Next Article in Journal
Characteristic Tearing Energy and Fatigue Crack Propagation of Filled Natural Rubber
Next Article in Special Issue
Utilizing Biomass-Based Graphene Oxide–Polyaniline–Ag Electrodes in Microbial Fuel Cells to Boost Energy Generation and Heavy Metal Removal
Previous Article in Journal
Efficient Congo Red Removal Using Porous Cellulose/Gelatin/Sepiolite Gel Beads: Assembly, Characterization, and Adsorption Mechanism
Previous Article in Special Issue
Silica Containing Composite Anion Exchange Membranes by Sol–Gel Synthesis: A Short Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Anion Exchange Membranes with 1D, 2D and 3D Fillers: A Review

by
Riccardo Narducci
1,*,
Emanuela Sgreccia
1,
Philippe Knauth
2 and
Maria Luisa Di Vona
1
1
Department Industrial Engineering and International Laboratory “Ionomer Materials for Energy”, University of Rome Tor Vergata, I-00133 Rome, Italy
2
CNRS, Madirel (UMR 7246) and International Laboratory “Ionomer Materials for Energy”, Aix Marseille University, F-13013 Marseille, France
*
Author to whom correspondence should be addressed.
Polymers 2021, 13(22), 3887; https://doi.org/10.3390/polym13223887
Submission received: 22 October 2021 / Revised: 5 November 2021 / Accepted: 8 November 2021 / Published: 10 November 2021
(This article belongs to the Special Issue Polymeric Materials for Fuel Cell Applications)

Abstract

:
Hydroxide exchange membrane fuel cells (AEMFC) are clean energy conversion devices that are an attractive alternative to the more common proton exchange membrane fuel cells (PEMFCs), because they present, among others, the advantage of not using noble metals like platinum as catalysts for the oxygen reduction reaction. The interest in this technology has increased exponentially over the recent years. Unfortunately, the low durability of anion exchange membranes (AEM) in basic conditions limits their use on a large scale. We present in this review composite AEM with one-dimensional, two-dimensional and three-dimensional fillers, an approach commonly used to enhance the fuel cell performance and stability. The most important filler types, which are discussed in this review, are carbon and titanate nanotubes, graphene and graphene oxide, layered double hydroxides, silica and zirconia nanoparticles. The functionalization of the fillers is the most important key to successful property improvement. The recent progress of mechanical properties, ionic conductivity and FC performances of composite AEM is critically reviewed.

Graphical Abstract

1. Introduction

The anion exchange membranes (AEM) are the central element of many technologically relevant devices [1,2,3], first and foremost alkaline membrane fuel cells (FC) [4,5,6]. These fuel cells can significantly reduce the amount of noble metal catalysts for the oxygen reduction reaction (ORR) and may represent the future of FC development. However, although AEM may revolutionize future fuel cell technology, currently they lack several properties that are paramount for the viability of this technology, including long-term stability in the alkaline medium that impacts their durability in operating conditions [7,8,9,10]. One favoured way to improve the properties of ionomeric and more generally polymeric membranes is the development of composite materials, especially using inorganic fillers that are supposed to enhance the mechanical properties and the FC performance. The reduction in ionic conductivity consecutive to the addition of an inert second phase can be mitigated by fillers that present an intrinsic ionic conductivity or by space charge effects near the interfaces between ionomer and inorganic nanoparticles. Furthermore, the composite material may in certain cases show higher long-term stability in alkaline conditions by synergistic effects, due, e.g., to a physical cross-linking (such as acid–base interactions) or a partial crystallization of the polymer near the inorganic phase. Although the application of nanoparticle fillers is promising, their practical investigation started in earnest only a decade ago, as shown in the number of publications and citations on the topic “composite anion exchange membranes” (Figure 1). No comprehensive and critical review exists on this topic, although some related topics have been addressed in the last years [11,12,13,14,15].
This review intends thus to summarize the main findings reported in the literature since 2010, including AEM for use in fuel cells, but also in redox flow batteries and water purification applications. We chose a simple and coherent presentation of this wide field by subdividing the literature according to the dimensionality of the fillers. We start from 1D materials, essentially nanotubular solids, including titanate and carbon nanotubes (NTs). Titanate NT exhibit high chemical stability, high specific surface area and good performance in alkaline environments and their aspect ratio can increase the mechanical properties. Carbon NT are also characterized by extreme strength and flexibility so that an increase in mechanical properties is expected. Organic functionalization is the key to obtaining high ionic conductivity. The following chapter summarizes 2D materials with a large part on graphene and graphene oxide (GO). Graphene and GO are mechanically extremely resistant and at the same time present great flexibility. Composites, especially with functionalized graphene and/or reduced GO, are expected to have improved FC performance and conductivity thanks to a better filler distribution and sometimes with the help of ionic liquids. Layered double hydroxides (LDHs) can mitigate the loss of hydroxide ion conductivity, being themselves anion conducting materials. They can also improve the mechanical properties even at 100% RH. The final part discusses 3D fillers, with a large section on SiO2 and silicates, as well as other metal oxides including alumina, titania and zirconia nanoparticles as well as metal–organic frameworks (MOFs). The organic functionalization of the oxides plays a decisive role in increasing the performance of composite materials. The first type of polymers used as matrix are inert, non-ionic, used as an absorbent for alkaline solutions, including synthetic poly(vinyl alcohol) with the general formula [-CH2CH(OH)-]n and chitosan, an aminated polysaccharide, composed of N-acetyl-D-glucosamine and D-glucosamine linked by β(1-4) bonds. The second type of matrix are ionomers that present an intrinsic ionic conductivity. They include fluorinated and hydrocarbon polymers, especially commercial, inexpensive aromatic polymers such as poly (2,6-dimethyl-1,4-phenylene oxide) (PPO) and polysulfone (PSU). Both are thermoplastics, known for their toughness, resistance at high temperature and especially PPO for alkaline stability.
We cover in this review particularly ionic conductivity, mechanical properties and device performance, especially in fuel cells. Several tables in the manuscript summarize information on the reviewed AEM.

2. 1D Materials

2.1. Titanate (TNTs) and Halloysite Nanotubes (HNTs)

Titanate nanotubes (TNTs) were first described in the 1990s. These nanostructured tubular materials present excellent characteristics and performances [16]. Among them, titanium dioxide TiO2 NT exhibit high chemical stability and catalytic activity, a high specific surface area, strong metal–support interaction, and good performances in alkaline and acidic environments. TiO2 NT have a one-dimensional (1D) structure and exists as bundled tubes with a length between 1 and 220 µm and a diameter between 30 and 80 nm. The morphology of TNT depends on which synthesis method was used, including electrochemical anodization [17,18], a template-assisted method, sol–gel chemistry [19], etc. given all these characteristics, among other applications, TNT are used as filler in fuel cell membranes [16].
One of the earliest examples of composite AEM with TNT was presented in 2018 by Elumalai and Sangeetha [20]. The same authors incorporated TNT covalently linked with an imidazolium-based ionic liquid (IL) into quaternized polysulfone (QPSU) with triethylamine (TEA) [21]. The presence of the IL increased the IEC; the best conductivity was observed for 5 wt% IL-TNT (21 mS cm−1, 30 °C), due to the uniform distribution of the filler without any cluster formation. The same composite membrane reached a maximum tensile strength (TS) of 43 MPa and a fuel cell power density of 302 mW cm−2 at 60 °C under 80% RH. Quaternized TNT (QTNT) were also incorporated in quaternized poly(arylene ether ketone) (QPAEK) [22].
Halloysite nanotubes (HNTs), natural two-layered aluminosilicate clay minerals, were quaternized (QHNTs) with imidazolium groups bearing different functionalities (butyl, decyl, carbethoxyl, and benzyl moieties) and then embedded into a chitosan (CS) matrix [23]. Figure 2 shows the TEM images of HNTs and QHNTs. HNTs presented a clear tubular structure while after modification QHNTs showed a well-defined polymeric layer.
The incorporation of 5 wt% of HNTs elevated the TS and Young’s modulus (E) to 52 MPa and 1100 MPa, respectively. The membranes containing QHNTs functionalized with carbethoxyl groups showed the best hydroxide conductivity under hydrated conditions: with 7.5 wt% the conductivity reached 17 mS cm−1 at 90 °C, which is still modest.

2.2. Carbon Nanotubes (CNTs)

Carbon nanotubes are a relaxation of the spherical structure of fullerene rolled up on itself, obtaining the typical cylindrical structure. They can be divided into two types:
(a)
SWCNT (Single-Walled Carbon NanoTubes) consist of a single graphitic sheet wound on itself.
(b)
MWCNT (Multi-Walled Carbon NanoTubes) are formed by several sheets coaxially wound one on top of the other.
Very promising in their electrical properties, they are also characterized by extreme strength and flexibility, making them suitable for use as reinforcing fibers in composites. Several authors incorporated functionalized multi-walled carbon nanotubes (MWCNTs) into various types of AEM.

2.2.1. Poly(vinyl alcohol) (PVA)

Poly(vinyl alcohol) (PVA) is an inert polymer, soluble in warm water, that acquires ion conductivity through the addition of KOH or other alkaline solutions. One of the first examples of composite with PVA and MWCNT was reported by Pan et al. in 2011 and applied in direct methanol alkaline fuel cells (DMAFC). The methanol permeability decreased from ~3.6 × 10−7 cm2 s−1 for pristine PVA to 3.0 × 10−7 cm2 s−1 for composite membranes. The conductivity increased with the presence of MWCNT and was reported between 52 and 118 mS cm−1 (in 2 M and 6 M KOH at 30–60 °C). Figure 3 presents the DMAFC performance and the improvement due to the incorporation of CNTs [24].
PVA-functionalized MWCNTs were also prepared using an ozone-mediated method [25]. The MWCNTs were added to PVA solution and stirred at 80 °C for 3 h to graft the PVA polymer onto the MWCNT. The dry PVA/MWCNT films were then immersed in 1–8 M KOH solutions. PVA containing 0.05% of functionalized MWCNT exhibited the maximum power density of 39 mW cm−2 obtained with 2 M MeOH and 6 M KOH at 60 °C, consistent with the previous work.
Functionalized MWCNTs with pendant Fe3O4 nanoparticles (FeCNT) were incorporated into a PVA film [26]. High resolution HRTEM images showed a diameter of the MWCNT of 20−30 nm with Fe3O4 nanoparticles on the MWCNT surface. For a PVA–0.15 FeCNT sample an open-circuit potential of 0.87 V and a maximum power density of 88 mW cm−2 were obtained in DMAFC (2 M MeOH in 6 M KOH at 60 °C). To improve the chemical compatibility between CNTs and PVA, the CNTs were functionalized with PVA chains [27]. The maximum power density for 3 M EtOH in 5 M KOH solution was 65 mW cm−2 for the composite and 31 mW cm−2 for pristine PVA film.
Membranes composed of a semi-interpenetrating network structure of cross-linked PVA, poly-diallyldimethylammonium chloride (PDDA) and hydroxylated MWCNTs-OH were tested in AEMFC [28]. The cross-link was obtained by thermal treatment of PVA and subsequent immersion in acid solution with 10 wt% glutaraldehyde (GA). The tensile stress at break of PVA/PDDA membranes with 3 wt% MWCNTs-OH was 40 MPa, the elongation at break (%) and Young’s modulus were 12% and 783 MPa, respectively. The maximum OH conductivity was 30 mS cm−1 at RT with an improved alkaline stability. The (PVA/PDDA/1 wt% MWCNTs-OH) MEAs in an AEMFC with 0.5 mg Pt cm−2 at the anode and cathode side attained a power density of 66 mW cm−2 at 40 °C; using 3.2 mg cm−2 of cobalt phthalocyanine at the cathode side, the power density decreased to 14.0 mW cm−2.

2.2.2. Chitosan (CS)

Only a few examples of composites containing chitosan and CNTs are present in the literature. CS was used as a matrix for MWCNTs-OH and ionic liquids quaternized with isoquinoline moieties bearing ammonium groups [29]. Quaternized chitosan (QCS) and functionalized CNTs were prepared with the aim to improve the mechanical properties [30]. CNTs were also functionalized with quaternized silica obtained by in situ sol–gel methods [31]. The incorporation of 5 wt% of functionalized CNTs enhanced the conductivity to 43 mS cm−1 at 80 °C, about two times higher than bare QCS. DMAFC tests with composite membranes produced a maximum power density at 60 °C of 81 mW cm−2, higher than pure QCS (52 mW cm−2).

2.2.3. Aromatic Polymers

Except in one case [32], the imidazolium group (Im) was the preferred choice in all composites containing aromatic polymers.
Jin and Bai in 2013 blended methyl-Im quaternized PPO (PPO–MIm) with MWCNTs functionalized with a polymeric ionic liquid containing Im moieties (PIL(BF4)–MWCNTs) [33]. The conductivity increased in the presence of the filler (95% more with respect to PPO–MIm) together with an enhancement in TS up to 13 MPa.
Imidazolium ionic liquids (ImILs) modified carbon nanotubes (IL@CNT) were also inserted into imidazolium-modified poly(ether ether ketone) (Im-PEEK) [34]. Two types of ionic liquids (IL-M and IL-B) with different alkaline stability and chain length were chemically attached to CNTs (Figure 4). The authors stated that the introduction of IL@CNT provided additional ion hopping positions and 1D long range ion-conducting channels.
The composite containing IL-B@CNT showed a higher conductivity (135 mS cm−1 at 70 °C and 100% RH) confirming the beneficial effect of the long chain (pure Im-PEEK 80 mS cm−1), while the residual hydroxide conductivity (Im-PEEK/IL-B@CNT-8, after 48 h in 2 M KOH at 50 °C) was 77%. The peak power density in AEMFC was 81 mW cm−2 at 50 °C [34].
Im-PEEK was also used as a hosting matrix for poly(vinyl imidazole) functionalized carbon nanotubes (PVI@CNT) [35]. The imidazolium groups of the filler reacted with the Im-PEEK to form a cross-linking structure along the nanotubes. The nano-hybrid membrane with 15 wt% of the filler (Im-PEEK/PVI@CNT-15) showed a conductivity of 121 mS cm−1 at 70 °C and 100% RH. According to the authors, the improvement of the hydroxide conductivity was due to additional anion transport pathways built along the interface between the PVI@CNT and the polymer matrix, with an increase in the connectivity and the ion transport. The maximum power density in a single H2/O2 fuel cell was 129 mW cm−2 at 60 °C under 100% RH.
Via co-electrospinning, an imidazolium functionalized MWCNTs was incorporated into imidazolium functionalized polysulfone (Im-PSU) [36]. In fully hydrated conditions, the tensile stress increased from 5.6 MPa for a cast Im-PSU to 24.4 MPa for electrospun ImPSU with 0.2 wt% of MWCNTs. The maximum value of OH-conductivity was obtained for 0.4 wt% of FMWCNT with 100 mS cm−1 at 60 °C, but the residual conductivity after immersion in 1 M NaOH at 60 °C for 168 h was 62%.
Table 1 summarizes information on the various composites, including polymer and 1D filler type, remarks and references.
We have seen in this section different 1D nanofillers such as CNT, halloysite and titanate NT, commonly used for the fabrication of hybrid materials. One-dimensional tubular and well designed nanofillers can provide ion conducting groups, construct ion conducting pathways along the interphase region and improve the anti-swelling property of the nanohybrid membrane. For these reasons, as shown previously, both the mechanical properties and the residual conductivities after the alkaline test increase, and above all the performances in FC show an improvement compared to the bare polymers.

3. 2D Materials

3.1. Layered Double Hydroxides (LDHs)

Layered double hydroxides (LDHs) are inorganic lamellar ionic materials belonging to the group of anionic clays and their synthesis has a low cost. The structure of LDH is based on Mg(OH)2 brucite-type blocks where the replacement of M2+ with M3+ cations gives positively charged layers, balanced by mobile anions in the interlayer, which can be reversibly inserted. The lamellae are linked by Van der Waals forces. LDHs with an acceptable anionic conductivity and excellent stability in alkaline media can be used successfully as filler in AEM [37]. Well dispersed LDHs increase the mechanical properties of the matrix and can help to mitigate the loss of conductivity observed in the case of non-conducting fillers.

3.1.1. Poly(vinyl alcohol) (PVA) and Chitosan (CS)

One of the first attempts to use LDH in a composite date back to 2012 when Zhao et al. dispersed LDH in crosslinked PVA [38]. SEM images showed a good homogeneity without aggregates or chunks until 30 wt% of LDH (30LDH). FC with a PVA/20LDH membrane achieved a maximum power density of 82 mW cm−2 at 80 °C.
Exfoliated LDH nanosheets, obtained via the filtration process, were inserted in quaternized PVA [39]. The conductivity of the LDH(NO3) membrane after exfoliation and restacking was around 200 mS cm−1 at 80 °C, which seems surprisingly high especially for nitrate ions and also higher than the OH form. The improvement of conductivity was attributed to the exposed surfaces of the nanosheets, the enhanced water uptake, and long-range ordered ionic channels. For the composite with 10 wt% of PVA, (LDH(NO3)/QPVA-10) was reached at the same temperature 172.1 mS cm−1.
Other examples concern the use of mixed membranes composed of QCS and PVA; they hosted LDH intercalated with glycine betaine [40], an LDH flower-like hierarchical structure wrapped on SiO2 nanospheres (LDH@SiO2) [41], and carbon nanotubes coated with LDH (LDH@CNTs) [42]. In the last example, LDH nanosheets were anchored on carbon nanotubes by a NH4F-assisted in situ coating method (Figure 5) [42].
The conductivity of the composite membrane was 47 mS cm−1 at 80 °C, in comparison with 29 mS cm−1 for the pristine membrane. With the increase in the content of LDH@CNTs the alkaline stability improved and QCS/PVA-4%-LDH@CNTs membrane showed a residual ionic conductivity of about 73% after 192 h in 1 M KOH at 40 °C. The performance with 1 wt% of filler in a DMAFC (2 M MeOH, 5 M KOH) at 80 °C gave a maximum power density of 107 mW cm−2 [42].

3.1.2. Polysulfone (PSU)

An attempt to use an inexpensive commercial polymer was proposed in 2017 by Di Vona et al. [37]. The composites were prepared from QPSU grafted with TMA or 1,4-diazabicyclo[2.2.2]octane (DABCO) and 14 wt% of LDH, synthesized by the urea method with a composition Mg0.62Al0.38(OH)2(Cl)0.38·0.6H2O. The ionic conductivity at 60 °C as a function of RH is reported in Figure 6: composite membranes showed lower conductivity values due to the lower water uptake, while polymers with TMA were more conductive because of the higher water content. The mechanical properties in fully humidified conditions were improved, with a three-fold increase in Young’s modulus for composites (~600 MPa vs. ~200 MPa). The membranes were treated in alkaline conditions at 60 °C without losing their properties.
Meanwhile, Zhang prepared LDH by the urea method, which was incorporated from 3 to 10 wt% into TMA-PSU membranes [43]. The 5 wt% LDH nanocomposite membrane showed the highest TS (21 MPa), the lowest elongation at break (11%), and the highest ionic conductivity (24 mS cm−1 at 60 °C). Pizzoferrato et al. proposed a composite membrane based on PSU-DABCO and ZnAl-LDH containing an ionic liquid (1-butyl-3-methylimidazolium hydrogen sulfate, BmimSO4) [44]. This strategy enabled the intercalation of IL into the interlamellar space of LDH with the aim of enhanced ionic conductivity. In fact, at 100% RH and 25 °C, the membranes with 7 wt% of filler reached a conductivity of 25 mS cm−1 in OH form and maintained 16 mS cm−1 after alkaline treatment in 2 M KOH at 60 °C for 24 h [44].

3.1.3. Poly(phenylene oxide) (PPO)

A porous-sandwich structure was synthesized by electrostatic spraying of ammonium layered double hydroxide (QLDH) on the surface of triple-cation side chains of PPO (TC-PPO) [45,46]. The composite membranes with higher IEC values (from 3.12 to 3.39 meq g−1), due to the presence of ammonium groups in QLDH layers, showed a high hydroxide conductivity (122 mS cm−1 at 80 °C). The FC performance at 60 °C achieved a maximum power density of 267 mW cm−2 at a current density of 554 mA cm−2, values higher than those of membranes without QLDH but still far from those recorded for from Nafion 112 [46]. The same authors synthetized LDH functionalized with 3-hydroxy-6-azaspiro [5.5] undecane cations (OH-ASU) to increase the ion exchange capacity [45]. The ASU-LDH filler was combined with TC-PPO to fabricate three-decker ASU-LDH/TC-PPO hybrids with IEC up to 3.90 meq g−1. The OH conductivity, shown in Figure 7, increased from 79 mS cm−1 for TC-PPO to 112 mS cm−1 for ASU-LDH/TC-PPO-50. After stability tests in 1 M NaOH at 80 °C for 588 h, only an 11.5% drop in OH conductivity was observed.
An electric field was sometimes applied to induce the formation of ion channels along the through-plane direction [47,48]. Aligned LDH functionalized with N-spirocyclic ammonium groups and PPO showed a maximum ion conductivity of 110 mS cm−1 at 80 °C and an enhanced alkali resistance with 83% residual conductivity [48]. The good stability was attributed to the formation of electric field-induced ion channels that reduced the local alkaline concentration near quaternized groups and decreased the possibility of the hydroxide to attack ammonium groups.
Composite membranes with DABCO-PPO and 30 wt% MgAl-LDH were prepared in 2020 by Pasquini et al. [49]. The high amount of the second phase was chosen in order to understand the effect of the inorganic filler on alkaline and hydrolytic degradation. The stability behaviour, studied by different techniques, was comparable to the initial ionomer showing a fast degradation in the first hours.

3.1.4. Poly(vinylidene fluoride) (PVDF)

Only one example was reported in the literature concerning PVDF: a bioinspired geometrical templating of PVDF substrate containing electrospun MgAl-LDH for solid-state AFC [50]. The hybrid substrate was morphologically similar to the coconut flower (Cocosnucifera). Efficient pore-filling by LDH was demonstrated by SEM and the inorganic phase enhanced TS two-fold (14 MPa). The Cl conductivity at 80 °C was 92 mS cm−1 at 100% RH and 18 mS cm−1 at 50% RH.

3.2. MXenes

MXenes are a family of 2D layered materials which include transition metal nitrides, carbides, and carbonitrides produced by selective exfoliation of MAX phases, where “M” refers to early d-block transition metals, “A” to groups 13 and 14 (main sp elements) and “X” to carbon or nitrogen atoms [51]. MXenes present several properties, such as hydrophilicity, fast ion transport and intercalation, large surface area, etc., provoking interest in many scientific applications including FC.
By solution blending method two kinds of 2D nanomaterial Ti3C2Tx (Tx stands for the hydrophilic surface terminations-O, -OH, -F, etc.) were incorporated into QPSU [52]. The mechanical strength and the conductivity were slightly increased adding 3 wt% of LiF-Ti3C2Tx and NH4HF2-Ti3C2Tx. The maximum power density in a single cell at 60 °C with H2/O2 atmospheric pressure and Pt loading of 0.4 mg cm−2 in both sides for QPSU/LiF-Ti3C2Tx and QPSU/NH4HF2-Ti3C2Tx attained 74 mW cm−2 and 101 mW cm−2, respectively. In 2018, Wang et al. dispersed imidazolium functionalized MXene in a CS matrix [53]. The pure CS membrane had a TS of 28 MPa, after the addition of QMXene-NH2, the TS improved significantly up to 41 MPa. The conductivity of AEM with 7.5 wt% of QMXene-NH2 at 100% RH increased from 1.5 to 4 mS cm−1 [53].
The AEM composites with 2D fillers are summarized in Table 2, including the type of polymer and 2D filler, remarks and references.

3.3. Graphene Oxide (GO) and Graphene

A very large amount of work and effort was devoted to composite AEM containing graphene oxide (GO) or graphene. The main features for graphene are an extreme mechanical resistance and great flexibility, for GO its hydrophilicity and dispersibility in water, due to epoxy and hydroxyl groups on the basal planes, and carboxyl and carbonyl groups at the edges of layers. In particular, GO was functionalized with many groups creating original and sometimes complex structures.

3.3.1. Poly(vinyl alcohol) (PVA)

PVA-based matrices were often used to host different carbon materials such as exfoliated graphene nanosheets [54], or graphene and sulfonated graphene nanoparticles (PVA/chitosan blend membranes) [55]. Cross-linked PVA and quaternized polyethyleneimine (QPEI) hosted GO functionalized with silica by epoxide ring-opening reactions using APTEOS [56]. Cross-linked composites composed of PVA, quaternized with glycidyltrimethylammonium chloride (GTMAC), and GO (QPVA/GO) was prepared to decrease the EtOH permeability [57].
Magnetite nanoparticles on GO were distributed along the through-plane direction in a QPVA matrix by applying an external magnetic field (MF) during the film drying step [58]. The QPVA/0.1%Fe3O4@GO nanocomposites presented the highest ionic conductivity with 55 mS cm−1 and a maximum power density of 200 mW cm−2 at 60 °C and 2 M MeOH [59].

3.3.2. Polysulfone (PSU)

Many functionalized graphene or graphene oxide-based composites contained PSU or QPSU as a host matrix [60,61,62,63,64,65,66].
TEM images of quaternized polymer brush-functionalized graphene (QPbGs) into QPSU showed a uniform distribution with the amount of QPbGs below 1 wt%. The HCO3 conductivity was 56 mS cm−1 at 80 °C with 1 wt% of filler [67]. Reduced GO (rGO) functionalized with polydopamine (PDArGO) and QPSU quaternized with 3-(dimethylamino)-1-propylamine reached a hydroxide conductivity of 61 mS cm−1 at 80 °C (QPSU-1.5%-PDArGO) [68]. The Young’s modulus raised from 868 MPa for QPSU to 1843 MPa, and the WU increased from 18% to 30% at 80 °C.
TMA-PSU was cross-linked (XL-QPSU) by rGO modified with short- and long-chain tertiary amines (SrGO and LrGO) [69]. Membranes with elastic long-chain (LrGO) displayed higher ion conductivity due to major nanophase separated morphology. The XL-QPSU-2%-LrGO OH conductivity was 76 mS cm−1 at RT, with respect to a conductivity of 49 mS cm−1 for the pristine AEM. The same authors modified rGO with pyrene-containing tertiary amines (TrGO) and polymers (PrGO) via π–π interactions [70]. The functionalized rGO was used to cross-link TMA-PSU. The XL-QPSU-2%-PrGO reached a hydroxide conductivity at 80 °C of 118 mS cm−1 and after alkaline treatment in 1 M NaOH for 500 h at 60 °C, the residual conductivity was 88% higher than un-crosslinked QPSU.
Composite AEMs were also expected to reduce the vanadium permeability in vanadium redox flow batteries (VRFB), due to the Donnan exclusion effect. A long alkyl chain (C16) grafted to quaternized GO was bonded to DABCO-PSU and the resulting composites used in VRFB [71]. The long alkyl chain increased the hydrophilic–hydrophobic nanophase separation. The selectivity, assessed by a ratio of ion conductivity and vanadium ion permeability, for 5 wt% of the modified GO, was 19 × 105 S min cm−3 and the Coulombic efficiency (CE), was 98% at 60 mA cm−2. The good results were attributed to the high ionic conductivity and the extremely low vanadium cross-over.

3.3.3. Poly(phenylene oxide) (PPO)

Although PPO is widely employed as a polymer in AEMs, only a few examples concern its use in composites with GO.
Brominated PPO and polyethyleneimine (PEI)-modified GO was proposed by Kulshrestha et al. [72]. A blend of GO (QGO) and cellulose (QCel) quaternized with DABCO, and subsequently XL with DABCO-PPO, displayed an OH conductivity of 114 mS cm−1 at 25 °C [73]. The stability decreased with the increase in QCel amount, due to its oxygenated functionalities susceptible for the nucleophilic attack, and increased with the content of QGO, very stable in alkaline conditions.
An imidazolium-functionalized PPO (Im-PPO) and an ionic liquid-functionalized GO were used to prepare composite membranes with a hydroxide conductivity of 79 mS cm−1 at 80 °C [74]. The alkaline test in 2 M NaOH at 80 °C over 480 h showed a decrease in IEC from 1.90 to 1.34 meq g−1. FC tests at 60 °C for 0.5 wt% of IL-GO composite presented a maximum power density of 136 mW cm−2 with a current density of 300 mA cm−2 under 100% RH.
Recently, commercial membranes based on PPO (Fumion® FAA) were filled with different amounts of commercial graphene with a surface area of 500 m2/g [75]. The SEM morphology showed a uniform graphene inclusion into the polymer matrix; no crack formation and increased roughness were found. The maximum value of OH conductivity was reached for the sample containing 50 mg of graphene and was 113 mS cm−1 in 0.01 M KOH at 80 °C.
Another commercial membrane (JAM-II-07, Yanrun, China) was modified with sulfonated reduced graphene oxide (S-rGO) nanosheets and used in the electrodialysis process [76].

3.3.4. Polybenzimidazole (PBI)

Polybenzimidazole (PBI) is a well-known matrix in proton exchange membranes due to its performance at high temperatures; in AEM, PBI can be used after alkaline doping. To overcome the barrier due to the release of alkaline dopants in operative conditions, various strategies were used with the support of GO or rGO.
Zeng et al. proposed a MEA formed by a porous PBI membrane (sp-PBI) and a composite as binder formed by PBI and rGO on which the electrocatalyst was deposited [77]. The MEA reached a maximum power density of 544 mW cm−2, higher than conventional MEA (397 mW cm−2), as shown in Figure 8.
GO nanosheets were coated onto a PBI surface using a spin coater. The thickness of spin-coated with 2 wt% GO was 1–2 μm [78]. The Young’s modulus was improved to 1040 MPa and the tensile strength reached 50 MPa, effects attributed to hydrogen bonds between the edges of GO nanosheets. In DMFC the maximum power density at 80 °C reached 200 mW cm−2 with Pt-based catalysts [78]. Two types of GO nanosheets were prepared by microwaves (MGO) and by the modified Hummer’s method (NGO) [79]. MGO was less hydrophilic and more thermally resistant due to it having fewer ether groups and more sp3 C–C bonds. The PBI composite with 1 wt% of filler showed an increased conductivity by 38% for NGO and 29% for MGO with respect to pure PBI. Voc and Pmax for NGO were 0.7 V and 310 mW cm−2, respectively.

3.3.5. Other Aromatic Polymers

Quaternized PAEK was filled with different amounts of rGO [80] while poly(phthalazinone ether ketone) (PPEK) hosted cyclodextrins modified with trimethylammonium groups on GO (QA-CDβ@GO) [81]. Imidazolium-functionalized Im-GO inserted into Im-PEEK showed a tensile strength of 36 MPa, 37% higher than pristine Im-PEEK [82]. The conductivity at 70 °C and 100% RH reached 140 mS cm−1. The hybrid showed in H2/O2 FC tests a power density of 50 mW cm−2 at 50 °C, 122% higher than the pristine membrane.
Recently, DABCO quaternized poly(arylene ether) (QPAE) was filled with ammonium functionalized GO (GO-(APTS-c-PTMA)) using (3-aminopropyl)triethoxysilane (APTS) and (3-bromopropyl)trimethyl ammonium bromide (PTMA) as chemical cross linkers [83]. The hybrid with 0.7 wt% of filler showed an OH conductivity of 114 mS cm−1 at 90 °C and the remaining ionic conductivity, after soaking in 2 M KOH at 80 °C, was 75% of the initial value. The H2/O2 FC performance at 70 °C and 100% RH showed a maximum power density of 136 mW cm−2 and a current density of 317 mA cm−2.

3.3.6. Other Aliphatic Polymers

Some aliphatic polymeric matrices not containing ether bonds were explored, with the aim of decreasing degradation under alkaline conditions.
Liu et al. synthesized a perfluorinated anion-conducting polymer (I-PFSO2NH2-Cl) in 2017 with ionic liquid functionalized graphene nanoribbons (IGNRs). The composite with 1.0 wt% of filler attained a conductivity of 121 mS cm−1 at 80 °C in liquid water. The single cell performance with IGNRs/Pt electrocatalysts showed a maximum power density of 197 mW cm−2 and a current density of 372 mA cm−2 [84]. Cross-linked quaternized poly(styrene-b-isobutylene-b-styrene) (QSIBS) was the host matrix for organo-modified graphene oxide (GOAN) quaternized with TMHDA [85] and for covalently linked graphene [86]. GO modified with butylvinylimidazolium (GO/IM) was incorporated inside a complex matrix composed by para-methyl styrene/butylvinylimidazolium (PMS/b-VIB) and poly(4,4′-diphenylether-5,5′-bibenzimidazole) (DPEBI) [87]. The synthesis of GO/IM is reported in Figure 9. The nanohybrid membranes attained an OH conductivity of 102 mS cm−1 at 100 °C. The activation energy was between 16 and 19 kJ mol−1 corresponding to a Grotthuss-type mechanism; the lower value was observed for 2 and 3 wt% filler content.
A GO multilayer paper was also employed as a matrix. The membranes were treated with KOH and used in FC [88].

3.4. Carbon and Boron Nitride (BN)

Recently, some examples were reported in the literature on the use of nitrides, characterized by extreme hardness, to improve the mechanical properties and decrease the crossover of membranes.
Graphitic carbon nitride (g-C3N4) nanosheets were used as filler in modified quaternary aminated poly(arylene ether sulfone) (QPAES) [89]. The Young’s moduli were in the range of 1150–1670 MPa and increased with respect to bare QPAES (1110 MPa). AEMFC tests at 60 °C, for 0.6 wt% filler, showed the highest power density with 68 mW cm−2. g-C3N4 nanosheets were also inserted into PAEK functionalized by Menshutkin reaction with TMA (QPAEK) [90]. The Young’s modulus and the TS increased as expected; the maximum power density in a single H2/O2 FC was reached for the QPAEK-CN-0.5 membrane with 49 mW cm−2 at 80 °C. PPO quaternized with N-methyl morpholine was used by Rathod et al. in 2020 to prepare composite membranes with a different amount (1–5 wt%) of functionalized boron nitride (BN) [91]. Membranes with 5 wt% exhibited the best performances; the ionic conductivity was acceptable (6.3 mS cm−1 at 30 °C) and the MeOH permeability decreased (5.15 × 10−8 cm2 s−1) indicating that BN acted as a barrier for methanol limiting the cross-over.
Table 3 summarizes information on the various composites, including the type of polymer and other 2D filler, remarks and references.
We have seen in this section various 2D nanofillers, such as LDH, MXene, graphene and graphene oxide, etc., commonly used for the fabrication of hybrid materials especially with commercial and low cost polymers such as PPO and PSU. The purpose was to maintain good conductivity, also by functionalizing the inorganic fillers with ionic conducting groups, while trying to enhance the mechanical properties and reduce the permeability of gases in the FC, thanks to the 2D geometry of these materials.

4. 3D Materials

4.1. Silica and Silicates

Silicon dioxide and silicates are the most abundant classes of minerals in nature, and their cost is practically negligible. Phyllosilicates, such as montmorillonite and palygorskite, are hydrated silicates of aluminium and/or magnesium with the ability to exchange ions. Notable successes were obtained by using Si and its derivatives as fillers in proton exchange membranes [92,93]; later their use was extended to AEM. In this review, we will limit the description of the progress achieved to silica and derivatives not obtained via the sol–gel process. Sol–gel silica composites were described in a recent review by Sgreccia et al. [94].

4.1.1. Poly(vinylidene fluoride) (PVDF) and Poly(vinyl alcohol) (PVA)

One of the first examples was proposed by Zuo et al. who prepared a composite with TMA-PVDF and SiO2 [95]. The membrane with 2 wt% of SiO2 showed the best specific conductivity (3 mS cm−1). More recently silica-coated PVDF (SiO2@PVDF) electrospun nanofibers were quaternized and inserted in QCS [96]. Tests in alkaline DMFC, performed at various MeOH concentrations, showed a maximum power density at 80 °C up to 99 mW cm−2 (2 M MeOH), presenting only 4% of performance loss after 100 h in chronoamperometry test.
PVA was often used as matrix for silica hybrids [97,98]. A DMAFC maximum power density (97 mW cm−2, 80 °C with 2 M MeOH + 6 M KOH) was achieved with quaternized PVA and 5 wt% of nanosized (14 nm) fumed silica (FS) due to its higher free volume and hydroxide ion transfer phenomenon [99]. A different approach was proposed by Lu et al. where cellulose nanocrystal (CNC)-based composite films were mixed with 40% of hydrophobic binder (PVA : silica gel = 1:2) [100]. Due to the hydrophilicity and dimensional stability of CNCs, membranes exhibited high water uptake (~80%) but low water swelling (~5%).
For PEM electrochemical reactors, a blend of PVA and CS was doped with organic ionomers (4VP, methyl chloride quaternary salt resin), commercial ionomer filler (AS4, structure not disclosed by the industry), inorganic titanosilicate (AM-4, containing Na+ ions) and layered stannosilicate (UZAR-S3, Na7Sn2Si9O25) [101]. The best performances were obtained for UZAR-S3/CS:PVA and 4VP/CS:PVA composites.

4.1.2. Aromatic Polymers

Aromatic polymers were also a preferred choice in the building of silica-containing membranes for fuel cells, diffusion dialysis (DD), and redox flow batteries. Pan et al. in 2015 improved the efficiency of acid recovery by DD with the fabrication of a quaternized PPO/SiO2 hybrid material (QPPO–SiO2) obtained by electrospinning and post-treatment (solvent fumigation and hot-press) [102]. The main fiber diameters were between 200–300 nm with an average of ~250 nm as shown in Figure 10.
In a simulated polishing waste solution containing 1 M HCl and 0.225 M FeCl2, the membrane exhibited both a higher acid permeability (UH) with 0.053 m h−1 and selectivity (S, 68.05) in comparison with the direct casting QPPO–SiO2 hybrid membrane. In addition, compared with a commercial DF-120 membrane the hot-pressed electrospun QPPO–SiO2 membrane showed more than seven times higher UH and approximately three times higher S, underling its advantage in DD application [102]. Quaternized polyethersulfone (TMA-PES) composite membranes were fabricated with the incorporation of three functionalized SiO2 nanoparticles containing propylamine (ASi-I), trimethylpropylamine (ASi-II) and melamine-based dendrimer amine groups (ASi-III) [103].
Recently Chen et al. prepared composites with a “hamburger structure” using quaternized PPO with a triple-cation precursor (TA-PPO) as matrix and 1,2-dimethylimidazolium-silica (Im-SiO2) as dopant. Im-SiO2 was placed on the surface of TA-PPO membrane to protect the matrix from the attack of OH and radicals [104]. The composite membrane showed a low swelling ratio (8.2%), and a high OH conductivity (105 mS cm−1 at 80 °C).
Quaternized mesoporous silica nanoparticles (QMSNs) were mixed with PSU quaternized with TEA [105] and TMA [106]. The homogeneous dispersion of QMSN in the QPSU matrix, highlighted by the SEM images, improved the mechanical properties: the Young’s modulus increased from 1990 for pure QPSU to 2250 MPa for 20 wt% of QMSNs while the composite with 15 wt% of filler showed the highest CO32− conductivity, reaching 20 mS cm−1 at 80 °C [106]. Afterwards mesoporous silica (SBA-15) was functionalized with imidazolium-based ionic liquid and inserted in QPSU. The presence of covalent bonds between IL and SBA-15 was verified by solid-state NMR [107]. The QPSU/3%IL-SBA-15 showed the best fuel cell performance with an OCV of 0.87 V and a maximum power density of 278 mW cm−2.
TMA-PSU was a host membrane for two functionalized montmorillonites, containing cetyl trimethyl ammonium chloride (MMT-1) and (3-aminopropyl) triethoxysilane (MMT-2) [108]. The membrane containing 5 wt% of MMT-1 showed the highest conductivity with 47 mS cm−1 at 95 °C. The alkaline stability was evaluated in 2 M NaOH for 120 h at 60 °C and the residual ionic conductivity was around 80% [108]. TMA-PSU (QPSU) was also used as matrix for natural hydrophilic “nanofiber-like” palygorskite (Pal) particles [109]. SAXS showed a good hydrophilic–hydrophobic microphase separation that led to a high OH conductivity and QPSU/Pal-0.5 reached 93 mS cm−1 at 80 °C. The TS was between 24 and 45 MPa, the Young’s modulus 1177–1848 MPa and the elongation at break 3–12%.
Another aromatic polymer, quaternized cardo-poly(ether ketone) (QPEK-C) were doped with a quaternized ORMOSIL (TMSP-TMA+Cl) [110]. The composites were used in all vanadium redox flow batteries; Figure 11 shows the sulfate ion conductivity with different percentages of TMSP-TMA+ as a function of time. The best results were obtained with 10–20 wt% of TMSP-TMA+.
In VRFB, the CE at 100 mA cm−2 was 99% for both, QPEK-C and QPEK-C/20 wt% TMSP-TMA+, compared to 95% for Nafion® 212. The battery capacity was 10% lower over 30 charge/discharge cycles (~60 h) while for Nafion® 212 in similar conditions a loss of 30% was observed [110].
Table 4 summarizes information on the various composites, including the type of polymer and 3D filler, remarks and references.

4.2. Metal Oxides and Derivatives

One of the major uses of metal oxides as filler in AEM is in improving the mechanical properties of soft polymers such as PVA. Various oxides are used, including Al2O3, a well known hard material characterized by poor thermal and electrical conductivity, TiO2, non-expensive and non-toxic with high chemical stability, and ZrO2, a polymorphous crystalline oxide with high ionic conductivity when doped with acceptor cations and low electronic conductivity. The chosen matrices were mainly PVA and aromatic polymers.

4.2.1. Aluminum Oxides

PVA quaternized with glycidytrimethyl ammonium chloride and 10 wt% Al2O3 (QPVA/Al2O3) displayed in DMAFC a maximum power densities of 36 mW cm−2 with 4 M KOH + 4 M CH3OH, as shown in Figure 12 [111]. Although SEM images indicated some chunks and aggregates randomly distributed on the surface, the storage modulus slightly increased with respect to pure PVA (172 vs. 151 MPa at 100 °C).
Composite membranes with smooth and dense morphology were formed by quaternary trimethylammonium PSU with different loadings of Al2O3 powder (1–4 wt%) [112]. The presence of filler increased the swelling degree, the water uptake, and the ionic conductivity.

4.2.2. Zirconium Oxides

The morphology of composite membranes of TMA-PSU and zirconia was investigated by TEM, revealing a homogenous distribution when the particles size were around 10 nm [113]. The membrane with 10 wt% of ZrO2 showed the best performance, reaching a maximum conductivity of 15 mS cm−1. The maximum power density presented a value of 250 mW cm−2 with an OCV of 0.91 V at 60 °C [113]. Based on Im-PSU, recently, Rambabu et al. prepared composites with different percentages of zirconia (2.5, 5, 7.5 and 10 wt%), through a phase inversion method [114]. The conductivity attained was 80 mS cm−1 at 50 °C, 47% higher than pure Im-PSU. The TS showed values of 47 and 43 MPa before and after the alkaline treatment, while pure Im-PSU presented 31 and 25 MPa, respectively. The FC performance, using Pt/C catalysts, showed with 10 wt% of ZrO2 an OCP of 1.04 V, a maximum power density of 270 mW cm−2 and a current density of 640 mA cm−2.
Another quaternized aromatic polymer (QPAES) was used to prepare composite membranes with nanozirconia [115,116]. The introduction of the second phase enhanced the OH conductivity of cross-linked multiblock copoly(arylene ether sulfone) (XL TMA-coPAES), with 7.5 wt% of ZrO2 the composites showing the best conductivity (55 mS cm−1 at 80 °C) and the best alkaline stability, retaining 94% of conductivity after immersion in 1 M NaOH solution at 60 °C for 340 h. The highest Young’s modulus was reached for 10 wt% of ZrO2 with 492 MPa, around two times higher than virgin polymer. Fluoropolymer-based membranes were obtained from a free-radical desulfurization coupling reaction on Nafion NR-50 followed by grafting and filling with ZrO(ClO4)2 [117]. The chemical stability was evaluated in a reactor with a 7 M KOH solution of EtOH/H2O 15%/85% at 120 °C for 2200 h. The IEC after the stability test was practically unchanged. The high stability was ascribed to the fluorinated polymeric structure and to covalent C–C bonds that link the ZrO(ClO4)2 nanoclusters and polymer side chains [117].

4.2.3. Titanium Dioxide and Titanates

Many aromatic polymers were used as host matrices: PSU and quaternized polystyrene-block-poly(ethylene-ran-butylene)-block-polystyrene (PSEBS) [118], TMA-PSU [119], vinylbenzyl chloride-divinylbenzene copolymers (AEH) [120], PSU and TMA-PPO [121].
The effect of the doping with hydrophilic (tri(hydroxymethyl)propane, TMP) and hydrophobic (polymethyl-hydrosiloxane, PMHS) TiO2 nanoparticles was studied using DABCO-PSU as the matrix [122]. The DMA and DSC studies showed a Tg around 250 °C, due to a partial crystallization of the polymer, with a significant decrease in the contact angle (67°) for composites with hydrophilic character. The ionic conductivity was higher for the hydrophobic filler (PMHS-TiO2) related to a more homogeneous dispersion, as observed by AFM images.
Some examples of composites concern the use of ionic liquids and nanoparticles with the aim of simultaneously increasing conductivity and mechanical properties. 1-Methy-3-methylimidazolium IL and nano-TiO2 were mixed to TEA-PPO [123]. The conductivity increased with the amount of IL and TiO2 and the best mechanical properties were reached for 15 wt% IL and 1 wt% of TiO2 with Young’s modulus of 921 MPa and 6% of elongation at break. The degradation rates of the ion conductivity measured after treatment in 4 M NaOH solution at RT for 280 h, showed for 1 wt% of TiO2 a reduction of around 30% instead of 80% without the nanofiller, confirming the role of TiO2 in stabilizing the IL in the membranes [123]. The same authors inserted methyl, ethyl, hydroxyethyl IL and TiO2 into a TEA–PPO matrix [124]. The stability tests, in 4 M NaOH at RT for 288 h, revealed for the ethyl derivative a 92% retention of conductivity while the pristine membrane retained 78%.
PVA based-membranes were used in microbial fuel cells (MFCs) [125]. The comparison between TiO2–PVA quaternized with trimethylammonium chloride (QAPVA), commercial Nafion 117, and LeHoAM-III (Hangzhou Lvhe Environmental Technology Co., Ltd., China) showed a maximum power density at 35 °C for TiO2-QAPVA (125 mW cm−2), two and three times higher than Nafion 117 and LeHOAM-III, respectively. The TiO2-QAPVA superior performances were attributed to a good oxygen resistance of the membrane.
PVA membranes, filled with spherical calcium titanate nanoparticles (CaTiO3) with orthorhombic perovskite structure, were proposed for redox flow batteries [126]. The distribution was uniform and homogenous when the nano-CaTiO3 filler was less than 20 wt%.

4.3. Other Inorganics

Homogeneous composite membranes with PTFE and Sn0.92Sb0.08P2O7 exhibited conductivities from ~10 to 100 mS cm−1 between 75 and 200 °C [127]. In the fuel cell tests, the maximum power density reached 94 mW cm−2 at 100 °C.
In Direct Borohydride Fuel Cells (DBFCs), PVA membranes with CoOOH functionalization showed a better performance than pristine membranes [128]. The maximum power densities obtained in the test cells at 30 °C with and without CoSO4 were 144 and 72 mW cm− 2, respectively (Figure 13).
PVA was employed to prepare composites based on TMA-PVA/chitosan/molybdenum-disulfide (QPVA/CS/MoS2) with different amounts of MoS2 nanosheets [129]. The membranes with 1.0 wt% MoS2 showed a TS of 33 MPa. The MeOH permeability of pure QPVA/CS was 1.0 × 10−7 cm2 s−1 and decreased to 0.2 × 10−7 cm2 s−1 with the addition of 1.0 wt% of MoS2.
Zinc oxide nanoparticles (ZnO) were used as fillers in blended TMA-PPO and PSU [130]. The maximum power density at RT was 69 mW cm−2 (current density 220 mA cm−2), three times high than pure QPPO at the same conditions.
AEM with flame-resistance properties were prepared with PBI and 1-butyl-3-methyl imidazolium phosphotungstate (PWA-IL) [131]. SEM evidenced an irregular crystal structure of PWA-IL hybrids. After modification, the PBI/(PWA-IL) showed a lower TS (65 MPa) than pristine PBI (80 MPa), while the elongation at break (PWA-IL 1:4) increased from 6% to 13%. The anionic conductivity at 80 °C of PBI/(PWA-IL 1:5) containing 20 wt% of PWA-IL, was 76 mS cm−1, higher than pristine PBI (36 mS cm−1).

4.4. Metal Organic Frameworks (MOFs)

Crystalline metal–organic frameworks (MOFs) are formed by metal ions coordinated with rigid organic ligands to form high porosity structures and are therefore characterized by very large internal surface areas. With tunable functionality and well-defined channels, MOFs have inspired a new class of ion-conductive compounds. In contrast to the extensive studies on proton-conductive MOFs [132] and related membranes, rare reports focus on MOFs in the preparation of AEM.
Vinyl benzyl chloride (VBC) monomers were impregnated into nanopores of zeolitic imidazolate framework (ZIF-8) and then aminated to obtain poly vinyl benzyl trimethylammonium chloride (PVBTAC) as described in Figure 14 [133]. The Brunauer–Emmett–Teller (BET) surface analysis showed, after incorporation of the polymer, a pore volume of ZIF-8 from 0.49 to 0.26 mL g−1. The absence of a cross-linking agent in the polymerization ensured the formation of linear chains in the structure.
ZIF-8 was also added in a different amount to PVA [134]. To enhance the electrochemical properties at low humidity, choline hydroxide ionic liquids were used in PVA-ZIF-8 system [135]. The conductivities of IL@ZIF-8/IL/PVA composites were tested at 33% of RH and improved from 0.11 mS cm−1 at 25 °C to about 1 mS cm−1 at 60 °C for 20 wt% of IL@ZIF-8.
PVA-ZIF-8 composite membranes were also used for DMAFCs [136]. The alkaline stability test for PVA-40.5% ZIF-8 after immersion in 6 M KOH for 24 h and 168 h showed a decrease in conductivity by 14%, for pure PVA 31%. For the same composition, a maximum power density of 173 mW cm−2 was achieved in DMAFC at 60 °C.
A sandwiched AEM formed by porous bromomethylated-PPO with entrapped cationic MOFs coated with PVA on the two sides was prepared for DMAFCs [137]. PVA coating limited the MeOH crossover and entrapped cationic MOFs worked as OH conductive channels. The OH conductivity was 145 mS cm−1 at 80 °C and the MeOH permeability was 3.68 × 10−7 cm2 s−1 [137].
The incorporation of chloromethylated MIL-101(Cr) into chloromethylated PEEK, followed by quaternization with imidazolium, gave Im-PEEK/ImMIL-101(Cr) membranes [138]. ImMIL-101(Cr) was uniformly distributed into the Im-PEEK matrix up to 1.0 wt%; above this value a slight aggregation of the filler was observed. The TS of the composite membranes reached the highest value of 35 MPa with the incorporation of 10 wt% of ImMIL-101(Cr), 47.5% higher than pure Im-PEEK, the presence of filler reducing the elongation at break and increased the rigidity of membranes.

4.5. Carbon Dots (CDs)

Although the CD particle is classified with zero dimensions, we will deal with it in this section for simplicity.
Yuan et al. in 2019 prepared nanocomposites formed by an Im-PSU matrix and quaternized carbon dots (QCDs) derived from citric acid and ethylenediamine as shown in Figure 15 [139].
The Young’s modulus rose to 1600 MPa; the TS 70 MPa and the elongation at break rose to 15.5 %. For Im-PSU-1.0%-QCDs, the stability tests showed 61% of OH conductivity retention, (original value 109.3 mS cm−1), after 500 h in 1 M NaOH at 60 °C.
Composite AEMs based on chloromethylated PSU were developed for the DD process in the recovery of acid from the waste stream in 2020 [140]. The composites were prepared with different wt% (0.1, 0.5, and 1.0) of graphene quantum dots (GQDs) synthesized by chemical oxidation of MWCNTs, followed by quaternization with TMA (called SCE-0.1, SCE-0.5, and SCE-1.0). The conductivity for SCE-0.1 attained was 15 mS cm−1. TS and elongation at break increased, respectively, by 37% and 28% for a composite with 1.0 wt% GQDs with respect to pristine SCE. Acid recovery with 3 M HCl solutions was 30% for pristine membranes, while for SCE-1.0 it increased to 44%.
Table 5 summarizes information on the various composites, including the type of polymer and other 3D filler, remarks and references.
We have seen in this section different 3D nanofillers such as silica, ZrO2, MOF and CD commonly used for fabrication of hybrid materials especially with commercial, low-cost polymers such as PPO, PSU and very attractive PVA. The greatest efforts have focused on silica and functionalized derivatives and oxides, especially zirconia. The purpose was to increase conductivity, slightly improve IEC, enhance mechanical properties especially Young’s modulus and increase the power density in FC tests. Generally, the stability in an alkaline environment improves, but not significantly except for fluoropolymer-based anion-conducting membranes containing ZrO(ClO4)2 where the stability seems to exceed 2000 h.

5. Conclusions

This review on composite anion exchange membranes with inorganic fillers allows some general conclusions. We divided the results according to the dimensionality of the fillers. Some interesting synthetic efforts versus the innovative functionalization of nanoparticles can be recognized. Generally, an enhancement of mechanical properties (TS, Young’s modulus) is observed in composites. The ionic conductivity is often reduced with some possible mitigation by an intrinsic ion conducting fillers or little improvement with functionalized fillers.
Frequently an enhancement of FC performances is reported, but not much improvement on membrane stability in alkaline conditions.
For 1D fillers, the major work was perfromed on carbon nanotubes with various functionalization and some papers on titanate nanotubes. One-dimensional tubular nanofillers can provide ion conducting groups, construct ion conducting pathways along the interphase region and improve the anti-swelling property of the nanohybrid membrane. The carbon nanotubes are also characterized by extreme strength and flexibility; the increase in mechanical properties is evident but the increase in FC performance is modest. The use of imidazole as functionalizing molecule contributes to attainment of a high conductivity; a composite with imidazole PEEK and imidazole CNT reached a hydroxide conductivity of 135 mS cm−1 at 70 °C.
For 2D fillers, by far the largest amount of results was reported on graphene oxides; significant work was also performed on inexpensive layered double hydroxides which are intrinsically anionic conductors. A perfluorinated anion-conducting polymer with ionic-liquid-functionalized graphene nanoribbons attained a conductivity of 121 mS cm−1 at 80 °C in water. Sandwiched-porous PBI and reduced graphene oxide attained a maximum power density at 90 °C of 544 mW cm−2. The membrane GO/cellulose/PPO displayed a conductivity of ~114 mS cm−1 at 25 °C and ~215 mS cm−1 at 80 °C. PVA and exfoliated LDH(NO3) membrane exhibited a conductivity of 172.1 mS cm−1 at 80 °C. A porous-sandwich structure based on triple-cation side chain PPO and quaternary-ammonium-modified LDH exhibited a maximum power density of 267 mW cm−2 at 60 °C.
For 3D fillers, most works were reported on functionalized silica and zirconia nanoparticles. The ionic conductivity generally did not improve much, remaining at acceptable values around 10 mS cm−1 at 60 °C for most composites with silica. A maximum power density of 298 mW cm−2 at 60 °C was attained for a QPSU with triethylamine functionalized mesoporous silica. The stability in an alkaline environment was improved, but not significantly except for a fluoropolymer-based AEM containing ZrO(ClO4)2 where the measured stability seems to exceed 2000 h. As for the MOFs, we can see a good conductivity for MIL 101 with 145 mS cm−1 at 80 °C.
Still open challenges that must be overcome for a large-scale application of AEMFCs as alternative to PEMFCs are: (i) the creation of alkaline stable ionomers, (ii) the creation of efficient and stable non-PGM catalysts for HOR/ORR necessary to make a low-cost final device. Other points must be optimized including the components of the MEA and their assembly, support for the catalyst, etc. As far as composites are concerned, the improvement from a mechanical point of view is evident, but they still suffer from low alkaline stability at high temperatures. Other points of discussion are certainly: (i) clarifying the interaction at the interface between the composite membrane and the catalyst, (ii) how to improve water management, (iii) how to avoid carbonation and (iv) the use of easily reproducible test conditions for a true comparison of membrane performances. Concerning the composite membranes reported in the review, few cases are highlighted with relatively new materials such as MOFs, CDs and MXenes; we expect in the future an increase in the use of these materials by optimizing their properties as needed and making them more stable, e.g., with the formation of covalent bonds between organic and inorganic phases. Another point to explore is replacing the commercial AEM that suffers from a significant degradation both at the level of the cationic group and of the backbone, with more stable polymers such as polynorbornene, polystyrene, and ionic groups, such as n-dimethyl piperidinium-type. Even if these materials are more expensive, this disadvantage can be mitigated by a future wide commercialization of AEM capable of operating with high current densities at temperatures of 80–95 °C.

Author Contributions

Conceptualization, M.L.D.V., P.K. and R.N.; resources, R.N., E.S., P.K. and M.L.D.V.; data curation, R.N., E.S., P.K. and M.L.D.V.; writing—review and editing, R.N., E.S., P.K. and M.L.D.V.; visualization, R.N., E.S., P.K. and M.L.D.V.; supervision, P.K. and M.L.D.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Varcoe, J.R.; Slade, R.C.T. Prospects for Alkaline Anion-Exchange Membranes in Low Temperature Fuel Cells. Fuel Cells 2004, 5, 187–200. [Google Scholar] [CrossRef] [Green Version]
  2. Varcoe, J.R.; Atanassov, P.; Dekel, D.R.; Herring, A.M.; Hickner, M.A.; Kohl, P.A.; Kucernak, A.R.; Mustain, W.E.; Nijmeijer, K.; Scott, K.; et al. Anion-exchange membranes in electrochemical energy systems. Energy Environ. Sci. 2014, 7, 3135–3191. [Google Scholar] [CrossRef] [Green Version]
  3. Maurya, S.; Shin, S.-H.; Kim, Y.; Moon, S.-H. A review on recent developments of anion exchange membranes for fuel cells and redox flow batteries. RSC Adv. 2015, 5, 37206–37230. [Google Scholar] [CrossRef]
  4. Merle, G.; Wessling, M.; Nijmeijer, K. Anion exchange membranes for alkaline fuel cells: A review. J. Membr. Sci. 2011, 377, 1–35. [Google Scholar] [CrossRef]
  5. Antolini, E. Composite materials for polymer electrolyte membrane microbial fuel cells. Biosens. Bioelectron. 2015, 69, 54–70. [Google Scholar] [CrossRef]
  6. Vijayakumar, V.; Nam, S.Y. Recent advancements in applications of alkaline anion exchange membranes for polymer electrolyte fuel cells. J. Ind. Eng. Chem. 2018, 70, 70–86. [Google Scholar] [CrossRef]
  7. Cheng, J.; He, G.; Zhang, F. A mini-review on anion exchange membranes for fuel cell applications: Stability issue and addressing strategies. Int. J. Hydrogen Energy 2015, 40, 7348–7360. [Google Scholar] [CrossRef]
  8. Dekel, D.R. Review of cell performance in anion exchange membrane fuel cells. J. Power Sources 2018, 375, 158–169. [Google Scholar] [CrossRef]
  9. You, W.; Noonan, K.J.; Coates, G.W. Alkaline-stable anion exchange membranes: A review of synthetic approaches. Prog. Polym. Sci. 2019, 100, 101177. [Google Scholar] [CrossRef]
  10. Knauth, P.; Pasquini, L.; Narducci, R.; Sgreccia, E.; Becerra-Arciniegas, R.-A.; Di Vona, M. Effective ion mobility in anion exchange ionomers: Relations with hydration, porosity, tortuosity, and percolation. J. Membr. Sci. 2020, 617, 118622. [Google Scholar] [CrossRef]
  11. Omasta, T.J.; Park, A.M.; LaManna, J.M.; Zhang, Y.; Peng, X.; Wang, L.; Jacobson, D.L.; Varcoe, J.R.; Hussey, D.S.; Pivovar, B.S.; et al. Beyond catalysis and membranes: Visualizing and solving the challenge of electrode water accumulation and flooding in AEMFCs. Energy Environ. Sci. 2018, 11, 551–558. [Google Scholar] [CrossRef] [Green Version]
  12. Zheng, Y.; Omasta, T.J.; Peng, X.; Wang, L.; Varcoe, J.R.; Pivovar, B.S.; Mustain, W.E. Quantifying and elucidating the effect of CO2 on the thermodynamics, kinetics and charge transport of AEMFCs. Energy Environ. Sci. 2019, 12, 2806–2819. [Google Scholar] [CrossRef] [Green Version]
  13. Mustain, W.E.; Chatenet, M.; Page, M.; Kim, Y.S. Durability challenges of anion exchange membrane fuel cells. Energy Environ. Sci. 2020, 13, 2805–2838. [Google Scholar] [CrossRef]
  14. Chen, N.; Lee, Y.M. Anion exchange polyelectrolytes for membranes and ionomers. Prog. Polym. Sci. 2020, 113, 101345. [Google Scholar] [CrossRef]
  15. Lim, K.; Wong, C.; Wong, W.; Loh, K.; Selambakkannu, S.; Othman, N.; Yang, H. Radiation-Grafted Anion-Exchange Membrane for Fuel Cell and Electrolyzer Applications: A Mini Review. Membranes 2021, 11, 397. [Google Scholar] [CrossRef] [PubMed]
  16. Abdullah, M.; Kamarudin, S. Titanium dioxide nanotubes (TNT) in energy and environmental applications: An overview. Renew. Sustain. Energy Rev. 2017, 76, 212–225. [Google Scholar] [CrossRef]
  17. Premchand, Y.D.; Djenizian, T.; Vacandio, F.; Knauth, P. Fabrication of self-organized TiO2 nanotubes from columnar titanium thin films sputtered on semiconductor surfaces. Electrochem. Commun. 2006, 8, 1840–1844. [Google Scholar] [CrossRef]
  18. Gong, D.; Grimes, C.A.; Varghese, O.K.; Hu, W.; Singh, R.S.; Chen, Z.; Dickey, E. Titanium oxide nanotube arrays prepared by anodic oxidation. J. Mater. Res. 2001, 16, 3331–3334. [Google Scholar] [CrossRef] [Green Version]
  19. Kasuga, T.; Hiramatsu, M.; Hoson, A.; Sekino, T.; Niihara, K. Formation of Titanium Oxide Nanotube. Langmuir 1998, 14, 3160–3163. [Google Scholar] [CrossRef]
  20. Elumalai, V.; Sangeetha, D. Preparation of anion exchangeable titanate nanotubes and their effect on anion exchange membrane fuel cell. Mater. Des. 2018, 154, 63–72. [Google Scholar] [CrossRef]
  21. Elumalai, V.; Sangeetha, D. Synergic effect of ionic liquid grafted titanate nanotubes on the performance of anion exchange membrane fuel cell. J. Power Sources 2018, 412, 586–596. [Google Scholar] [CrossRef]
  22. Lu, Y.; Hu, Z.; Buregeya, I.; Pan, X.; Li, N.; Chen, S.; Liu, Y. Composite Anion Exchange Membranes from Quaternized Poly(arylene ether ketone) and Quaternized Titanate Nanotubes with Enhanced Ion Conductivity. Fuel Cells 2019, 19, 663–674. [Google Scholar] [CrossRef]
  23. Shi, B.; Li, Y.; Zhang, H.; Wu, W.; Ding, R.; Dang, J.; Wang, J. Tuning the performance of anion exchange membranes by embedding multifunctional nanotubes into a polymer matrix. J. Membr. Sci. 2016, 498, 242–253. [Google Scholar] [CrossRef]
  24. Pan, W.-H.; Lue, S.J.; Chang, C.-M.; Liu, Y.-L. Alkali doped polyvinyl alcohol/multi-walled carbon nano-tube electrolyte for direct methanol alkaline fuel cell. J. Membr. Sci. 2011, 376, 225–232. [Google Scholar] [CrossRef]
  25. Lue, S.J.; Pan, W.-H.; Chang, C.-M.; Liu, Y.-L. High-performance direct methanol alkaline fuel cells using potassium hydroxide-impregnated polyvinyl alcohol/carbon nano-tube electrolytes. J. Power Sources 2012, 202, 1–10. [Google Scholar] [CrossRef]
  26. Lo, C.-F.; Wu, J.-F.; Li, H.-Y.; Hung, W.-S.; Shih, C.-M.; Hu, C.-C.; Liu, Y.-L.; Lue, S.J. Novel polyvinyl alcohol nanocomposites containing carbon nano-tubes with Fe3O4 pendants for alkaline fuel cell applications. J. Membr. Sci. 2013, 444, 41–49. [Google Scholar] [CrossRef]
  27. Huang, C.-Y.; Lin, J.-S.; Pan, W.-H.; Shih, C.-M.; Liu, Y.-L.; Lue, S.J. Alkaline direct ethanol fuel cell performance using alkali-impregnated polyvinyl alcohol/functionalized carbon nano-tube solid electrolytes. J. Power Sources 2016, 303, 267–277. [Google Scholar] [CrossRef]
  28. Zhou, T.; Wang, M.; He, X.; Qiao, J. Poly(vinyl alcohol)/Poly(diallyldimethylammonium chloride) anion-exchange membrane modified with multiwalled carbon nanotubes for alkaline fuel cells. J. Mater. 2019, 5, 286–295. [Google Scholar] [CrossRef]
  29. Zhou, T.; He, X.; Lu, Z. Studies on a novel anion-exchange membrane based on chitosan and ionized organic compounds with multiwalled carbon nanotubes for alkaline fuel cells. J. Appl. Polym. Sci. 2018, 135, 46323. [Google Scholar] [CrossRef]
  30. Jang, S.C.; Chuang, F.S.; Tsen, W.C.; Kuo, T.W. Quaternized chitosan/functionalized carbon nanotubes composite anion exchange membranes. J. Appl. Polym. Sci. 2019, 136, 47778. [Google Scholar] [CrossRef]
  31. Jang, S.-C.; Tsen, W.-C.; Chuang, F.-S.; Gong, C. Simultaneously enhanced hydroxide conductivity and mechanical properties of quaternized chitosan/functionalized carbon nanotubes composite anion exchange membranes. Int. J. Hydrogen Energy 2019, 44, 18134–18144. [Google Scholar] [CrossRef]
  32. Vinodh, R.; Sangeetha, D. Quaternized Poly(Styrene Ethylene Butylene Poly Styrene)/Multiwalled Carbon Nanotube Composites for Alkaline Fuel Cell Applications. J. Nanosci. Nanotechnol. 2013, 13, 5522–5533. [Google Scholar] [CrossRef] [PubMed]
  33. Li, Q.; Liu, L.; Liang, S.; Dong, Q.; Jin, B.; Bai, R. Preparation and characterization of composite membranes with ionic liquid polymer-functionalized multiwalled carbon nanotubes for alkaline fuel cells. RSC Adv. 2013, 3, 13477–13485. [Google Scholar] [CrossRef]
  34. Qiu, M.; Zhang, B.; Wu, H.; Cao, L.; He, X.; Li, Y.; Li, J.; Xu, M.; Jiang, Z. Preparation of anion exchange membrane with enhanced conductivity and alkaline stability by incorporating ionic liquid modified carbon nanotubes. J. Membr. Sci. 2018, 573, 1–10. [Google Scholar] [CrossRef]
  35. Yang, P.; Zhang, B.; Wu, H.; Cao, L.; He, X.; Jiang, Z. Imidazolium-functionalized carbon nanotubes crosslinked with imidazole poly(ether ether ketone) for fabricating anion exchange membranes with high hydroxide conductivity and dimension stability. Electrochim. Acta 2019, 318, 572–580. [Google Scholar] [CrossRef]
  36. Gong, X.; Dai, Y.; Yan, X.; Wu, X.; Wang, Q.; Zhen, D.; Li, T.; Chen, W.; He, G. Electrospun imidazolium functionalized multiwalled carbon nanotube/polysulfone inorganic-organic nanofibers for reinforced anion exchange membranes. Int. J. Hydrogen Energy 2018, 43, 21547–21559. [Google Scholar] [CrossRef]
  37. Di Vona, M.L.; Casciola, M.; Donnadio, A.; Nocchetti, M.; Pasquini, L.; Narducci, R.; Knauth, P. Anionic conducting composite membranes based on aromatic polymer and layered double hydroxides. Int. J. Hydrogen Energy 2017, 42, 3197–3205. [Google Scholar] [CrossRef]
  38. Zeng, L.; Zhao, T.; Li, Y. Synthesis and characterization of crosslinked poly (vinyl alcohol)/layered double hydroxide composite polymer membranes for alkaline direct ethanol fuel cells. Int. J. Hydrogen Energy 2012, 37, 18425–18432. [Google Scholar] [CrossRef]
  39. He, X.Y.; Cao, L.; He, G.W.; Zhao, A.Q.; Mao, X.L.; Huang, T.; Li, Y.; Wu, H.; Sun, J.; Jiang, Z.Y. A highly conductive and robust anion conductor obtained via synergistic manipulation in intra- and inter-laminate of layered double hydroxide nanosheets. J. Mater. Chem. A 2018, 6, 10277–10285. [Google Scholar] [CrossRef]
  40. Hu, Y.; Tsen, W.-C.; Chuang, F.-S.; Jang, S.-C.; Zhang, B.; Zheng, G.; Wen, S.; Liu, H.; Qin, C.; Gong, C. Glycine betaine intercalated layered double hydroxide modified quaternized chitosan/polyvinyl alcohol composite membranes for alkaline direct methanol fuel cells. Carbohydr. Polym. 2018, 213, 320–328. [Google Scholar] [CrossRef]
  41. Zhao, S.; Tsen, W.-C.; Hu, F.; Zhong, F.; Liu, H.; Wen, S.; Zheng, G.; Qin, C.; Gong, C. Layered double hydroxide-coated silica nanospheres with 3D architecture-modified composite anion exchange membranes for fuel cell applications. J. Mater. Sci. 2019, 55, 2967–2983. [Google Scholar] [CrossRef]
  42. Gong, C.; Zhao, S.; Tsen, W.-C.; Hu, F.; Zhong, F.; Zhang, B.; Liu, H.; Zheng, G.; Qin, C.; Wen, S. Hierarchical layered double hydroxide coated carbon nanotube modified quaternized chitosan/polyvinyl alcohol for alkaline direct methanol fuel cells. J. Power Sources 2019, 441, 227176. [Google Scholar] [CrossRef]
  43. Liu, W.; Liang, N.; Peng, P.; Qu, R.; Chen, D.; Zhang, H. Anion-exchange membranes derived from quaternized polysulfone and exfoliated layered double hydroxide for fuel cells. J. Solid State Chem. 2017, 246, 324–328. [Google Scholar] [CrossRef]
  44. Pizzoferrato, R.; Ciotta, E.; Ferrari, I.V.; Narducci, R.; Pasquini, L.; Varone, A.; Richetta, M.; Antonaroli, S.; Braglia, M.; Knauth, P.; et al. Layered Double Hydroxides Containing an Ionic Liquid: Ionic Conductivity and Use in Composite Anion Exchange Membranes. ChemElectroChem 2018, 5, 2781–2788. [Google Scholar] [CrossRef]
  45. Chen, N.; Long, C.; Li, Y.; Wang, D.; Lu, C.; Zhu, H.; Yu, J. Three-Decker Strategy Based on Multifunctional Layered Double Hydroxide to Realize High-Performance Hydroxide Exchange Membranes for Fuel Cell Applications. ACS Appl. Mater. Interfaces 2018, 10, 18246–18256. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, N.J.; Long, C.; Li, Y.X.; Wang, D.; Zhu, H. High-performance layered double hydroxide/poly(2,6-dimethyl-1,4-phenyleneoxide) membrane with porous sandwich structure for anion exchange membrane fuel cell applications. J. Membr. Sci. 2018, 552, 51–60. [Google Scholar] [CrossRef]
  47. Zhu, H.; Li, R.; Chen, N.; Wang, F.; Wang, Z.; Han, K. Electrorheological effect induced quaternized poly(2,6-dimethyl phenylene oxide)-layered double hydroxide composite membranes for anion exchange membrane fuel cells. RSC Adv. 2016, 6, 85486–85494. [Google Scholar] [CrossRef]
  48. Wang, D.; Chen, N.; Long, C.; Lu, C.; Li, Y.; Zhu, H.; Wang, F. Electric-field-aligned functionalized-layered double hydroxide/polyphenyl ether composite membrane for ion transport. Int. J. Hydrogen Energy 2019, 44, 13852–13863. [Google Scholar] [CrossRef]
  49. Pasquini, L.; Becerra-Arciniegas, R.; Narducci, R.; Sgreccia, E.; Gressel, V.; Di Vona, M.; Knauth, P. Properties and Alkaline Stability of Composite Anion Conducting Ionomers Based on Poly(phenylene oxide) Grafted with DABCO and Mg/Al Lamellar Double Hydroxide. ChemElectroChem 2020, 7, 2917–2924. [Google Scholar] [CrossRef]
  50. Sailaja, G.S.; Zhang, P.; Anilkumar, G.M.; Yamaguchi, T. Aniosotropically Organized LDH on PVDF: A Geometrically Templated Electrospun Substrate for Advanced Anion Conducting Membranes. ACS Appl. Mater. Interfaces 2015, 7, 6397–6401. [Google Scholar] [CrossRef]
  51. Sinha, A.; Dhanjai; Zhao, H.; Huang, Y.; Lu, X.; Chen, J.; Jain, R. MXene: An emerging material for sensing and biosensing. TrAC Trends Anal. Chem. 2018, 105, 424–435. [Google Scholar] [CrossRef]
  52. Zhang, X.; Fan, C.; Yao, N.; Zhang, P.; Hong, T.; Xu, C.; Cheng, J. Quaternary Ti3C2Tx enhanced ionic conduction in quaternized polysulfone membrane for alkaline anion exchange membrane fuel cells. J. Membr. Sci. 2018, 563, 882–887. [Google Scholar] [CrossRef]
  53. Wang, L.; Shi, B. Hydroxide Conduction Enhancement of Chitosan Membranes by Functionalized MXene. Materials 2018, 11, 2335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Ye, Y.-S.; Cheng, M.-Y.; Xie, X.-L.; Rick, J.; Huang, Y.-J.; Chang, F.-C.; Hwang, B.-J. Alkali doped polyvinyl alcohol/graphene electrolyte for direct methanol alkaline fuel cells. J. Power Sources 2013, 239, 424–432. [Google Scholar] [CrossRef]
  55. Yang, J.-M.; Wang, S.-A. Preparation of graphene-based poly(vinyl alcohol)/chitosan nanocomposites membrane for alkaline solid electrolytes membrane. J. Membr. Sci. 2015, 477, 49–57. [Google Scholar] [CrossRef]
  56. Sharma, P.P.; Gahlot, S.; Bhil, B.M.; Gupta, H.; Kulshrestha, V. An environmentally friendly process for the synthesis of an fGO modified anion exchange membrane for electro-membrane applications. RSC Adv. 2015, 5, 38712–38721. [Google Scholar] [CrossRef]
  57. Zakaria, Z.; Kamarudin, S.K.; Timmiati, S.N. Influence of Graphene Oxide on the Ethanol Permeability and Ionic Conductivity of QPVA-Based Membrane in Passive Alkaline Direct Ethanol Fuel Cells. Nanoscale Res. Lett. 2019, 14, 1–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Ma, W.-T.; Kumar, S.R.; Hsu, C.-T.; Shih, C.-M.; Tsai, S.-W.; Yang, C.-C.; Liu, Y.-L.; Lue, S.J. Magnetic field-assisted alignment of graphene oxide nanosheets in a polymer matrix to enhance ionic conduction. J. Membr. Sci. 2018, 563, 259–269. [Google Scholar] [CrossRef]
  59. Lin, J.-S.; Kumar, S.R.; Ma, W.-T.; Shih, C.-M.; Teng, L.-W.; Yang, C.-C.; Lue, S.J. Gradiently distributed iron oxide@graphene oxide nanofillers in quaternized polyvinyl alcohol composite to enhance alkaline fuel cell power density. J. Membr. Sci. 2017, 543, 28–39. [Google Scholar] [CrossRef]
  60. Liu, L.; Tong, C.; He, Y.; Zhao, Y.; Lü, C. Enhanced properties of quaternized graphenes reinforced polysulfone based composite anion exchange membranes for alkaline fuel cell. J. Membr. Sci. 2015, 487, 99–108. [Google Scholar] [CrossRef]
  61. Luo, Z.; Gong, Y.; Liao, X.; Pan, Y.; Zhang, H. Nanocomposite membranes modified by graphene-based materials for anion exchange membrane fuel cells. RSC Adv. 2016, 6, 13618–13625. [Google Scholar] [CrossRef]
  62. Liu, J.; Qu, R.; Peng, P.; Liu, W.; Chen, D.; Zhang, H.; Liu, X. Covalently functionalized graphene oxide and quaternized polysulfone nanocomposite membranes for fuel cells. RSC Adv. 2016, 6, 71305–71310. [Google Scholar] [CrossRef]
  63. Liu, Y.; Dai, J.; Zhang, K.; Ma, L.; Qaisrani, N.A.; Zhang, F.; He, G. Hybrid anion exchange membrane of hydroxyl-modified polysulfone incorporating guanidinium-functionalized graphene oxide. Ionics 2017, 23, 3085–3096. [Google Scholar] [CrossRef]
  64. Hu, B.; Miao, L.; Zhao, Y.; Lü, C. Azide-assisted crosslinked quaternized polysulfone with reduced graphene oxide for highly stable anion exchange membranes. J. Membr. Sci. 2017, 530, 84–94. [Google Scholar] [CrossRef]
  65. Msomi, P.F.; Nonjola, P.; Ndungu, P.G.; Ramontja, J. Quaternized poly (2.6 dimethyl—1.4 phenylene oxide)/Polysulfone anion exchange membrane reinforced with graphene oxide for methanol alkaline fuel cell application. J. Polym. Res. 2018, 25, 143. [Google Scholar] [CrossRef]
  66. Mao, X.; Li, Z.; He, G.; Li, Z.; Zhao, J.; Zhang, Y.; Jiang, Z. Enhancing hydroxide conductivity of anion exchange membrane via incorporating densely imidazolium functionalized graphene oxide. Solid State Ionics 2019, 333, 83–92. [Google Scholar] [CrossRef]
  67. Hu, B.; Liu, L.; Zhao, Y.; Lü, C. A facile construction of quaternized polymer brush-grafted graphene modified polysulfone based composite anion exchange membranes with enhanced performance. RSC Adv. 2016, 6, 51057–51067. [Google Scholar] [CrossRef]
  68. Miao, L.; Bai, Y.; Yuan, Y.; Lü, C. Mussel-inspired strategy towards functionalized reduced graphene oxide-crosslinked polysulfone-based anion exchange membranes with enhanced properties. Int. J. Hydrogen Energy 2018, 43, 17461–17474. [Google Scholar] [CrossRef]
  69. Bai, Y.; Yuan, Y.; Miao, L.; Lü, C. Functionalized rGO as covalent crosslinkers for constructing chemically stable polysulfone-based anion exchange membranes with enhanced ion conductivity. J. Membr. Sci. 2018, 570–571, 481–493. [Google Scholar] [CrossRef]
  70. Bai, Y.; Yuan, Y.; Yang, Y.; Lü, C. A facile fabrication of functionalized rGO crosslinked chemically stable polysulfone-based anion exchange membranes with enhanced performance. Int. J. Hydrogen Energy 2019, 44, 6618–6630. [Google Scholar] [CrossRef]
  71. Shukla, G.; Shahi, V.K. Amine functionalized graphene oxide containing C16 chain grafted with poly(ether sulfone) by DABCO coupling: Anion exchange membrane for vanadium redox flow battery. J. Membr. Sci. 2019, 575, 109–117. [Google Scholar] [CrossRef]
  72. Yadav, V.; Sharma, P.P.; Kulshrestha, V. Facile synthesis of Br-PPO/f GO based polymer electrolyte membranes for electrochemical applications. Int. J. Hydrogen Energy 2017, 42, 26511–26521. [Google Scholar] [CrossRef]
  73. Das, G.; Park, B.J.; Kim, J.; Kang, D.; Yoon, H.H. Quaternized cellulose and graphene oxide crosslinked polyphenylene oxide based anion exchange membrane. Sci. Rep. 2019, 9, 9572. [Google Scholar] [CrossRef] [PubMed]
  74. Yang, Q.; Lin, C.X.; Liu, F.H.; Li, L.; Zhang, Q.G.; Zhu, A.M.; Liu, Q.L. Poly (2,6-dimethyl-1,4-phenylene oxide)/ionic liquid functionalized graphene oxide anion exchange membranes for fuel cells. J. Membr. Sci. 2018, 552, 367–376. [Google Scholar] [CrossRef]
  75. Ion-Ebrasu, D.; Pollet, B.G.; Caprarescu, S.; Chitu, A.; Trusca, R.; Niculescu, V.; Gabor, R.; Carcadea, E.; Varlam, M.; Vasile, B.S. Graphene inclusion effect on anion-exchange membranes properties for alkaline water electrolyzers. Int. J. Hydrogen Energy 2020, 45, 17057–17066. [Google Scholar] [CrossRef]
  76. Zhao, Y.; Tang, K.; Ruan, H.; Xue, L.; Van der Bruggen, B.; Gao, C.; Shen, J. Sulfonated reduced graphene oxide modification layers to improve monovalent anions selectivity and controllable resistance of anion exchange membrane. J. Membr. Sci. 2017, 536, 167–175. [Google Scholar] [CrossRef]
  77. Zeng, L.; Zhao, T.S.; An, L.; Zhao, G.; Yan, X.H. A high-performance sandwiched-porous polybenzimidazole membrane with enhanced alkaline retention for anion exchange membrane fuel cells. Energy Environ. Sci. 2015, 8, 2768–2774. [Google Scholar] [CrossRef]
  78. Yu, B.-C.; Wang, Y.-C.; Lu, H.-C.; Lin, H.-L.; Shih, C.-M.; Kumar, S.R.; Lue, S.J. Hydroxide-ion selective electrolytes based on a polybenzimidazole/graphene oxide composite membrane. Energy 2017, 134, 802–812. [Google Scholar] [CrossRef]
  79. Chang, W.T.; Chao, Y.H.; Li, C.W.; Lin, K.L.; Wang, J.J.; Kumar, S.R.; Lue, S.J. Graphene oxide synthesis using microwave-assisted vs. modified Hummer’s methods: Efficient fillers for improved ionic conductivity and suppressed methanol permeability in alkaline methanol fuel cell electrolytes. J. Power Sources 2019, 414, 86–95. [Google Scholar] [CrossRef]
  80. Chu, J.Y.; Lee, K.H.; Kim, A.R.; Yoo, D.J. Graphene-mediated organic-inorganic composites with improved hydroxide conductivity and outstanding alkaline stability for anion exchange membranes. Compos. Part B Eng. 2018, 164, 324–332. [Google Scholar] [CrossRef]
  81. Zhang, S.; Zhu, X.; Jin, C.; Zhang, Z. Cyclodextrin edge functionalized graphene oxide modified poly(phthalazinone ether ketone) composite membrane with enhanced properties for anion exchange membrane. Solid State Ion. 2018, 320, 360–368. [Google Scholar] [CrossRef]
  82. Li, J.; Zhang, B.; Wu, H.; Cao, L.; He, X.; Li, Y.; Xu, M.; Jiang, Z. Incorporating imidazolium-functionalized graphene oxide into imidazolium-functionalized poly(ether ether ketone) for enhanced hydroxide conductivity. J. Membr. Sci. 2018, 565, 233–240. [Google Scholar] [CrossRef]
  83. Chu, J.Y.; Lee, K.H.; Kim, A.R.; Yoo, D.J. Improved electrochemical performance of composite anion exchange membranes for fuel cells through cross linking of the polymer chain with functionalized graphene oxide. J. Membr. Sci. 2020, 611, 118385. [Google Scholar] [CrossRef]
  84. Liu, X.; Chen, X.; Hu, Y.; Gong, T.; Li, H.; Zhang, Y. Ionic-Liquid-Functionalized Graphene Nanoribbons for Anion Exchange Membrane Fuel Cells. J. Electrochem. Soc. 2017, 164, F433–F440. [Google Scholar] [CrossRef]
  85. Dai, P.; Mo, Z.-H.; Xu, R.-W.; Zhang, S.; Lin, X.; Lin, W.-F.; Wu, Y.-X. Development of a cross-linked quaternized poly(styrene-b-isobutylene-b-styrene)/graphene oxide composite anion exchange membrane for direct alkaline methanol fuel cell application. RSC Adv. 2016, 6, 52122–52130. [Google Scholar] [CrossRef] [Green Version]
  86. Mo, Z.-H.; Yang, R.; Hong, S.; Wu, Y.-X. In-situ preparation of cross-linked hybrid anion exchange membrane of quaternized poly (styrene-b-isobutylene-b-styrene) covalently bonded with graphene. Int. J. Hydrogen Energy 2018, 43, 1790–1804. [Google Scholar] [CrossRef]
  87. Ouadah, A.; Luo, T.; Wang, J.; Gao, S.; Wang, X.; Zhang, X.; Fang, Z.; Wu, Z.; Wang, J.; Zhu, C. Imidazolium-grafted graphene oxide via free radical polymerization: An efficient and simple method for an interpenetrating polymer network as electrolyte membrane. Compos. Sci. Technol. 2018, 164, 204–213. [Google Scholar] [CrossRef]
  88. Bayer, T.; Cunning, B.; Selyanchyn, R.; Daio, T.; Nishihara, M.; Fujikawa, S.; Sasaki, K.; Lyth, S. Alkaline anion exchange membranes based on KOH-treated multilayer graphene oxide. J. Membr. Sci. 2016, 508, 51–61. [Google Scholar] [CrossRef] [Green Version]
  89. Ingabire, P.B.; Pan, X.; Haragirimana, A.; Li, N.; Hu, Z.; Chen, S. Enhanced conduction capability of nanocomposite membrane of quaternized poly (arylene ether sulfone)s covalently bonded with graphitic carbon nitride nanosheets for fuel cells. React. Funct. Polym. 2019, 144, 104260. [Google Scholar] [CrossRef]
  90. Lu, Y.; Pan, X.; Li, N.; Hu, Z.; Chen, S. Improved performance of quaternized poly(arylene ether ketone)s/graphitic carbon nitride nanosheets composite anion exchange membrane for fuel cell applications. Appl. Surf. Sci. 2019, 503, 144071. [Google Scholar] [CrossRef]
  91. Rathod, N.H.; Yadav, V.; Rajput, A.; Sharma, J.; Shukla, D.; Kulshrestha, V. New class of composite anion exchange membranes based on Quaternized poly (phenylene oxide) and functionalized boron nitride. Colloid Interface Sci. Commun. 2020, 36, 100265. [Google Scholar] [CrossRef]
  92. Li, C.; Sun, G.; Ren, S.; Liu, J.; Wang, Q.; Wu, Z.; Sun, H.; Jin, W. Casting Nafion–sulfonated organosilica nano-composite membranes used in direct methanol fuel cells. J. Membr. Sci. 2006, 272, 50–57. [Google Scholar] [CrossRef]
  93. Adjemian, K.T.; Lee, S.J.; Srinivasan, S.; Benziger, J.; Bocarsly, A.B. Silicon oxide Nafion composite membranes for proton-exchange membrane fuel cell operation at 80–140 °C. J. Electrochem. Soc. 2002, 149, A256–A261. [Google Scholar] [CrossRef]
  94. Sgreccia, E.; Narducci, R.; Knauth, P.; Di Vona, M. Silica Containing Composite Anion Exchange Membranes by Sol–Gel Synthesis: A Short Review. Polymers 2021, 13, 1874. [Google Scholar] [CrossRef]
  95. Zuo, X.; Yu, S.; Xu, X.; Xu, J.; Bao, R.; Yan, X. New PVDF organic–inorganic membranes: The effect of SiO2 nanoparticles content on the transport performance of anion-exchange membranes. J. Membr. Sci. 2009, 340, 206–213. [Google Scholar] [CrossRef]
  96. Liu, G.; Tsen, W.-C.; Jang, S.-C.; Hu, F.; Zhong, F.; Zhang, B.; Wang, J.; Liu, H.; Wang, G.; Wen, S.; et al. Composite membranes from quaternized chitosan reinforced with surface-functionalized PVDF electrospun nanofibers for alkaline direct methanol fuel cells. J. Membr. Sci. 2020, 611, 118242. [Google Scholar] [CrossRef]
  97. Wang, E.D.; Zhao, T.S.; Yang, W.W. Poly (vinyl alcohol)/3-(trimethylammonium) propyl-functionalized silica hybrid membranes for alkaline direct ethanol fuel cells. Int. J. Hydrogen Energy 2010, 35, 2183–2189. [Google Scholar] [CrossRef]
  98. Yang, C.-C.; Chiu, S.-S.; Kuo, S.-C.; Liou, T.-H. Fabrication of anion-exchange composite membranes for alkaline direct methanol fuel cells. J. Power Sources 2012, 199, 37–45. [Google Scholar] [CrossRef]
  99. Kumar, S.R.; Juan, C.-H.; Liao, G.-M.; Lin, J.-S.; Yang, C.-C.; Ma, W.-T.; You, J.-H.; Lue, S.J. Fumed Silica Nanoparticles Incorporated in Quaternized Poly(Vinyl Alcohol) Nanocomposite Membrane for Enhanced Power Densities in Direct Alcohol Alkaline Fuel Cells. Energies 2015, 9, 15. [Google Scholar] [CrossRef] [Green Version]
  100. Lu, Y.; Armentrout, A.A.; Li, J.; Tekinalp, H.L.; Nanda, J.; Ozcan, S. A cellulose nanocrystal-based composite electrolyte with superior dimensional stability for alkaline fuel cell membranes. J. Mater. Chem. A 2015, 3, 13350–13356. [Google Scholar] [CrossRef]
  101. García-Cruz, L.; Casado-Coterillo, C.; Iniesta, J.; Montiel, V.; Irabien, A. Chitosan:poly (vinyl) alcohol composite alkaline membrane incorporating organic ionomers and layered silicate materials into a PEM electrochemical reactor. J. Membr. Sci. 2016, 498, 395–407. [Google Scholar] [CrossRef] [Green Version]
  102. Pan, J.F.; He, Y.B.; Wu, L.; Jiang, C.X.; Wu, B.; Mondal, A.N.; Cheng, C.L.; Xu, T.W. Anion exchange membranes from hot-pressed electrospun QPPO-SiO2 hybrid nanofibers for acid recovery. J. Membr. Sci. 2015, 480, 115–121. [Google Scholar] [CrossRef]
  103. Moghadasi, M.; Mortaheb, H.R. Incorporating functionalized silica nanoparticles in polyethersulfone-based anion exchange nanocomposite membranes. J. Appl. Polym. Sci. 2016, 134. [Google Scholar] [CrossRef]
  104. Chen, N.; Long, C.; Li, Y.; Wang, D.; Zhu, H. A hamburger-structure imidazolium-modified silica/polyphenyl ether composite membrane with enhancing comprehensive performance for anion exchange membrane applications. Electrochimica Acta 2018, 268, 295–303. [Google Scholar] [CrossRef]
  105. Vijayakumar, E.; Sangeetha, D. A quaternized mesoporous silica/polysulfone composite membrane for an efficient alkaline fuel cell application. RSC Adv. 2015, 5, 42828–42835. [Google Scholar] [CrossRef]
  106. Liu, L.; Tong, C.; He, Y.; Zhao, Y.; Hu, B.; Lü, C. Novel quaternized mesoporous silica nanoparticle modified polysulfone-based composite anion exchange membranes for alkaline fuel cells. RSC Adv. 2015, 5, 43381–43390. [Google Scholar] [CrossRef]
  107. Elumalai, V.; Dharmalingam, S. Synthesis characterization and performance evaluation of ionic liquid immobilized SBA-15/quaternised polysulfone composite membrane for alkaline fuel cell. Microporous Mesoporous Mater. 2016, 236, 260–268. [Google Scholar] [CrossRef]
  108. Liao, X.; Ren, L.; Chen, D.; Liu, X.; Zhang, H. Nanocomposite membranes based on quaternized polysulfone and functionalized montmorillonite for anion-exchange membranes. J. Power Sources 2015, 286, 258–263. [Google Scholar] [CrossRef]
  109. Li, Y.; Li, M.; Zhou, S.; Xue, A.; Zhang, Y.; Zhao, Y.; Zhong, J.; Zhang, Q.; Yang, D. Enhancement of hydroxide conductivity by incorporating nanofiber-like palygorskite into quaternized polysulfone as anion exchange membranes. Appl. Clay Sci. 2020, 195, 105702. [Google Scholar] [CrossRef]
  110. Yun, S.; Parrondo, J.; Ramani, V. Composite anion exchange membranes based on quaternized cardo-poly(etherketone) and quaternized inorganic fillers for vanadium redox flow battery applications. Int. J. Hydrogen Energy 2016, 41, 10766–10775. [Google Scholar] [CrossRef] [Green Version]
  111. Yang, C.C.; Chiu, S.J.; Chien, W.C.; Chiu, S.S. Quaternized poly(vinyl alcohol)/alumina composite polymer membranes for alkaline direct methanol fuel cells. J. Power Sources 2010, 195, 2212–2219. [Google Scholar] [CrossRef]
  112. Pérez-Prior, M.T.; García-García, T.; Varez, A.; Levenfeld, B. Preparation and characterization of ammonium-functionalized polysulfone/Al2O3 composite membranes. J. Mater. Sci. 2015, 50, 5893–5903. [Google Scholar] [CrossRef]
  113. Vinodh, R.; Purushothaman, M.; Sangeetha, D. Novel quaternized polysulfone/ZrO2 composite membranes for solid alkaline fuel cell applications. Int. J. Hydrogen Energy 2011, 36, 7291–7302. [Google Scholar] [CrossRef]
  114. Rambabu, K.; Bharath, G.; Arangadi, A.F.; Velu, S.; Fawzi, B.; Show, P.L. ZrO2 incorporated polysulfone anion exchange membranes for fuel cell applications. Int. J. Hydrogen Energy 2020, 45, 29668–29680. [Google Scholar] [CrossRef]
  115. Li, X.; Yu, Y.; Meng, Y. Novel Quaternized Poly(arylene ether sulfone)/Nano-ZrO2 Composite Anion Exchange Membranes for Alkaline Fuel Cells. ACS Appl. Mater. Interfaces 2013, 5, 1414–1422. [Google Scholar] [CrossRef]
  116. Li, X.H.; Tao, J.X.; Nie, G.H.; Wang, L.C.; Li, L.H.; Liao, S.J. Cross-linked multiblock copoly(arylene ether sulfone) ionomer/nano-ZrO2 composite anion exchange membranes for alkaline fuel cells. RSC Adv. 2014, 4, 41398–41410. [Google Scholar] [CrossRef]
  117. Xu, S.; Jiang, R.; Jiang, S.; Gao, Y. Communication—Anion-Conductive Perfluoroheteroaromatic Composite Membranes: High Chemical Stability under Strong Alkaline Conditions. J. Electrochem. Soc. 2016, 163, F688–F690. [Google Scholar] [CrossRef]
  118. Vinodh, R.; Sangeetha, D. Comparative study of composite membranes from nano-metal-oxide-incorporated polymer electrolytes for direct methanol alkaline membrane fuel cells. J. Appl. Polym. Sci. 2012, 128, 1930–1938. [Google Scholar] [CrossRef]
  119. Nonjola, P.T.; Mathe, M.K.; Modibedi, R.M. Chemical modification of polysulfone: Composite anionic exchange membrane with TiO2 nano-particles. Int. J. Hydrogen Energy 2013, 38, 5115–5121. [Google Scholar] [CrossRef]
  120. Xie, F.; Gao, X.; Hao, J.; Yu, H.; Shao, Z.; Yi, B. Preparation and properties of amorphous TiO2 modified anion exchange membrane by impregnation-hydrolysis method. React. Funct. Polym. 2019, 144, 104348. [Google Scholar] [CrossRef]
  121. Msomi, P.; Nonjola, P.; Ndungu, P.; Ramontja, J. Poly (2, 6-dimethyl-1, 4-phenylene)/polysulfone anion exchange membrane blended with TiO2 with improved water uptake for alkaline fuel cell application. Int. J. Hydrogen Energy 2020, 45, 29465–29476. [Google Scholar] [CrossRef]
  122. Derbali, Z.; Fahs, A.; Chailan, J.-F.; Ferrari, I.; Di Vona, M.; Knauth, P. Composite anion exchange membranes with functionalized hydrophilic or hydrophobic titanium dioxide. Int. J. Hydrogen Energy 2017, 42, 19178–19189. [Google Scholar] [CrossRef]
  123. Chu, Y.; Chen, Y.; Chen, N.; Wang, F.; Zhu, H. A new method for improving the ion conductivity of anion exchange membranes by using TiO2 nanoparticles coated with ionic liquid. RSC Adv. 2016, 6, 96768–96777. [Google Scholar] [CrossRef]
  124. Chen, Y.; Li, Z.; Chen, N.; Li, R.; Zhang, Y.; Li, K.; Wang, F.; Zhu, H. A new method for improving the conductivity of alkaline membrane by incorporating TiO2-ionic liquid composite particles. Electrochim. Acta 2017, 255, 335–346. [Google Scholar] [CrossRef]
  125. Tao, H.-C.; Sun, X.-N.; Xiong, Y. A novel hybrid anion exchange membrane for high performance microbial fuel cells. RSC Adv. 2014, 5, 4659–4663. [Google Scholar] [CrossRef]
  126. Moly, P.P.; Jeena, C.B.; Elsa, P.J.; Ambily, K.J.; Joy, V.T. High performance polyvinyl alcohol/calcium titanate nanocomposite anion-exchange membranes as separators in redox flow batteries. Polym. Bull. 2018, 75, 4409–4428. [Google Scholar] [CrossRef]
  127. Hibino, T.; Kobayashi, K. An intermediate-temperature alkaline fuel cell using an Sn0.92Sb0.08P2O7-based hydroxide-ion-conducting electrolyte and electrodes. J. Mater. Chem. A 2012, 1, 1134–1140. [Google Scholar] [CrossRef]
  128. Qin, H.; Hu, Y.; Zhu, C.; Chu, W.; Sheng, H.; Dong, Z.; He, Y.; Wang, J.; Li, A.; Chi, H.; et al. Functionalization of polyvinyl alcohol composite membrane by CoOOH for direct borohydride fuel cells. Electrochem. Commun. 2017, 77, 1–4. [Google Scholar] [CrossRef]
  129. Jiang, X.C.; Sun, Y.J.; Zhang, H.X.; Hou, L.X. Preparation and characterization of quaternized poly(vinyl alcohol)/chitosan/MoS2 composite anion exchange membranes with high selectivity. Carbohydr. Polym. 2018, 180, 96–103. [Google Scholar] [CrossRef]
  130. Msomi, P.F.; Nonjola, P.; Ndungu, P.; Ramonjta, J. Quaternized poly(2.6 dimethyl-1.4 phenylene oxide)/polysulfone blend composite membrane doped with ZnO-nanoparticles for alkaline fuel cells. J. Appl. Polym. Sci. 2017, 135, 45959. [Google Scholar] [CrossRef]
  131. Li, J.; Wang, S.; Liu, F.; Chen, H.; Wang, X.; Mao, T.; Wang, D.; Liu, G.; Wang, Z. Flame-retardant AEMs based on organic-inorganic composite polybenzimidazole membranes with enhanced hydroxide conductivity. J. Membr. Sci. 2019, 591, 117306. [Google Scholar] [CrossRef]
  132. Escorihuela, J.; Narducci, R.; Compañ, V.; Costantino, F. Proton Conductivity of Composite Polyelectrolyte Membranes with Metal-Organic Frameworks for Fuel Cell Applications. Adv. Mater. Interfaces 2018, 6, 1801146. [Google Scholar] [CrossRef]
  133. Gao, L.; Li, C.-Y.V.; Chan, K.-Y.; Chen, Z.-N. Metal–Organic Framework Threaded with Aminated Polymer Formed in Situ for Fast and Reversible Ion Exchange. J. Am. Chem. Soc. 2014, 136, 7209–7212. [Google Scholar] [CrossRef] [PubMed]
  134. Liu, C.; Zhang, G.; Zhao, C.; Li, X.; Li, M.; Na, H. MOFs synthesized by the ionothermal method addressing the leaching problem of IL–polymer composite membranes. Chem. Commun. 2014, 50, 14121–14124. [Google Scholar] [CrossRef]
  135. Liu, C.; Feng, S.; Zhuang, Z.; Qi, D.; Li, G.; Zhao, C.; Li, X.; Na, H. Towards basic ionic liquid-based hybrid membranes as hydroxide-conducting electrolytes under low humidity conditions. Chem. Commun. 2015, 51, 12629–12632. [Google Scholar] [CrossRef] [PubMed]
  136. Hsu, P.-Y.; Hu, T.-Y.; Kumar, S.R.; Chang, C.-H.; Wu, K.C.-W.; Tung, K.-L.; Lue, S.J. Highly Zeolite-Loaded Polyvinyl Alcohol Composite Membranes for Alkaline Fuel-Cell Electrolytes. Polymers 2018, 10, 102. [Google Scholar] [CrossRef] [Green Version]
  137. Wu, B.; Ge, L.; Yu, D.; Hou, L.; Li, Q.; Yang, Z.; Xu, T. Cationic metal–organic framework porous membranes with high hydroxide conductivity and alkaline resistance for fuel cells. J. Mater. Chem. A 2016, 4, 14545–14549. [Google Scholar] [CrossRef]
  138. He, X.; Gang, M.; Li, Z.; He, G.; Yin, Y.; Cao, L.; Zhang, B.; Wu, H.; Jiang, Z. Highly conductive and robust composite anion exchange membranes by incorporating quaternized MIL-101(Cr). Sci. Bull. 2017, 62, 266–276. [Google Scholar] [CrossRef] [Green Version]
  139. Yuan, Y.; Hu, B.; Tong, C.; Bai, Y.; Lü, C. Novel quaternized carbon dots modified polysulfone-based anion exchange membranes with improved performance. Int. J. Hydrogen Energy 2019, 44, 22181–22193. [Google Scholar] [CrossRef]
  140. Yadav, V.; Raj, S.K.; Rathod, N.H.; Kulshrestha, V. Polysulfone/graphene quantum dots composite anion exchange membrane for acid recovery by diffusion dialysis. J. Membr. Sci. 2020, 611, 118331. [Google Scholar] [CrossRef]
Figure 1. (a) Number of publications and (b) number of citations on the topic “composite anion exchange membranes” (from ISI Web of Science 2021).
Figure 1. (a) Number of publications and (b) number of citations on the topic “composite anion exchange membranes” (from ISI Web of Science 2021).
Polymers 13 03887 g001
Figure 2. TEM images of (a) halloysite nanotubes and (b) quaternized halloysite nanotubes. Reproduced with permission from Ref. [23].
Figure 2. TEM images of (a) halloysite nanotubes and (b) quaternized halloysite nanotubes. Reproduced with permission from Ref. [23].
Polymers 13 03887 g002
Figure 3. Effect of MWCNT addition in PVA on DMAFC performance at 30 °C (anode: 2 M MeOH in 6 M KOH, flow rate of 5 mL min−1; cathode: humidified O2, flow rate of 100 mL min−1). Reproduced with permission from Ref. [24].
Figure 3. Effect of MWCNT addition in PVA on DMAFC performance at 30 °C (anode: 2 M MeOH in 6 M KOH, flow rate of 5 mL min−1; cathode: humidified O2, flow rate of 100 mL min−1). Reproduced with permission from Ref. [24].
Polymers 13 03887 g003
Figure 4. Synthesis of ionic liquids functionalized CNTs (IL-(M;B)@CNT). Reproduced with permission from Ref. [34].
Figure 4. Synthesis of ionic liquids functionalized CNTs (IL-(M;B)@CNT). Reproduced with permission from Ref. [34].
Polymers 13 03887 g004
Figure 5. Preparation of LDH coated carbon nanotubes (LDH@CNTs). Reproduced with permission from Ref. [42].
Figure 5. Preparation of LDH coated carbon nanotubes (LDH@CNTs). Reproduced with permission from Ref. [42].
Polymers 13 03887 g005
Figure 6. Conductivity measurements at 60 °C as a function of RH% of membranes after treatment at 25 °C in 2 M KOH for 24 h. Reproduced with permission from Ref. [37].
Figure 6. Conductivity measurements at 60 °C as a function of RH% of membranes after treatment at 25 °C in 2 M KOH for 24 h. Reproduced with permission from Ref. [37].
Polymers 13 03887 g006
Figure 7. Hydroxide conductivity variations vs. time of TC-PPO and hybrid membranes with ASU-LDH at 30 °C. Reprinted with permission from Ref. [45]. Copyright (2018) American Chemical Society.
Figure 7. Hydroxide conductivity variations vs. time of TC-PPO and hybrid membranes with ASU-LDH at 30 °C. Reprinted with permission from Ref. [45]. Copyright (2018) American Chemical Society.
Polymers 13 03887 g007
Figure 8. Comparison between conventional and novel MEAs. (a) Polarization and power–density curves; (b) constant current discharge curves (O2, current density 700 mA cm−2). Reproduced from Ref. [77] with permission from The Royal Society of Chemistry.
Figure 8. Comparison between conventional and novel MEAs. (a) Polarization and power–density curves; (b) constant current discharge curves (O2, current density 700 mA cm−2). Reproduced from Ref. [77] with permission from The Royal Society of Chemistry.
Polymers 13 03887 g008
Figure 9. Modification process of GO via free radical polymerization-grafting with a b-VIB group. Reproduced with permission from Ref. [87].
Figure 9. Modification process of GO via free radical polymerization-grafting with a b-VIB group. Reproduced with permission from Ref. [87].
Polymers 13 03887 g009
Figure 10. SEM images (A) and fiber diameter distribution (B) of QPPO–SiO2; optical photographs of nanofiber mats (C) and membrane (D) after hot-press. Reproduced with permission from Ref. [102].
Figure 10. SEM images (A) and fiber diameter distribution (B) of QPPO–SiO2; optical photographs of nanofiber mats (C) and membrane (D) after hot-press. Reproduced with permission from Ref. [102].
Polymers 13 03887 g010
Figure 11. Effect of exposure to (1.5 M VO2+ + 3 M H2SO4) at 30 °C on sulfate ion conductivity of QPEK-C and QPEK-C/10–40 wt% TMSP-TMA+ composite membranes. Reproduced with permission from Ref. [110].
Figure 11. Effect of exposure to (1.5 M VO2+ + 3 M H2SO4) at 30 °C on sulfate ion conductivity of QPEK-C and QPEK-C/10–40 wt% TMSP-TMA+ composite membranes. Reproduced with permission from Ref. [110].
Polymers 13 03887 g011
Figure 12. DMFC curves of QPVA/10 wt% Al2O3 composite membrane with various fuels (4 M KOH + x M CH3OH) at 25 °C and in ambient air. Reproduced with permission from Ref. [111].
Figure 12. DMFC curves of QPVA/10 wt% Al2O3 composite membrane with various fuels (4 M KOH + x M CH3OH) at 25 °C and in ambient air. Reproduced with permission from Ref. [111].
Polymers 13 03887 g012
Figure 13. DBFCs performances for CoOOH functionalized and pristine membranes. V (a) and power density (b) as a function of discharge current density at 30 °C. Reproduced with permission from Ref. [128].
Figure 13. DBFCs performances for CoOOH functionalized and pristine membranes. V (a) and power density (b) as a function of discharge current density at 30 °C. Reproduced with permission from Ref. [128].
Polymers 13 03887 g013
Figure 14. Synthesis of ion-exchange polymer (PVBTAH) inside the porous network of ZIF-8. Reprinted with permission from Ref. [133]. Copyright (2014) American Chemical Society.
Figure 14. Synthesis of ion-exchange polymer (PVBTAH) inside the porous network of ZIF-8. Reprinted with permission from Ref. [133]. Copyright (2014) American Chemical Society.
Polymers 13 03887 g014
Figure 15. Synthesis of CDs and QCDs. Reproduced with permission from Ref. [139].
Figure 15. Synthesis of CDs and QCDs. Reproduced with permission from Ref. [139].
Polymers 13 03887 g015
Table 1. Type of polymer and 1D filler in composite AEM with TNT and CNT.
Table 1. Type of polymer and 1D filler in composite AEM with TNT and CNT.
Polymer1D FillerRemarkRef
TEA-PSUtriethylammonium TNTmaximum power density 285 mW cm−2 at 60 °C [20]
TEA-PSU1-methyl-3-(3-trimethoxysilylpropyl) Im-chloride TNT 302 mW cm−2 at 60 °C (5 wt%)[21]
TMA-PAEKtrimethylammonium (3-chloropropyl)-trimethoxysilane TNTconductivity 52.5 mS cm−1 at 80 °C[22]
CS QHNTTS 52 MPa, Young’s modulus 1100 MPa
(5 wt%)
[23]
PVA MWCNTmethanol permeability 3.57 × 10−7 cm2 s−1[24]
PVAgrafted MWCNTKOH 8 M solution uptake 108 %[25]
PVAFeCNT maximum power density 88 mW cm−2 at 60 °C (DMAFC)[26]
PVAgrafted MWCNTmaximum power density 65 mW cm−2 at 60 °C (DEAFC)[27]
PVA/poly(diallyldimethylammonium chloride)MWCNTs-OH TS 40 MPa (3 wt%)[28]
CSMWCNTs-OH conductivity 6 mS cm−1 at RT[29]
glycidyltrimethylammonium chloride-CS (GTA, AR)CNTTS 32 MPa[30]
2,3-epoxypropyl trimethyl ammonium chloride-CSQSiO2-CNT maximum power density 81 mW cm−2 at 60 °C (DMAFC)[31]
Q-trimethylamine polystyrene-block-poly(ethylene-ran-butylene)-block-polystyreneTMA-MWCNTmaximum power density 187 mW cm−2 at 60 °C (DMAFC)[32]
Im-PPOPIL(BF4) MWCNTconductivity 56 mS cm−1 at 75 °C[33]
Im-PEEKIm-CNTconductivity 135 mS cm−1 at 70 °C, 100% RH[34]
Im-PEEKpoly(vinyl imidazole)-CNTmaximum power density129 mW cm−2 at 60 °C
(AEMFC)
[35]
Im-PSUIm-MWCNTsTS 24 MPa
(0.2 wt%)
[36]
Table 2. Type of polymer and 2D filler in composite AEM with LDH or MXene.
Table 2. Type of polymer and 2D filler in composite AEM with LDH or MXene.
Polymer2D FillerRemarksRef
(TMA, DABCO)-PSUMgAl-LDH (Cl)Young’s modulus 620 MPa
(100% RH)
[37]
XL glutaraldehyde-PVAMgAl-LDH (CO32−)EtOH permeability 1.8 × 10−7 cm2 s−1[38]
2,3-epoxypropyltrimethylammonium
chloride-PVA
MgAl-LDH(CO32−, NO3)NO3 conductivity 156 mS cm−1 at 80 °C, TS 48 MPa[39]
glycidyltrimethylammonium chloride-CS/PVAMgAl-LDH (NO3)TS 24 MPa[40]
(2,3-epoxypropyl trimethyl ammonium chloride)-CS/PVACNT coated with MgAl-LDH (CO32−)maximum power density 107 mW cm−2 (DMAFC)[42]
glycidyl trimethyl ammonium chloride-CS/PVAMgAl-LDH (CO32−) wrapped on quaternized SiO2 nanospheresCO32−
conductivity 11 mS cm−1 at 80 °C
[41]
TMA-PSUMgAl-LDH (NO3)TS 21 MPa[43]
DABCO-PSUZnAl-LDH (BmimSO4)conductivity 16 mS cm−1 at 25 °C[44]
1-methylimidazole PPOMgAl-LDH (CO32−)tensile strength 29.5 MPa[47]
1-(N′,N′-dimethylamino)-6,11-(N, N,N-trimethylammonium) undecane-PPO MgAl-LDH with N,N,N-trimethylpropyltriethoxysilane ammonium chloride conductivity 122 mS cm−1 at 80 °C,
maximum power density 267 mW cm−2 at 60 °C
[46]
1-(N′,N′-dimethylamino)-6-(N,N′-dimethylammonium)-11-(N,N′,N″-trimethyl ammonium)undecane-PPO
(TC-PPO)
MgAl-LDH, 3-hydroxy-6-azaspiro [5.5] undecane, N,N,N-trimethyl-3-(triethoxysilyl)propan-1-aminium bromide
(ASU-LDH)
IEC values 3.11–3.90 meq g−1[45]
TC-PPOMgAl-LDH (ASU-LDH) electric-field-alignedconductivity 110 mS cm−1 at 80 °C[48]
DABCO-PPOMgAl-LDHYoung’s modulus 320 ± 60 MPa (100% RH)[49]
electrospun PVDFMg4Al2(OH)12CO3·3H2O conductivity 87 mS cm–1 at 70 °C, 100% RH[50]
triethylene diamine-PSULiF-Ti3C2Tx; NH4HF2-Ti3C2Txmaximum power density 101 mW cm−2 at 60 °C[52]
CSIm brush-functionalized MXeneTS 41.0 MPa
(7.5 wt%)
[53]
Table 3. Type of polymer and 2D filler in composite AEM with graphene, GO, carbon nitride or boron nitride.
Table 3. Type of polymer and 2D filler in composite AEM with graphene, GO, carbon nitride or boron nitride.
Polymer2D FillerRemarksRef
PVAexfoliated graphenemaximum power density 46 mW cm−2 at 60 °C[54]
PVA/CSgraphene, sulfonated grapheneconductivity 47 mS cm−1 at 25 °C[55]
QPEI/PVAsilica functionalized GOconductivity 72 mS cm−1 at 30 °C[56]
XL glycidyl-trimethylammonium chloride-PVAGOEtOH permeability 3.65  × 10−7 cm2 s−1 at
60 °C
[57]
glycidytrimethyl ammonium chloride PVAaligned GO-Fe3O4maximum power density 172 mW cm−2 at 60 °C (0.1 wt%)[58]
glycidytrimethyl ammonium-PVA (QPVA)GO-Fe3O4 conductivity 47–55 mS cm−1 at 30–60 °C (0.1 wt%)[59]
TMA-PSU QGsTS 205 MPa (0.25 wt%)[60]
TMA-QPSUdopamine modified GO (DGO) TS 13 MPa, elongation at break 33% (1 wt%)[61]
TMA-QPSUXL-GO swelling ratio 3.9%, water uptake 19% at 60 °C (2 wt%)[62]
diethanolamine-modified PSU (HPSU)guanidinium-GOswelling of 9% at 60 °C[63]
3-azidopropyl-N,N-dimethylamine (Ap-DMA) and TMA (AMPSU) QPSU(azide XL) rGOYoung’s modulus 2890 MPa, RH 100% (0.1 wt%)[64]
TMA-PPO/PSU GOIEC 3.21 meq g−1
(2 wt%)
[65]
Im-functionalized bisphenol PSUIm-GOmaximum power density 79 mW cm−2 at 60 °C (0.2 wt%) [66]
TMA-PSUQPbGs (polymer brush funzionalized graphenes)conductivity 56 mS cm−1 at 80 °C (1 wt%)[67]
XL 3-(dimethylamino)-1-propylamine PSUpolydopamine-reduced GO (PDArGO)Young’s modulus 1843 MPa (1.5 wt%)[68]
TMA-PSU XL reduced and functionalized GO (rGO) conductivity 140 mS cm−1 at 80 °C
(2 wt%)
[69]
TMA-PSU rGO modified with pyrene-containing tertiary aminesconductivity 140 mS cm−1 80 °C
(2 wt%)
[70]
DABCO-PSUC16 GO19 × 105 S min cm−3 selectivity for VRFB[71]
BrPPO(PEI)-GOIEC 3.59 meq g−1[72]
cellulose/DABCO-PPOGOconductivity 215 mS cm−1 at 80 °C[73]
Im-PPO1-(3-aminopropyl)-3-methylimidazolium bromine (IL-GO)IEC decreased from 1.90 to 1.34 meq g−1 in 2 M NaOH, 80 °C, 480 h [74]
Fumion® FAA-3graphene with surface area of 500 m2/gconductivity 113 mS cm−1 at 80 °C[75]
JAM-II-07 (Yanrun, China)sulfonated rGOarea specific resistance 3.72 Ω cm2 [76]
sp-PBIrGOmaximum power density
544 mW cm−2 at 90 °C
[77]
PBIspin-coated GOTS 50 MPa[78]
PBIMGO and NGOconductivity 24 mS cm−1 (NGO) at 80 °C (1 wt%)[79]
TMA-PAEKrGOConductivity 115 mS cm−1 at 90 °C (1 wt%)[80]
PPEK QA-CDβ@GOYoung’s modulus 1243 MPa (10 wt%) [81]
Im-PEEKIm-GO maximum power density 50 mW cm−2 at 50 °C (4 wt%)[82]
DABCO-PAE QGOmaximum power density 136 mW cm−2 at 70 °C
(0.7 wt%)
[83]
perfluorinated AEM
(I-PFSO2NH2-Cl)
IGNRs
(GNRs grafted APTMS and MIMC)
conductivity 121 mS cm−1 at 80 °C (1 wt%)[84]
XL-QSIBS GO quaternized with: octadecylamine (GOA), octadecylamine
+ N,N-dimethyl-1,3-propanediamine (GOAN)
conductivity
19.5 mS cm−1
(GOAN 0.50 wt%) at 60 °C
[85]
QSIBSpoly (vinylbenzyl chloride) grafted graphene (GN-g-PVBC)conductivity 18 mS cm−1 at 60 °C, storage modulus 418 MPa (0.55 wt%)[86]
copolymer PMS/b-VIB/DPEBIbutylvinylimidazolium GOconductivity 102 mS cm−1 at 100 °C[87]
GO multilayer membranesGO and GOKOHwater uptake
GOKOH 1099 wt%
[88]
TMA-PAES g-C3N4maximum power density 68 mW cm−2 at 60 °C
(0.6 wt%)
[89]
TMA-PAEKg-C3N4 nanosheetsconductivity 35 mS cm−1 at 80 °C (0.5 wt%)[90]
N-methyl morpholine-PPOf-BNYield stress 37 MPa
(5 wt%)
[91]
Table 4. Type of polymer and of 3D filler in composite AEM with silica and silicate fillers.
Table 4. Type of polymer and of 3D filler in composite AEM with silica and silicate fillers.
Polymer3D FillerRemarksRef
XL TMA-PVDF SiO2 conductivity 3 mS cm−1 at RT
(2 wt%)
[95]
QCS SiO2 coated PVDF grafted with trimethyl-3-(trimethoxysilyl) propyl ammonium chloride conductivity 41 mS cm−1 at 80 °C
(10.6 wt%)
[96]
XL PVA/3-(trimethylammonium) propyl-functionalized silica SiO2 maximum power density 50 mW cm−2 at 60 °C (DEAFC) [97]
3-(trimethyl ammonium)-PVA 3-(trimethyl ammonium) propyl-functionalized SiO2storage modulus 172 MPa at 100 °C (20 wt%)[98]
GTMAC-PVA FS conductivity 35 mS cm−1 at 60 °C (5 wt%) [99]
CNC-PVA SiO2 conductivity 65 mS cm−1 at 60 °C (40 wt%) [100]
PVA and CS AM-4, 4VP, AS4, UZAR-S3 conductivity 1 mS cm−1 at RT (4VP/CS:PVA) [101]
TMA-PPO SiO2 UH 0.041 m h−1, S 49 [102]
TMA-PES functionalized SiO2 ASi-I, ASi-II, ASi-III conductivity 46 mS cm−1 at 25 °C (3 wt%, ASi-II) [103]
PPO Im-SiO2 conductivity 105 m S cm−1 at 80 °C [104]
TMA-PSU TEA-SiO2 (QSBA) OCV 0.86 V, power density 298 mW cm−2 (3 wt%) [105]
TMA-PSU TMA-SiO2 (QMSNs) Young’s modulus 2250 MPa (20 wt%) [106]
TEA-PSU Im-mesoporous SiO2 maximum power density 278 mW cm−2 at 60 °C [107]
TMA-PSU modified montmorillonite conductivity 47 mS cm−1 at 95 °C (5 wt%) [108]
TMA-PSU palygorskite conductivity 93 mS cm−1 at 80 °C (0.5 wt%) [109]
TMA-cardo-poly(etherketone) (QPEK-C) N-(trimethoxysilylpropyl)-N,N,N-trimethylammonium TS 26 MPa, elongation at break 32% [110]
Table 5. Type of polymer and 3D filler in composite AEM with metal oxides and other inorganics.
Table 5. Type of polymer and 3D filler in composite AEM with metal oxides and other inorganics.
Polymer3D FillerRemarksRef
GTMAC-PVA Al2O3 conductivity 48 mS cm−1 at 70 °C [111]
TMA-PSU Al2O3 TS 31 MPa (4 wt%) [112]
TMA-PSU ZrO2 conductivity 15 mS cm−1 at RT [113]
Im-PSU ZrO2 conductivity 80 mS cm−1 at 50 °C [114]
TMA-PAES nano ZrO2 conductivity 48 mS cm–1 at 80 °C
(10 wt%)
[115]
XL TMA-coPAES nano ZrO2Young’s modulus 492 MPa
(10 wt%)
[116]
Perfluoro(phenyl 2,2:6,2-terpyridine); 2,2:6,2-terpyridineZrO(ClO4)2 IEC 0.76 meq g−1[117]
TMA polystyrene-block-poly(ethylene-ran-butylene)-block-polystyrene (PSEBS)/TMA-PSU SiO2, ZrO2, TiO2maximum power density 75 mW cm−2 at 60 °C (7.5% TiO2) [118]
TMA-PSU TiO2conductivity 13 mS cm−1 at 21 °C
(10 wt% TiO2)
[119]
vinylbenzyl chloride-divinylbenzene copolymers amorphous TiO2 conductivity 43 mS cm−1 at 30 °C [120]
TMA-PPO/PSU TiO2 maximum power density 118 mW cm−2 at 60 °C (2 wt%) [121]
DABCO-PSUTMP-TiO2
PMHS-TiO2
conductivity 39 mS cm−1 (PMHS-TiO2), 34 mS cm−1 (TMP-TiO2) at 25 °C in KOH 2M [122]
TEA-PPO 1-methy-3-methylimidazolium-TiO2Young’s modulus 921 MPa
(TEA-PPO-1TiO2-15IL)
[123]
TEA-PPO 1-methy-3-methylimidazolium-TiO2 conductivity 52 mS cm−1 at 80 °C [124]
TMA-PVA TiO2 maximum power density
125 mW cm−2 at 35 °C
[125]
PVA CaTiO3 conductivity 66 mS cm−1 at RT [126]
PTFESn0.92Sb0.08P2O7maximum power density 147 mW cm−2 at 200 °C [127]
CoOOH-PVACoSO4maximum power density 144 mW cm−2 at 30 °C [128]
TMA-PVA/CSMoS2TS 33 MPa [129]
TMA-PPO/PSUZnOmaximum power density 69 mW cm−2 at RT [130]
PBI1-butyl-3-methyl imidazolium phosphotungstateconductivity 76 mS cm−1 at 80 °C
(20 wt%, PBI/(PWA-IL1:5))
[131]
poly vinyl benzyl trimethylammonium hydroxideZIF-8BET 1045 to 600 m2 g−1[133]
PVAZIF-8conductivity 0.3 mS cm−1 at 60 °C[134]
IL/PVAIL@ZIF-8conductivity 1.0 mS cm−1 at 60 °C
(20 mol%)
[135]
PVAZIF-8maximum power density 173 mW cm−2 at 60 °C (PVA/40.5% ZIF-8) (DMAFC)[136]
BrPPO/PVAMIL-101-Fe-NH2-Fconductivity 145 mS cm−1 at 80 °C[137]
Im-PEEKimidazolium MIL-101(Cr)TS 35 MPa, (10 wt%) [138]
Im-PSUQCDsYoung’s modulus 1600 MPa[139]
TMA-PSUTMA-GQDs conductivity 15 mS cm−1 at RT
(0.1 wt%)
[140]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Narducci, R.; Sgreccia, E.; Knauth, P.; Di Vona, M.L. Anion Exchange Membranes with 1D, 2D and 3D Fillers: A Review. Polymers 2021, 13, 3887. https://doi.org/10.3390/polym13223887

AMA Style

Narducci R, Sgreccia E, Knauth P, Di Vona ML. Anion Exchange Membranes with 1D, 2D and 3D Fillers: A Review. Polymers. 2021; 13(22):3887. https://doi.org/10.3390/polym13223887

Chicago/Turabian Style

Narducci, Riccardo, Emanuela Sgreccia, Philippe Knauth, and Maria Luisa Di Vona. 2021. "Anion Exchange Membranes with 1D, 2D and 3D Fillers: A Review" Polymers 13, no. 22: 3887. https://doi.org/10.3390/polym13223887

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