Skip to content
BY 4.0 license Open Access Published by De Gruyter August 13, 2020

Sulfonated poly(arylene ether sulfone) functionalized polysilsesquioxane hybrid membranes with enhanced proton conductivity

  • Rajdeep Mukherjee , Arun Kumar Mandal and Susanta Banerjee EMAIL logo
From the journal e-Polymers

Abstract

Sulfopropylated polysilsesquioxane and –COOH containing fluorinated sulfonated poly(arylene ether sulfone) composite membranes (SPAES-SS-X) have been prepared via an in situ sol–gel reaction through the solution casting technique. The composite membranes showed excellent thermal and chemical stability, compared to the pristine SPAES membrane. The uniform dispersion of the sulfonated SiOPS nanoparticles on the polymer matrix was observed from the scanning electron microscope images. Atomic force microscopy and transmission electron microscopy images indicated significantly better phase-separated morphology and connectivity of the ionic domains of the composite membranes than the pristine SPAES membrane. The composite membranes showed considerable improvement in proton conductivity and oxidative stability than the pristine copolymer membrane under similar test conditions.

Graphical abstract

1 Introduction

Proton exchange membrane fuel cells (PEMFCs) are used in residential and automobile sectors because of their eco-friendly nature and energy conversion efficiency. Proton exchange membrane (PEM) in PEMFCs acts as a separator between electrodes, facilitates proton transfer, and provides a barrier for fuel gas, oxidant, and electrons. Over the past few decades, a perfluorosulfonic acid polymer such as Nafion® is the most cited PEM material due to its high proton conductivity and good chemical and mechanical stabilities (1,2). But high gas permeability, low operating temperature (<80°C), and high cost of the Nafion® membrane have led to rigorous research efforts in the development of an alternative aromatic semifluorinated sulfonated polymer, such as poly(arylene ether)s (3,4,5), poly(arylene ether sulfone)s (6,7,8,9,10,11,12), poly(aryl ether ether nitrile)s (13), poly(arylene sulfide sulfone)s (14,15), poly(arylene ether ketone)s (16,17,18), poly(aryl ether ether ketone)s (19,20), poly(aryl ether ether ketone ketone)s (21,22), polyimides (23,24), polytriazoles (25,26), and polybenzimidazoles (27,28).

To improve some of the inadequate properties of PEMs, an efficient and convenient approach has been considered by forming hybrid materials such as organic–inorganic nanocomposites. Usually, sulfonated polymers/copolymers filled with inorganic nanofillers show lower ion exchange capacity (IEC) values and subsequently reduce the proton conductivity of the PEMs (29,30,31). The hydrophilic fillers, such as SiO2, TiO2, and ZrO2, are generally added to improve the water retention ability and enhance the proton conductivity of the PEMs (32,33,34). It envisioned that the introduction of these hygroscopic inorganic nanoparticles into the polymer matrix improves the hydration of the membrane and reduces the evaporation of water. Marani et al. synthesized titania nanosheets (TNS) and investigated the effects of TNS as an additive in sulfonated (poly ether ether ketone) (SPEEK) nanocomposites. They reported that the composite membrane with 1.67 wt% TNS showed a maximum conductivity value at 100% relative humidity (RH), which is much higher than that of a pure SPEEK membrane under the same conditions. In contrast, the degree of volume swelling was reduced by 80% to that of the reference SPEEK membrane (35).

In recent times, sulfonated nanofillers are being used as an additive to compensate for the reduction in IECW values with the addition of inorganic fillers to PEMs. Rhee et al. prepared sulfonated montmorillonite (SO3H-MMT) and developed a composite electrolyte membrane with Nafion® 115. The proton conductivity of the composite membrane was slightly lower than that of pristine Nafion® 115 (36). Bebin et al. synthesized Nafion® 115/sulfonic acid groups containing Laponite nanocomposite membranes and reported higher proton conductivity and water retention ability than pristine Nafion® 115 (37). Sua et al. reported a similar result toward proton conductivity when sulfonated silica nanoparticle (s-SNP) was added to the poly(phthalazinone ether ketone) (SPPEK) membrane (38). A 3.6-fold increase in proton conductivity than pristine SPPEK membrane was reported when 7.5 phr of silica-SO3H was added. Nevertheless, the conductivity value of the composite membrane was still lower than that of Nafion® 117. Later on, Xu et al. synthesized a series of new Nafion-based composite membranes via an in situ sol–gel reaction of 3-(trihydroxysilyl)propane-1-sulfonic acid (THSPSA) through the solution casting technique. More significantly, the composite membranes exhibit an increased water uptake and two times higher proton conductivity, compared to unmodified Nafion at high temperatures and low humidity (39). All these studies indicated that the composite approach might be an alternate way to enhance the PEM properties (40,41).

The main objective of the present research work was to exploit the effect of incorporated inorganic compounds in the presence of –COOH functionalized sulfonated polymers and fabricated highly proton-conducting hybrid membranes. Accordingly, a –COOH functionalized fluorinated sulfonated copolymer was prepared. Different amounts of THSPSA were added to this –COOH functionalized copolymer to make hybrid membranes by in situ sol–gel reactions of THSPSA during the film casting step (shown in Figure 1). The oxide precursor THSPSA was chosen due to its multifunctional structure. In contrast, the –COOH functionalized fluorinated sulfonated copolymers were selected for better interaction between the inorganic phases (–OH groups) and the ionic domains of the polymer matrix through –COOH groups. Besides, the sulfonic acid groups of the sulfonated poly(arylene ether sulfone) chains and the resulting sulfopropylated polysilsesquioxane (SiOPS) interacted with each other and formed a network structure, which was responsible for efficient proton-conducting channels. Thus, the utility of the newly prepared composite membranes (regarded as SPAES-SS-X, where X represents the theoretical amount in wt% of sulfonated silica in the hybrid membrane) has been demonstrated by investigating the morphology and selected PEM properties such as proton conductivity, water uptake, mechanical strength, swelling ratio, and thermal and chemical stability.

Figure 1 Architectural representation of the preparation of SPAES-SS-X composite membranes through the formation of an IPN structure.
Figure 1

Architectural representation of the preparation of SPAES-SS-X composite membranes through the formation of an IPN structure.

2 Experimental section

2.1 Materials

The bishydroxy monomer, 2-[bis-(4-hydroxyphenyl)-methyl]-benzoic acid (TP-COOH), and the –CF3 functionalized monomer, 4,4′-bis(4′-fluoro-3′-trifluoromethyl benzyl)biphenyl (QBF), were prepared, according to procedures reported in the literature (42,43). THSPSA (35 wt% in water) and 3,3′-disodiumsulfonyl-4,4′-dichlorodiphenylsulfone (SDCDPS, >98%) were purchased from Fluorochem and Chemos GmbH, respectively, and were used as received. N-Methyl-2-pyrrolidone (NMP; E. Merck) was purified by stirring with NaOH and distilled (twice from phosphorus pentoxide [P2O5]) under vacuum before use. Toluene (Merck) was refluxed over filamentous sodium to remove water and was freshly distilled prior to use. Potassium carbonate was dried under vacuum at 120°C for 24 h before use. Other chemicals were used as received without any further purification.

2.2 Synthesis of the –COOH functionalized fluorinated sulfonated poly(arylene ether) copolymer

The –COOH functionalized fluorinated sulfonated poly(arylene ether) copolymer (TPQSH-COOH-50) was synthesized via standard K2CO3-mediated polycondensation of a bisphenol monomer (TP-COOH) with two bishalide monomers (QBF and SDCDPS) in the ratio of 50:50 where the NMP and toluene were used as solvent and cosolvent, respectively. The detailed synthetic procedure with characterization of TPQSH-COOH-50 has been demonstrated in an earlier publication (44). The copolymer structure used in this investigation is shown in Figure 2.

Figure 2 Chemical structure of the pristine copolymer TPQSH-COOH-50 (44).
Figure 2

Chemical structure of the pristine copolymer TPQSH-COOH-50 (44).

2.3 Composite membrane preparation

The composite membrane is prepared using the following procedure. A measured amount of TPQSH-COOH-50 copolymer was dissolved in 15 mL of DMAc and filtered to obtain a transparent polymer solution. THSPSA (2.5–10 wt%) was added dropwise under vigorous stirring to this solution. The resulting mixture was sonicated for 30 min using a water bath sonicator. The homogeneous mixture was transferred into a round-bottomed flask for partial esterification reaction (between the –COOH group of the polymer and the –OH group of the silica filler), which was carried out at 80°C for 12 h in DMAc with 5 mg of p-toluenesulfonic acid per gram of polymer for catalysis.

The resulting solution with different amounts of THSPSA was cast onto a glass petri dish and heated at 80°C for 12 h to complete the esterification process as well as to form a Si–O–Si network. Then the temperature was slowly increased to 100°C, 120°C, 140°C, and 160°C, and the membranes were kept for 2 h in each temperature to remove any residual solvent. After cooling down to room temperature, the composite membranes were peeled off from the glass plate and acidified by immersing in the 1.5 M H2SO4 solution at room temperature for 24 h, followed by washing with deionized (DI) water several times until pH 7. All the hybrid membranes were dried under vacuum overnight at 120°C before use. The typical thickness of the membranes was in the range of 45–50 µm.

3 Characterization

Infrared spectra of the composite membranes were recorded on a NEXUS 870 FTIR (Thermo Nicolet) spectrophotometer under a humidity-free atmosphere at room temperature. The DSC measurements were performed under nitrogen atmosphere using TA Instruments DSC Q20 at a heating rate of 20 K/min and the glass transition temperatures (Tg) determined by taking the middle point of the step transition in the second heating run. Low-temperature dynamic scanning calorimetry (DSC) measurements were performed to determine the state of water (the frozen water content and the nonfrozen water content) within the polymer membranes using a NETZSCH DSC 200PC instrument at a heating rate of 5 K/min under nitrogen atmosphere. Initially, 5–10 mg swollen membranes were sealed hermetically in a sample pan and cooled to −40°C. Then after reaching the equilibrium, the sample was heated to 40°C. The single melting enthalpy peak (enthalpy of fusion) that arose near 0°C was considered in order to determine the amount of free water present in the membrane during the course of the experiment (an average value was taken for calculation). Thermogravimetric analysis (TGA) was performed using a TA Instruments TGAQ50 thermal analyzer under synthetic air (N2:O2 = 80:20) at a heating rate of 10 K/min. The mechanical properties of the acid form of the hybrid membranes (10 mm × 25 mm) were measured using a TINIUS OLSEN tensile testing machine (Model No. H5KS). The measurements were performed at a crosshead speed of 5 mm/min of the specimen length. All the membranes were predried at 80°C for 4 h before testing and the measurements were performed at 30°C and 60% RH. The IECW, water uptake, swelling ratio, and oxidative stability of the membranes were determined, according to the reported protocol (11,14). The surface images of the composite membranes were taken using a field emission scanning electron microscope manufactured by Carl Zeiss, Model Supra 40. The instrument was operated at an accelerated voltage of 5–20 kV equipped with an energy-dispersive spectrometer. The atomic force microscopy (AFM) analysis was carried out using AFM 5500 (Agilent Technology) in tapping mode. Transmission electron microscopy (TEM) images of the ultra-microtome membranes were taken by using an FEI-TECNAI G2 20 S-TWIN TEM instrument at an operating voltage of 80 kV. The acid form of the membranes was stained with Pb2+ ions using 0.5 M lead acetate solution for overnight to enhance the contrast between phases. The membranes were then washed with DI water several times to remove the surface salt solution and dried at room temperature for 24 h. The stained samples were sectioned by ultra-microtome under cryogenic condition and placed on carbon-coated copper grids before TEM viewing. In plane, proton conductivity (σ, mS/cm) of all the composite membranes was measured at different temperatures by four-probe AC impedance spectroscopy using an electrode system connected to a Gamry instrument reference 3000 potentiostat/galvanostat/ZAR over the frequency range of 100 Hz–1 MHz under fully hydrated condition (11,12). The reproducibility in the measurements of proton conductivity was within 2–3%, and the maximum value obtained is reported.

4 Results and discussion

The formation of sulfopropylated polysilsesquioxane in all the composite membranes was confirmed by the FT-IR spectra, as shown in Figure 3. The characteristic symmetric stretching and asymmetric stretching bands at 809 and 1,018 cm−1, respectively, suggest the formation of the Si–O–Si bond in the polymer composite (45). Additionally, the formation of polysilsesquioxane was also confirmed by the disappearance of Si–OH bands around 930–940 cm−1. The aromatic C═C stretching band appeared at 1,584 cm−1, whereas the C–F stretching band was found at 1,240–1,127 cm−1 (46). The characteristic symmetric stretching band at 1,051 cm−1 corresponds to the ether linkage of the polymers. The symmetric stretching of the aromatic S–O bond of the –SO3H group overlaps with the asymmetric stretching of the Si–O–Si bond, but a strong absorption band was found at 1,095 cm−1 which corresponds to the asymmetric stretches of the aromatic S–O bond of the –SO3H group. Moreover, the absorption band of S–O stretching from the –SO3H group of sulfopropylated polysilsesquioxane at 1,125 cm−1 was superimposed with the C–F stretching band of the pristine copolymer. All these results suggest that the sulfopropylated polysilsesquioxane was successfully developed into the polymer matrix during the sol–gel process.

Figure 3 FT-IR spectra of SPAES-SS-X hybrid membranes.
Figure 3

FT-IR spectra of SPAES-SS-X hybrid membranes.

The hybrid membranes do not show any noticeable signature for glass transitions in the DSC analysis (temperature range of 100–350°C) attributed to the locking of the segmental motion caused by the ionic interactions of the hydrophilic sulfonic groups of both SiOPS and the sulfonated poly(arylene ether sulfone) chains (12,44,46). All hybrid membranes displayed two-step degradation profiles in TGA (Figure 4) under air. The first weight loss (280–320°C) was assigned to the degradation of sulfonic acid groups. The 10% weight loss temperature of hybrid membranes was around 320°C and was lower than that of the pristine membrane (44). However, no significant change in 10% weight loss temperature of the hybrid membranes with the loading percentage of sulfonated SiOPS nanoparticles was observed. The weight loss range from 495 to 508°C was attributed to the decomposition of the polymer backbone (Table 1).

Figure 4 TGA curves of the SPAES-SS-X composite membranes in air.
Figure 4

TGA curves of the SPAES-SS-X composite membranes in air.

Table 1

Thermal, mechanical, and oxidative stability of SPAES-SS-X membranes as well as selected reference polymers

PolymerTga (°C)TdbTSc (MPa)YMd (GPa)EBe (%)Oxidative stability τf (h)Ref.
τ1τ2
TPQSH-COOH-50510g52g1.68g8g21.2g>24g(44)
SPAES-SS-2.5508471.54622.5>24This study
SPAES-SS-5.0505391.41318.0>24This study
SPAES-SS-7.5501361.391015.5>24This study
SPAES-SS-10495271.6427.5>24This study
Nafion® 11712943521.9h0.16h288h>24h(12)
  1. a

    Glass transition temperature determined by the DSC.

  2. b

    Backbone degradation temperature.

  3. c

    Tensile strength, at a strain rate: 5%/min, 65 ± 2% RH, and 30°C.

  4. d

    Young’s modulus.

  5. e

    Elongation at break.

  6. f

    τ1 and τ2 refer to the initial breaking time and the time to complete dissolution in Fenton’s reagent (2 ppm FeSO4 in 3% H2O2) at 80°C, respectively.

  7. g

    Data from (44).

  8. h

    Data from (12).

Mechanical stability is critical of the electrolyte membranes for effective utilization in a fuel cell as PEMs. Table 1 summarizes the mechanical properties of the composite membranes and the pristine membrane and Nafion® 117. Figure 5 shows that all the composite membranes exhibit a lower tensile strength value than the –COOH functionalized pristine copolymer membrane. The existence of additional –SO3H groups in the polymer matrix and the higher structural rigidity of the inorganic nanofillers make them mechanically weak (33). In general, all these hybrid membranes showed higher tensile strength and Young’s modulus with a lower elongation at break than Nafion® 117. In the dry state, the tensile strength values of SPAES-SS-X hybrid membranes decreased steadily with the increase in filler loading, but no absolute trend was observed in Young’s modulus and elongation at break values. The membranes with 7.5 wt% (SPAES-SS-7.5) filler loading showed an optimal set of mechanical properties with elongation at break as high and sufficient for fabricating a membrane electrode assembly (47,48).

Figure 5 Stress–strain plot of the hybrid membranes.
Figure 5

Stress–strain plot of the hybrid membranes.

Oxidative stability of the hybrid membranes was checked by immersing the samples in Fenton’s reagent (3 wt% H2O2, 2 ppm FeSO4) at 80°C (Table 1). The hybrid membrane with 2.5 wt% filler exhibited the highest oxidative stability, with τ1 = 22.5 h, whereas the least oxidative stability was observed for 10% filler loaded membrane (τ1 = 7.5 h). Generally, compared to the pristine membrane, the oxidative stability of the hybrid membranes increases with increasing filler loading. Here, the presence of additional –SO3H groups in the nanofiller reduces the oxidative stability of the membranes. Hydrophilic domains are prone to an oxidative attack by the free radical species (HO˙ and HOO˙). Thus, increasing the percentage of s-SNPs (THSPSA) increases the number of hydrophilic domains, and consequently, the oxidative stability of the hybrid membranes with higher filler loading decreases. However, due to the formation of the crosslinked and interpenetrating network structure, the hybrid membrane, even with 10 wt% filler, showed excellent oxidative stability, compared to many other pristine membrane materials (24,44,46). On the other hand, the –CF3 groups in QBF and the wholly aromatic nature of the copolymers are mainly responsible for enhancing the oxidative stability and making them resistant toward attack by the oxidative species (49).

The hydrolytic stability of the hybrid membranes was investigated at 80°C, keeping them for 24 h in water. The membranes were substantially stable in wet conditions, as no apparent changes were observed in their appearance and weight even after 24 h. Although the SiOPS alone dissolves immediately in hot water, the hybrid membranes exhibit a high hydrolytic stability. This stability arises from a strong interaction (possibly hydrogen bonding) between the sulfonic acid groups of SiOPS and the sulfonated poly(arylene ether sulfone) chains (50).

IECW usually signifies the number of exchangeable protons present in the polymer electrolytes and was measured by the classical titrimetric analysis with 0.1 M NaOH solution. The IECW values of SPAES-SS-X copolymer membranes and Nafion® 117 are given in Table 2. The titrimetric IECW value of the bare TPQSH-COOH-50 was found to be 1.32 mequiv./g, and with an increase of THSPSA loading, the IECW was found to increase rapidly up to 1.48 mequiv./g, but a further rise in THSPSA content from 7.5 to 10 wt% did not lead to a noticeable increase in IECW. It is attributed to the trapping of H+ ions (present in the inner sulfonic acid groups) by large SiOPS nanoclusters at high precursor concentrations (39). This result corroborates with the water uptake of the hybrid membranes. The water uptake values of the hybrid membranes were significantly higher than that of the pristine TPQSH-COOH-50 membrane, which is attributed to the hygroscopic nature of the SiOPS nanofiller (39). With increase in THSPSA loading up to 7.5 wt%, the water uptake values gradually increased from 47% to 61% (at 80°C). Beyond 7.5% loading of THSPSA, the rise in water uptake of the SPAES-SS-10 hybrid membrane increased marginally (64%), as shown in Figure 6a.

Table 2

IECW, hydration number, water uptake, and proton conductivity of pristine copolymer, SPAES-SS-X composite membranes, and Nafion® 117

PolymerIECW (meq/g)λb [H2O/SO3]WUc (wt%)Swelling ratio (%)σd (mS/cm)Ea (kJ/mol)
In plane
Theo.aTitr.30°C80°C30°C80°C80°C30°C80°C
TPQSH-COOH-501.381.3224321630559.6
SPAES-SS-2.51.381.36171943473389315.6
SPAES-SS-5.01.381.411921475254011418.8
SPAES-SS-7.51.381.482325576195313016.0
SPAES-SS-101.381.5024265964105413816.7
Nafion® 1170.910.909316510.2
  1. a

    IECW,Theo. = (1,000/MWrepeat unit) × DSTheo × 2, where DSTheo is calculated theoretically from a monomer feed ratio.

  2. b

    λ = WUW (%)/(100 × IECW,Theo. × MW,H2O), where MW,H2O = 18 g/mol.

  3. c

    WU(wt%) = [(WwetWdry)/Wdry] × 100.

  4. d

    Proton conductivity was measured under fully hydrated condition (in water).

Figure 6 (a) Variation in the water uptake (wt%) with % of THSPSA at different temperatures of SPAES-SS-X composite membranes. (b) DSC melting curves of water-swollen membranes of TPQSH-COOH-50 and SPAES-SS-2.5.
Figure 6

(a) Variation in the water uptake (wt%) with % of THSPSA at different temperatures of SPAES-SS-X composite membranes. (b) DSC melting curves of water-swollen membranes of TPQSH-COOH-50 and SPAES-SS-2.5.

A low-temperature DSC experiment was performed to investigate the state of water inside the hydrated membranes by measuring the latent heat of fusion (ΔHf) for water (44). Generally, a higher ΔHf value indicated a high concentration of free and weakly bound waters that are responsible for the melting enthalpy peaks near 0°C in the DSC curves. It is consistent with the literature reports claiming that the loosely bounded interfacial water molecules are mainly responsible for the interconnected channels between ionic domains and play a crucial role in proton conduction of Nafion® (2,51). In contrast with Nafion® 117 (39), the SPAES-SS-2.5 membrane showed a sharp melting peak near 0°C, which is attributed to a more significant number of free and weekly bound water molecules in it (Figure 6b). However, the same behavior was not observed with the hybrid membranes when the filler loading increased to 5 wt% or higher. It could be observed that due to the presence of highly dense sulfonic acid groups, the hybrid membranes with more filler loading contain a higher number of strongly bonded water molecules (probably through hydrogen bonding) and do not contribute to the melting endotherm in the DSC curves. The hydration number (λ, the average number of water molecules per sulfonic acid group) of the hybrid membranes was calculated to probe the state of water inside the hydrated membranes. From Table 2, it is seen that at 80°C, with an increase in THSPSA, the hydration number (λ) sharply increased from 19 to 25 for SPAES-SS-7.5 and then slowly to 26 for SPAES-SS-10.

Analogous to the water uptake, the swelling ratios (in-plane) of the SPAES-SS-X hybrid membranes gradually enhanced with increasing sulfonated silica nanofiller loading and temperatures. However, compared to the pristine copolymer membrane, the hybrid membranes showed a much lower-dimensional swelling attributed to the formation of a cage-like structure of the silica network. They suppressed the free movement of water molecules through the bulk of the polymer (52). At 80°C, all the hybrid membranes possess a much better dimensional stability than Nafion® 117, resulting from the bulky TP-COOH groups and hydrophobic –CF3 groups along with a rigid quadriphenyl moiety (18).

Scanning electron microscope (SEM) techniques have extensively studied the surface morphology of the hybrid membranes to examine the homogeneity of the silica sulfonic acid nanoparticles (THSPSA) in the polymer matrix. The SEM image (Figure 7) reveals that the s-SNPs were uniformly dispersed on the polymer surface, indicating homogeneity. The compatibility between the organic and inorganic components improved, probably due to the existence of two types of covalent bonds (the sulfonated silica particles formed the covalent bonds with a –COOH functionalized polymer and the bond was formed after the formation of SiOPS) present in the hybrid membranes. The density of the –SO3H groups on the membrane surface gradually increases with a gradual increase in silica loading (2.5–10 wt%). In addition to that, the surface morphology of the composite membranes was studied by AFM in tapping mode (Figure 8). There are two regions in the AFM image where the bright part corresponds to the hard hydrophobic domains and the dark area corresponds to the soft hydrophilic domains. The interconnectivity of the ionic domains was vividly observed in the composite membrane, with 7.5 wt% of silica loading.

Figure 7 SEM images of SPAES-SS-X composite membranes.
Figure 7

SEM images of SPAES-SS-X composite membranes.

Figure 8 2D tapping mode AFM images SPAES-SS-X composite membranes.
Figure 8

2D tapping mode AFM images SPAES-SS-X composite membranes.

The bulk morphology of the membranes played an important role in elucidating the proton conductivity and in being investigated by the TEM. The Pb2+-stained sample membranes showed (Figure 9) an excellent micro-phase separated morphology wherein the dark spherical regions correspond to the hydrophilic ionic domains. The brighter areas represent the hydrophobic nonionic domains of the polymer matrix. However, the connectivity of the ionic domains of the composite membranes was significantly better than that of the pristine SPAES membrane (44). The hybrid membranes with 2.5 wt% SiOPS consist of both small hydrophilic clusters (particle size ∼20–25 nm) and some interconnected medium-sized ionic clusters (particle size ∼40–45 nm). The connectivity of the ionic domains was more pronounced on increasing the percentage of hydrophilic silica nanoparticles. It is believed that the ionic domains resulted from the aggregation of sulfonic acid groups from the polymer electrolyte and THSPSA/SiOPS, and the maximum interconnectivity was found in the SPAES-SS-10 composite membrane. The result suggests that the SiOPS aggregates distributed relatively more uniformly to form a well-dispersed nanocomposite membrane.

Figure 9 TEM micrographs of lead (Pb2+) stained SPAES-SS-X composite membranes.
Figure 9

TEM micrographs of lead (Pb2+) stained SPAES-SS-X composite membranes.

Proton conductivity is the prime property for a PEM material for its effective utilization in fuel cells. In plane, proton conductivity of all the composite membranes was measured using AC impedance spectroscopy within the temperature ranging from 30 to 80°C under fully hydrated condition. In comparison with the pristine TPQSH-COOH-50, the hybrid membranes exhibited an improved proton conductivity due to the presence of additional sulfonic acid groups in nanoparticles. Table 2 summarizes the proton conductivity of the membranes, and Figure 10a represents the change in proton conductivity of the composite membranes as a function of temperatures along with Nafion® 117.

Figure 10 (a) Proton conductivity of bare copolymer and composite membranes. (b) Variation in proton conductivity with % of THSPSA at different temperatures under fully hydrated conditions of SPAES-SS-X hybrid membranes.
Figure 10

(a) Proton conductivity of bare copolymer and composite membranes. (b) Variation in proton conductivity with % of THSPSA at different temperatures under fully hydrated conditions of SPAES-SS-X hybrid membranes.

The proton conductivity increases progressively with the loading of the THSPSA content, as shown in Figure 10b and, in particular, the conductivity of the hybrid membranes (mainly XX = 7.5 and 10) at low temperature was significantly improved, compared to the pristine membrane (TPQSH-COOH-50). Depending on the amount of filler loading, the proton conductivity of the hybrid membranes was exhibited in the ranges of 38–54 mS/cm and 93–138 mS/cm, respectively, at 30°C and 80°C (Table 2). The 10 wt% SiOPS loaded membrane showed more than two times higher proton conductivity (138 mS/cm) than the pristine TPQSH-COOH-50 membrane (55 mS/cm) at 80°C under the same test conditions (44). It is noted that the proton conductivity of this 10 wt% filler loaded membrane did not increase that sharply, compared to that of the 7.5 wt% SiOPS charged membrane. This feature was attributed to the trapping of inner –SO3H groups by forming a high precursor concentration of the large SiOPS nanoclusters in 10 wt% filler loaded membrane and making them inert for the interaction with water (50). The proton conductivity values are in close agreement with the water uptake values of the hybrid membranes. The proton transportation of hybrid membranes occurred by both surface and bulk transport paths. The protons hopped on the surface of the pore channels in the form of H3O+ in the surface transport pathway, whereas, in bulk transport mechanisms, the protons jumped in the bulk of the pore channels in the form of H5O2+/H9O4+ (53,54,55). The density of the functional groups and the distance between two neighboring functional groups control the rate of surface transport pathway, whereas the rate of bulk transport pathway was controlled by the RH of the membrane surroundings and the cross-sectional area of the pore channels. The incorporation of functionalized silica (THSPSA) into the polymer matrix reduced the distance between two neighboring proton-conducting moieties, compared to the pristine TPQSH-COOH-50 membrane. It facilitated the surface transportation effectively through the hybrid membranes. For a better understanding, the temperature-dependent proton conductivity of the membranes is plotted in Figure 11.

The activation energies of composite membranes were calculated from the slope of different lines presented in Figure 11 using the Arrhenius equation and the values were found to be in the range of 15.6–18.8 kJ/mol, close to that of Nafion® 117. The activation energies of the composite membranes are higher than that of the pristine TPQSH-COOH-50. The hydrogen bond present in the crosslinked composite membranes needs to be broken at the high temperature to facilitate proton transport, resulting in relatively high activation energies (12). These results were in concurrence with the morphological data. They suggested that the incorporation of SiOPS evokes a significant improvement in the polymer electrolyte membranes’ water transport behavior by forming an interconnected ionic network that causes high proton conductivity.

Figure 11 Temperature-dependent proton conductivity (σ) of the hybrid membranes.
Figure 11

Temperature-dependent proton conductivity (σ) of the hybrid membranes.

5 Conclusions

A series of fluorinated sulfonated hybrid membranes were fabricated using a practical approach to improve the proton conductivity by a simple incorporation of an acid-functionalized THSPSA nanofiller. The prepared hybrid membranes showed excellent mechanical stability. The SEM studies suggest that the sulfonated SiOPS nanoparticles were dispersed uniformly on the polymer matrix in all the hybrid membranes with high homogeneity, probably due to the existence of two types of covalent bonds. However, the TEM micrographs of the composite membranes showed a clear phase-separated morphology with a significantly better connectivity of the ionic domains than the pristine SPAES membrane. The pendant sulfonic acids on the polysilsesquioxane fillers help to increase the interconnectivity of the proton transport channels by reducing the distance between two neighboring proton-conducting moieties, compared to the pristine TPQSH-COOH-50 membrane and facilitated the proton transportation effectively through the composite membranes. The SPAES-SS-10 membrane showed a maximum proton conductivity of 138 mS/cm at 80°C. This work demonstrates that the prepared nanocomposite conductive membrane materials are expected to find their applications in PEM fuel cells.

Acknowledgment

RM thanks the CSIR, New Delhi, for providing the fellowship to carry out this research.

  1. Conflict of interest: One of the authors (Susanta Banerjee) is a member of the Editorial Advisory Board of e-Polymers.

References

(1) Carrette L, Friedrich KA, Stimming U. Fuel cells-fundamentals and applications. Fuel cells. 2001;1:5–39.10.1002/1615-6854(200105)1:1<5::AID-FUCE5>3.0.CO;2-GSearch in Google Scholar

(2) Mauritz KA, Moore RB. State of understanding of Nafion. Chem Rev. 2004;104:4535–86.10.1021/cr0207123Search in Google Scholar

(3) Miyatake K, Oyaizu K, Tsuchida E, Hay AS. Synthesis and properties of novel sulfonated arylene ether/fluorinated alkane copolymers. Macromolecules. 2001;34:2065–71.10.1021/ma0019959Search in Google Scholar

(4) Harrison WL, Wang F, Mecham JB, Bhanu VA, Hill M, Kim YS, et al. Influence of the bisphenol structure on the direct synthesis of sulfonated poly(arylene ether) copolymers. I. J Polym Sci Part A Polym Chem. 2003;41:2264–76.10.1002/pola.10755Search in Google Scholar

(5) Kim DS, Robertson GP, Kim YS, Guiver MD. Copoly(arylene ether)s containing pendant sulfonic acid groups as proton exchange membranes. Macromolecules. 2009;42:957–63.10.1021/ma802192ySearch in Google Scholar

(6) Wang F, Hickner M, Ji Q, Harrison W, Mecham J, Zawodzinski TA, et al. Synthesis of highly sulfonated poly(arylene ether sulfone) random (statistical) copolymers via direct polymerization. Macromol Symp. 2001;175:387–96.10.1002/1521-3900(200110)175:1<387::AID-MASY387>3.0.CO;2-1Search in Google Scholar

(7) Wang F, Hickner M, Kim YS, Zawodzinski TA, McGrath JE. Direct polymerization of sulfonated poly(arylene ether sulfone) random (statistical) copolymers: candidates for new proton exchange membranes. J Membr Sci. 2002;197:231–42.10.1016/S0376-7388(01)00620-2Search in Google Scholar

(8) Kim YS, Einsla BR, Sankir M, Harrison W, Pivovar BS. Structure–property–performance relationships of sulfonated poly(arylene ether sulfone)s as a polymer electrolyte for fuel cell applications. Polymer. 2006;47:4026–35.10.1016/j.polymer.2006.02.032Search in Google Scholar

(9) Zhang N, Li J, Wang X, Xia Z, Liu H. Preparation and properties of bisphenol A-based sulfonated poly(arylene ether sulfone) proton exchange membranes for direct methanol fuel cell. J Appl Polym Sci. 2009;114:304–12.10.1002/app.30482Search in Google Scholar

(10) Li N, Shin DW, Hwang DS, Lee YM, Guiver MD. Polymer electrolyte membranes derived from new sulfone monomers with pendent sulfonic acid groups. Macromolecules. 2010;43:9810–20.10.1021/ma102107aSearch in Google Scholar

(11) Mohanty AK, Mistri EA, Ghosh A, Banerjee S. Synthesis and characterization of novel fluorinated poly(arylene ether sulfone)s containing pendant sulfonic acid groups for proton exchange membrane materials. J Membr Sci. 2012;145:409–10.10.1016/j.memsci.2012.03.048Search in Google Scholar

(12) Mukherjee R, Banerjee S, Komber H, Voit B. Highly proton conducting fluorinated sulfonated poly(arylene ether sulfone) copolymers with side chain grafting. RSC Adv. 2014;4:46723–36.10.1039/C4RA07291JSearch in Google Scholar

(13) Gao Y, Robertson GP, Guiver MD, Mikhailenko SD, Li X, Kaliaguine S. Synthesis of copoly(aryl ether ether nitrile)s containing sulfonic acid groups for PEM application. Macromolecules. 2005;38:3237–45.10.1021/ma047572eSearch in Google Scholar

(14) Bai Z, Dang TD. Direct synthesis of fully sulfonated polyarylenethioether sulfones as proton-conducting polymers for fuel cells. Macromol Rapid Commun. 2006;27:1271–7.10.1002/marc.200600275Search in Google Scholar

(15) Shin DW, Lee SY, Kang NR, Lee KH, Guiver MD, Lee YM. Durable sulfonated poly(arylene sulfide sulfone nitrile)s containing naphthalene units for direct methanol fuel cells (DMFCs). Macromolecules. 2013;46:3452–60.10.1021/ma400154tSearch in Google Scholar

(16) Xing P, Robertson GP, Guiver MD, Mikhailenko SD, Kaliaguine S. Sulfonated poly(aryl ether ketone)s containing the hexafluoroisopropylidene diphenyl moiety prepared by direct copolymerization, as proton exchange membranes for fuel cell application. Macromolecules. 2004;37:7960–7.10.1021/ma0494941Search in Google Scholar

(17) Shang X, Tian S, Kong L, Meng Y. Synthesis and characterization of sulfonated fluorene-containing poly(arylene ether ketone) for proton exchange membrane. J Membr Sci. 2005;266:94–101.10.1016/j.memsci.2005.05.014Search in Google Scholar

(18) Xing P, Robertson GP, Guiver MD, Mikhailenko SD, Kaliaguine S. Synthesis and characterization of poly(aryl ether ketone) copolymers containing (hexafluoroisopropylidene)-diphenol moiety as proton exchange membrane materials. Polymer. 2005;46:3257–63.10.1016/j.polymer.2005.03.003Search in Google Scholar

(19) Zhong S, Liu C, Dou Z, Li X, Zhao C, Fu T, et al. Synthesis and properties of sulfonated poly(ether ether ketone ketone) containing tert-butyl groups as proton exchange membrane materials. J Membr Sci. 2006;285:404–11.10.1016/j.memsci.2006.09.016Search in Google Scholar

(20) Shen L, Xiao G, Sun G. Sulfonated poly(arylene thioether ketone ketone sulfone)s for proton exchange membranes with high oxidative stability. e-Polymers. 2005;5:321–30.10.1515/epoly.2005.5.1.321Search in Google Scholar

(21) Liu B, Kim DS, Murphy J, Robertson GP, Guiver MD, Mikhailenko S, et al. Fluorenyl-containing sulfonated poly(aryl ether ether ketone ketone)s (SPFEEKK) for fuel cell applications. J Membr Sci. 2006;280:54–64.10.1016/j.memsci.2006.01.004Search in Google Scholar

(22) Zhang G, Fu T, Shao K, Li X, Zhao C, Na H, et al. Novel sulfonated poly(ether ether ketone ketone)s for direct methanol fuel cells usage: synthesis, water uptake, methanol diffusion coefficient and proton conductivity. J Power Sources. 2009;189:875–81.10.1016/j.jpowsour.2008.12.155Search in Google Scholar

(23) Mistri EA, Mohanty AK, Banerjee S, Komber H, Voit B. Naphthalene dianhydride based semifluorinated sulfonated copoly(ether imide)s: synthesis, characterization and proton exchange properties. J Membr Sci. 2013;441:168–77.10.1016/j.memsci.2013.03.015Search in Google Scholar

(24) Mandal AK, Bisoi S, Banerjee S, Komber H, Voit B. Sulfonated copolyimides containing trifluoromethyl and phosphine oxide moieties: synergistic effect towards proton exchange membrane properties. Eur Polym J. 2017;95:581–95.10.1016/j.eurpolymj.2017.08.050Search in Google Scholar

(25) Singh A, Kumar AG, Bisoi S, Saha S, Banerjee S. Hexafluoroisopropylidene based sulfonated new copolytriazoles: investigation of proton exchange membrane properties. e-Polymers. 2017;17:107–18.10.1515/epoly-2016-0285Search in Google Scholar

(26) Saha S, Mukherjee R, Singh A, Banerjee S. Synthesis, characterization and investigation of proton exchange properties of sulfonated polytriazoles from a new semifluorinated diazide monomer. Polym Eng Sci. 2017;57:312–23.10.1002/pen.24415Search in Google Scholar

(27) Kulkarni MP, Peckham JT, Thomas OD, Holdcroft S. Synthesis of highly sulfonated polybenzimidazoles by direct copolymerization and grafting. J Polym Sci Part A Polym Chem. 2013;51:3654–66.10.1002/pola.26764Search in Google Scholar

(28) Singha S, Jana T, Modestra JA, Kumar AN, Mohan SV. Highly efficient sulfonated polybenzimidazole as a proton exchange membrane for microbial fuel cells. J Power Sources. 2016;317:143–52.10.1016/j.jpowsour.2016.03.103Search in Google Scholar

(29) Shao ZG, Joghee P, Hsing IM. Preparation and characterization of hybrid Nafion–silica membrane doped with phosphotungstic acid for high temperature operation of proton exchange membrane fuel cells. J Membr Sci. 2004;229:43–51.10.1016/j.memsci.2003.09.014Search in Google Scholar

(30) Zhang J, Fei G, Liang Y, Zhang Y, Zhao J. Influence of silica content in sulfonated polysulfone/phosphotungstic acid hybrid membranes on properties for fuel cell application. e-Polymers. 2010;10:1–10. 10.1515/epoly.2010.10.1.1135Search in Google Scholar

(31) Ghosh A, Banerjee S. Sulfonated fluorinated-aromatic polymers as proton exchange membranes. e-Polymers. 2014;14:227–57.10.1515/epoly-2014-0049Search in Google Scholar

(32) Miyake N, Wainright JS, Savinell RF. Evaluation of a sol–gel derived Nafion/silica hybrid membrane for proton electrolyte membrane fuel cell applications: I. Proton conductivity and water content. J Electrochem Soc. 2001;148:A898–A904.10.1149/1.1383071Search in Google Scholar

(33) Adjemian KT, Dominey R, Krishnan L, Ota H, Majsztrik P, Zhang T, et al. Baocarsly, function and characterization of metal oxide-Nafion composite membranes for elevated-temperature H2/O2 PEM fuel cells. Chem Mater. 2006;18:2238–48.10.1021/cm051781bSearch in Google Scholar

(34) Pereira F, Valle K, Belleville P, Morin A, Lambert Sanchez C. Advanced mesostructured hybrid silica-Nafion membranes for high-performance PEM fuel cell. Chem Mater. 2008;20:1710–8.10.1021/cm070929jSearch in Google Scholar

(35) Marani D, D’Epifanio A, Traversa E, Miyayama M, Licoccia S. Titania nanosheets (TNS)/sulfonated poly ether ether ketone (SPEEK) nanocomposite proton exchange membranes for fuel cells. Chem Mater. 2010;22:1126–33.10.1021/cm902405pSearch in Google Scholar

(36) Rhee CH, Kim HK, Chang H, Lee JS. Nafion/sulfonated montmorillonite composite:  a new concept electrolyte membrane for direct methanol fuel cells. Chem Mater. 2005;17:1691–7.10.1021/cm048058qSearch in Google Scholar

(37) Bebin P, Caravanier M, Galiano H. Nafion®/clay-SO3H membrane for proton exchange membrane fuel cell application. J Membr Sci. 2006;278:35–42.10.1016/j.memsci.2005.10.042Search in Google Scholar

(38) Sua YH, Liu YL, Sun YM, Lai JY, Wang DM, Gao Y, et al. Proton exchange membranes modified with sulfonated silica nanoparticles for direct methanol fuel cells. J Membr Sci. 2007;296:21–8.10.1016/j.memsci.2007.03.007Search in Google Scholar

(39) Xu K, Chanthad C, Gadinski MR, Hickner MA, Wang Q. Acid-functionalized polysilsesquioxane-Nafion composite membranes with high proton conductivity and enhanced selectivity. ACS Appl Mater Interfaces. 2009;1:2573–9.10.1021/am900498uSearch in Google Scholar PubMed

(40) Singha S, Jana T. Structure and properties of polybenzimidazole/silica nanocomposite electrolyte membrane: influence of organic/inorganic interface. ACS Appl Mater Interfaces. 2014;6:21286–96.10.1021/am506260jSearch in Google Scholar PubMed

(41) Lin H, Zhao C, Cui Z, Ma W, Fu T, Na H, et al. Novel sulfonated poly(arylene ether ketone) copolymers bearing carboxylic or benzimidazole pendant groups for proton exchange membranes. J Power Sources. 2009;193:507–14.10.1016/j.jpowsour.2009.04.036Search in Google Scholar

(42) Kutcherlapati SR, Koyilapu R, Jana T. Poly(N-vinyl imidazole) grafted silica nanofillers: synthesis by RAFT polymerization and nanocomposites with polybenzimidazole. J Polym Sci Part A Polym Chem. 2018;56:365–75.10.1002/pola.28917Search in Google Scholar

(43) Banerjee S, Maier G. Novel high Tg high-strength poly(aryl ether)s. Chem Mater. 1999;11:2179–84.10.1021/cm9910153Search in Google Scholar

(44) Mukherjee R, Banerjee S, Komber H, Voit B. Carboxylic acid functionalized fluorinated sulfonated poly(arylene ether sulfone)copolymers with enhanced oxidative stability. J Membr Sci. 2016;510:497–509.10.1016/j.memsci.2016.03.028Search in Google Scholar

(45) Mauritz KA, Warren RM. Microstructural evolution of a silicon oxide phase in a perfluorosulfonic acid ionomer by an in situ sol–gel reaction. 1. Infrared spectroscopic studies. Macromolecules. 1989;22:1730–4.10.1021/ma00194a038Search in Google Scholar

(46) Mukherjee R, Mohanty AK, Banerjee S, Komber H, Voit B. Phthalimidine based fluorinated sulfonated poly(arylene ether sulfone)s copolymer proton exchange membranes. J Membr Sci. 2013;435:145–54.10.1016/j.memsci.2013.01.059Search in Google Scholar

(47) Hickner MA, Ghassemi H, Kim YS, Einsla BR, McGrath JE. Alternative polymer systems for proton exchange membranes (PEMs). Chem Rev. 2004;104:4587–612.10.1021/cr020711aSearch in Google Scholar PubMed

(48) Park SG, Chae KJ, Lee M. A Sulfonated poly(arylene ether sulfone)/polyimide nanofiber composite proton exchange membrane for microbial electrolysis cell application under the coexistence of diverse competitive cations and protons. J Membr Sci. 2017;540:165–73.10.1016/j.memsci.2017.06.048Search in Google Scholar

(49) Banerjee S. Handbook of specialty fluorinated polymers. 1st edn. New York: Elsevier; 2015.10.1016/B978-0-323-35792-0.00001-5Search in Google Scholar

(50) Miyatake K, Tombe T, Chikashige Y, Uchida H, Watanabe M. Enhanced proton conduction in polymer electrolyte membranes with acid-functionalized polysilsesquioxane. Angew Chem, Int Ed. 2007;46:6646–9.10.1002/anie.200701100Search in Google Scholar PubMed

(51) Schmidt KR, Chen Q. Parallel cylindrical water nanochannels in Nafion fuel-cell membranes. Nature Mater. 2008;7:75–83.10.1142/9789814317665_0033Search in Google Scholar

(52) So SY, Yoon YJ, Kim TH, Yoon K, Hong YT. Sulfonated poly(arylene ether sulfone)/functionalized silicate hybrid proton conductors for high-temperature proton exchange membrane fuel cells. J Membr Sci. 2011;381:204–10.10.1016/j.memsci.2011.07.024Search in Google Scholar

(53) Eikerling M, Kornyshev AA, Kuznetsov AM, Ulstrup J, Walbran S. Mechanisms of proton conductance in polymer electrolyte membranes. J Phys Chem B. 2011;105:3646–62.10.1021/jp003182sSearch in Google Scholar

(54) Spohr E, Commer P, Kornyshev AA. Enhancing proton mobility in polymer electrolyte membranes:  lessons from molecular dynamics simulations. J Phys Chem B. 2002;106:10560–9.10.1021/jp020209uSearch in Google Scholar

(55) Kornyshev AA, Kuznetsov AM, Spohr E, Ulstrup J. Kinetics of proton transport in water. J Phys Chem B. 2003;107:3351–66.10.1021/jp020857dSearch in Google Scholar

Received: 2020-06-30
Revised: 2020-07-21
Accepted: 2020-07-21
Published Online: 2020-08-13

© 2020 Rajdeep Mukherjee et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 28.3.2024 from https://www.degruyter.com/document/doi/10.1515/epoly-2020-0048/html
Scroll to top button