 56, 1105–1107. [Google Scholar] [CrossRef]
  54. Robeson, L.M.; Smith, Z.P.; Freeman, B.D.; Paul, D.R. Contributions of diffusion and solubility selectivity to the upper bound analysis for glassy gas separation membranes. J. Memb. Sci. 2014, 453, 71–83. [Google Scholar] [CrossRef]
  55. Rowe, B.W.; Robeson, L.M.; Freeman, B.D.; Paul, D.R. Influence of temperature on the upper bound: Theoretical considerations and comparison with experimental results. J. Memb. Sci. 2010, 360, 58–69. [Google Scholar] [CrossRef]
  56. Robeson, L.M. The upper bound revisited. J. Memb. Sci. 2008, 320, 390–400. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of the poly(ionic liquid)s (PILs) and ionic liquids (ILs) used in this work to prepare the PIL–IL membranes.
Figure 1. Chemical structures of the poly(ionic liquid)s (PILs) and ionic liquids (ILs) used in this work to prepare the PIL–IL membranes.
Membranes 08 00124 g001
Figure 2. Experimental CO2 and H2 permeabilities (P) through the prepared PIL–IL membranes. Error bars represent standard deviations based on three experimental replicas. In some cases, the standard deviations are very small leading to error bars that cannot be clearly represented.
Figure 2. Experimental CO2 and H2 permeabilities (P) through the prepared PIL–IL membranes. Error bars represent standard deviations based on three experimental replicas. In some cases, the standard deviations are very small leading to error bars that cannot be clearly represented.
Membranes 08 00124 g002
Figure 3. Gas solubilities (S) for the studied PIL–IL membranes at 293 K and 308 K.
Figure 3. Gas solubilities (S) for the studied PIL–IL membranes at 293 K and 308 K.
Membranes 08 00124 g003
Figure 4. CO2/H2 separation performance of the PIL–IL membranes studied. The experimental error is within the data points. Data are plotted on a log–log scale and the upper bound at 300 K was adapted from Rowe et al. [55]. Literature data points ( Membranes 08 00124 i001) from other reported PIL–IL membranes are also displayed for comparison [33].
Figure 4. CO2/H2 separation performance of the PIL–IL membranes studied. The experimental error is within the data points. Data are plotted on a log–log scale and the upper bound at 300 K was adapted from Rowe et al. [55]. Literature data points ( Membranes 08 00124 i001) from other reported PIL–IL membranes are also displayed for comparison [33].
Membranes 08 00124 g004
Table 1. Description of the poly(ionic liquid)–ionic liquid (PIL–IL) membrane’s composition and experimental preparation conditions of the solvent casting procedure.
Table 1. Description of the poly(ionic liquid)–ionic liquid (PIL–IL) membrane’s composition and experimental preparation conditions of the solvent casting procedure.
PIL–IL MembranePolymer (PIL)wt % of ILSolventT (K)Evaporation Time (Days)
PIL C(CN)3–40 [C2mim][C(CN)3]Poly([Pyr11][C(CN)3])40Acetonitrile2983
PIL C(CN)3–60 [C2mim][C(CN)3]60
PIL NTf2–40 [C4mpyr][NTf2]Poly([Pyr11][NTf2])40Acetone2982
PIL NTf2–60 [C4mpyr][NTf2]60
PIL NTf2–40 [C2mim][NTf2]Poly([Pyr11][NTf2])40Acetone2982
Table 2. Experimental gas diffusivities (D) through the studied PIL–IL membranes at T = 293 K and T = 308 K.
Table 2. Experimental gas diffusivities (D) through the studied PIL–IL membranes at T = 293 K and T = 308 K.
PIL–IL MembraneGas Diffusivity (×1012) (m2 s−1) (T = 293 K)Gas Diffusivity (×1012) (m2 s−1) (T = 308 K)
D CO2 ± σD H2 ± σD CO2 ± σD H2 ± σ
PIL C(CN)3–40 [C2mim][C(CN)3]64 ± 1.0970 ± 36.2112 ± 2.51146 ± 34.0
PIL C(CN)3–60 [C2mim][C(CN)3]127 ± 1.11130 ± 5.70181 ± 0.61211 ± 3.2
PIL NTf2–40 [C4mpyr][NTf2]34 ± 0.1484 ± 18.562 ± 1.8546 ± 20.6
PIL NTf2–60 [C4mpyr][NTf2]44 ± 0.7610 ± 6.3076 ± 0.5673 ± 16.9
PIL NTf2–40 [C2mim][NTf2]61 ± 0.4722 ± 1.80106 ± 1.5792 ± 3.70
Table 3. Single gas permeabilities (P) a and ideal permselectivities (α) of the PIL–IL membranes studied b.
Table 3. Single gas permeabilities (P) a and ideal permselectivities (α) of the PIL–IL membranes studied b.
PIL–IL MembraneGas Permeability (Barrer)
(T = 293 K)
Gas Permeability (Barrer)
(T = 308 K)
P CO2 ± σP H2 ± σα CO2/H2P CO2 ± σP H2 ± σα CO2/H2
PIL C(CN)3–40 [C2mim][C(CN)3]139 ± 0.514.5 ± 0.29.6 ± 0.2209 ± 0.925.7 ± 0.18.1 ± 0.1
PIL C(CN)3–60 [C2mim][C(CN)3]438 ± 2.129.1 ± 0.415.1 ± 0.3505 ± 0.340.3 ± 1.112.5 ± 0.3
PIL NTf2–40 [C4mpyr][NTf2]119 ± 0.221.9 ± 0.15.4 ± 0.1164 ± 1.634.4 ± 0.34.8 ± 0.1
PIL NTf2–60 [C4mpyr][NTf2]232 ± 2.229.8 ± 0.17.8 ± 0.1288 ± 1.646.0 ± 0.16.3 ± 0.1
PIL NTf2–40 [C2mim][NTf2]214 ± 0.626.2 ± 0.18.2 ± 0.1287 ± 2.443.8 ± 0.26.5 ± 0.1
a Barrer (1 Barrer = 10−10 cm(STP)3·cm·cm−2·s−1·cm·Hg−1). b The listed uncertainties represent the standard deviations (σ) based on three experiments.

Share and Cite

MDPI and ACS Style

Gouveia, A.S.L.; Ventaja, L.; Tomé, L.C.; Marrucho, I.M. Towards Biohydrogen Separation Using Poly(Ionic Liquid)/Ionic Liquid Composite Membranes. Membranes 2018, 8, 124. https://doi.org/10.3390/membranes8040124

AMA Style

Gouveia ASL, Ventaja L, Tomé LC, Marrucho IM. Towards Biohydrogen Separation Using Poly(Ionic Liquid)/Ionic Liquid Composite Membranes. Membranes. 2018; 8(4):124. https://doi.org/10.3390/membranes8040124

Chicago/Turabian Style

Gouveia, Andreia S. L., Lucas Ventaja, Liliana C. Tomé, and Isabel M. Marrucho. 2018. "Towards Biohydrogen Separation Using Poly(Ionic Liquid)/Ionic Liquid Composite Membranes" Membranes 8, no. 4: 124. https://doi.org/10.3390/membranes8040124

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop