Ganesan
Sriram
*ab,
Karmegam
Dhanabalan
ab,
Kanalli V.
Ajeya
b,
Kanakaraj
Aruchamy
a,
Yern Chee
Ching
c,
Tae Hwan
Oh
*a,
Ho-Young
Jung
*bde and
Mahaveer
Kurkuri
*f
aSchool of Chemical Engineering, Yeungnam University, Gyeongsan 38541, Republic of Korea. E-mail: sriramyu@yu.ac.kr; taehwanoh@ynu.ac.kr
bDepartment of Environment and Energy Engineering, Chonnam National University, 77 Yongbong-ro, Buk-gu, Gwangju 61186, Republic of Korea. E-mail: jungho@chonnam.ac.kr
cDepartment of Chemical Engineering, Faculty of Engineering, University of Malaya, Kuala Lumpur 50603, Malaysia
dCenter for Energy Storage System, Chonnam National University, 77 Yongbong-ro, Buk-gu, Gwangju, 61186, Republic of Korea
eHybrid Power Pack Specialized Research Center, Chonnam National University, 77 Yongbong-ro, Buk-gu, Gwangju, 61186, Republic of Korea
fCentre for Research in Functional Materials (CRFM), JAIN (Deemed-to-be University), Jain Global Campus, Bengaluru-562112, Karnataka, India. E-mail: mahaveer.kurkuri@jainuniversity.ac.in
First published on 2nd October 2023
The production of hydrogen through water electrolysis using anion exchange membranes (AEMs) is gaining popularity worldwide due to its cost-effectiveness and efficiency. AEMs play a crucial role in enabling hydroxide (OH–) conduction while preventing fuel crossover, thereby influencing the performance and durability of water electrolysis (WE) systems. This comprehensive review aims to explore recent scientific research on membrane development, specifically focusing on hydrocarbon and fluorocarbon-based AEMs. It analyzes AEMs' physiochemical structures, including morphologies, topologies, surface properties, functional groups, organic components, and thermal stability. AEMs have different characteristics, such as ion exchange capacity, ionic conductivity, water uptake, swelling ratio, mechanical stability, and pH stability, which are also discussed in detail based on their backbone structures. Additionally, the review delves into the current advancements in AEM-based water electrolysis by examining the latest literature on membranes. It also explores the progress in electrocatalysts used in hydrogen production for anion exchange membrane water electrolysis (AEMWE) systems. By discussing the prospects of AEM research and addressing the associated challenges, this review provides insights into the future direction of this field. Overall, this comprehensive review serves as a valuable resource for understanding the development, properties, and applications of AEMs in water electrolysis, shedding light on the advancements in membrane technology and their impact on hydrogen production.
![]() | ||
Fig. 1 Schematic representation of the world's CO2 emissions from fossil sources in 2021, reproduced from ref. 24, Copyright 2023, The Global Change Data Lab. |
Techniques | Hydrogen sources | Efficiency (%) | H2 cost ($ per kg) | Advantages | Disadvantages |
---|---|---|---|---|---|
Water electrolysis | Water | ∼60–80 | ∼10.30 | Pollution-free techniques, producing high-quality grade hydrogen, and O2 is a by-product | Splitting water requires a lot of energy, expensive processes, transportation, and storage crisis |
Steam methane reforming | Fossil fuel (natural gas) | ∼74–85 | ∼2.27 | High yield in the production of hydrogen | High endothermic reaction resulting in a large amount of heat and possibilities of CO and CO2 emission |
Biomass | Steam (woody biomass) | ∼35–50 | ∼3.80 | Using low-cost sources and obtaining methane gas is a by-product | Possibility of deforestation, a large amount of space required, and an inefficient process |
Coal gasification | Fossil fuel (coal) | ∼60–75 | ∼0.96 | The cleanest and most energy-efficient thermal process | Expensive and not efficient |
Nuclear energy | Nuclear (water) | ∼45–50 | ∼7.0 | No emission of CO2 gas | The waste product is radioactive, therefore tricky to dispose of and a costly technique |
Photolysis | Solar (water) | ∼0.5 | ∼18.32 | O2 is a by-product, rich feedstock, and has no gas emissions | Less efficiency, inefficient photocatalytic material, needs sunlight, etc. |
Thermolysis | Thermal (water) | ∼20–45 | ∼8.40 | Green and clean process, and produces O2 as a by-product | Operational expenses are high, and there is a corrosion issue |
Pyrolysis | Fossil fuel (natural gas) | ∼35–50 | ∼1.6–1.70 | Scale-up process, cheap sources, and no CO2 emission | Lack of availability and contaminants in the feedstock |
Partial oxidation | Thermal (hydrocarbons) | ∼60–75 | ∼1.48 | High infrastructure-based technology | Producing petroleum coke and other oils as by-products and producing H2 |
Auto thermal reforming | Thermal (hydrocarbons) | ∼60–75 | ∼1.48 | Well-proven technology | CO2 emission as a by-product |
Bio-photolysis | Water | ∼11 | ∼2.13 | O2 as a by-product and working under very mild conditions | Low yields of H2, the requirement of sunlight and a large reactor, and a high-cost process |
Similarly, the pH conditions affect the oxygen evolution reaction (OER) route. In this reaction, O2 is produced by oxidizing two molecules of H2O in acidic conditions. The OER process involves the oxidation of four OH− ions to H2O and O2 molecules in alkaline and neutral conditions.49 Nicholson and Carlisle primarily explained this electrolysis process in the eighteenth century.52 There are several benefits to producing hydrogen using water electrolysis, such as an easy process, environmental friendliness, and guaranteed product quality. In rural areas with abundant solar and wind energy, poverty-stricken villages can utilize this electrolysis method to generate hydrogen to meet their energy needs for lighting and heating the households, telecommunication stations, and small-scale level industries.53–55 The hydrogen generated by water electrolysis is considered as the best energy carrier to produce power for industries, domestic, etc.
Reaction in the process:
2H2O → 2H2 + O2, E0 = 1.229 V | (1) |
In an acidic environment:
Cathode (negative): 2H+ + 2e− → H2, E0 = 0 V | (2) |
Anode (positive): 2H2O → O2 + 4H+ + 4e−, E0 = 1.23 V | (3) |
In an alkaline environment:
Cathode (negative): 2H2O + 2e− → H2 + 2OH−, E0 = −0.83 V | (4) |
Anode (positive): 4OH− → O2 + 2H2O + 4e−, E0 = 0.40 V | (5) |
Water electrolysis is favorable in both conditions, such as acid and alkaline, because of the availability of water molecules in the deprotonated form for OER in alkaline solutions and hydrogen evaluation reaction (HER) in acid solutions. The production of hydrogen by water requires a minimum free energy of 237.2 kJ mol−1 for the overall process of water electrolysis.56 Industrial water electrolysis began in the 19th century. Since then, electrolysis has grown to 400 and above operation units.57 Three main water electrolysis types can be considered, namely, alkaline water electrolysis (AWE) and proton exchange membrane electrolysis (PEMWE), that generate hydrogen at low temperatures (less than 90 °C), and solid oxide electrolysis (SOE), which is performed at high temperatures (700 to 950 °C) and is under development because a low voltage-based cell is required.45,57 To this degree, alkaline electrolysis is a well-known technology for hydrogen production up to the megawatt scale, and it is the most employed electrolytic technology worldwide. The main challenges in this electrolysis is the corrosive electrolyte and minimal current densities because of the mild mobility of hydroxide (OH−). A few specialty applications of PEMWE can be found in the market with decent performance and stability. The disadvantages of this electrolysis are the requirements of specific materials to operate the corrosive acidic cell and expensive noble metal catalysts (Ru, Pt, and Ir). SOE has attracted special attention because it converts electrical energy into chemical energy, producing high-purity and efficient hydrogen. However, SOE needs high temperature and pressure to operate cells, has low stability, and leads to degradation of electrodes.58 On the other hand, anion exchange membrane water electrolysis is a low-cost and ongoing efficient technology that can replace PEMWE, AWE, and SOE for producing hydrogen due to its ability to operate at high pressures and its reliance on anion exchange membranes (AEMs) and non-noble metal catalysts.
The electrochemical cells in this water electrolysis utilize hydrocarbon AEMs and transition-type metal electrodes, and the electrolytes can be either water or a slightly less-concentrated alkaline solution. Moreover, its operation in an alkaline environment uses inexpensive metal catalysts for OER and HER at the anode and cathode, respectively.32 The catalytic activity of platinum (Pt) for HER and iridium (Ir) in OER is exceptional, but the metals are expensive. Also, low-cost and abundant transition metal electro-catalysts can be used instead of those noble metals. Therefore, these low-cost components of AEMWE significantly reduce the electrolyze system's production cost, making it more suitable for commercialization. Generally, AEM-based technology could be low-cost and incredibly stable for hydrogen generation. AEMWE's working performance must match or beat PEMWE's characteristics, such as hydrogen quality, lifetime, high current density, and clean water operation. AEMWE system is similar to PEMWE, constructed in a heap with membranes to offer higher energy density. It is compatible with several renewable energy sources, such as solar and wind.62 This electrolysis process is based on alkaline conditions with the functional groups immobilized on polymer chains as the interface of the membrane and, for the development of AEMWE systems, should be considered significant on account of membrane stability, ionic conductivity, and integration of catalysts into AEMWE systems. In recent years, the literature has focused on the progress of catalyst-based fabrications rather than anion exchange membranes. There are few research reports on AEMWE cell performance using water until 2021. However, researchers are working continuously to develop AEMWE for hydrogen production.
AEMs are one of the most important components of AEMWE. Its main function is to transfer OH− ions from the cathode to the anode chamber, limit gas movement, and transmit electrons during electrochemical reactions.97 AEMs consist of a hydrocarbon polymer and an anion exchange functional group as the main and side chains. An AEM's main backbones are the polymeric materials normally used, such as polystyrene (PS) and polysulfone (PSF), to link divinylbenzene (DVB) and the relative ion-exchange groups that have ammonium or phosphonium groups.27,98 For WE applications, the ion-exchange membrane must have good permselectivity, ionic conductivity, thermal and chemical stability, mechanical stability, etc.99,100 It is still relatively challenging to fabricate AEM for useful applications to stabilize the tricky issues of decent mechanical stability and high ionic conductivity. Moreover, excess water may diminish the polymer's mechanical strength, while the polymer's chemical stability is not guaranteed, given the possibility of attack by hydroxide ions.27,101,102 Besides, AEM's chemical stability is low at higher pH, a major concern.29,103 There are two mechanisms, namely, Hofmann elimination, which involves degrading the anion-exchange groups under a basic environment through nucleophilic attack of OH− ions on N-alkyl groups at high temperatures,104 while the second degradation mechanism involves electrochemical oxidation.105 Fluorocarbons replace hydrocarbon-based AEMs in WE applications to solve the above-mentioned problems. Incorporating fluorine atoms into the polymer's backbone is more desirable. Fluorocarbon-based polymers may exhibit exceptional properties, such as high stability in alkaline environments, high thermal stability, high hydrophobicity, high microphase separation, mechanical durability, low swelling, and high conductivity due to the strong C–F bond and low polarizability, which can be attributed to fluorine's smaller size and greater electronegativity.106–108 More details on fluorocarbon-based membranes can be seen in section 2.2.
There have been recent developments in AEMs materials such as poly(arylene ether), polytetrafluoroethylene, poly(olefin), poly(phenylene), poly(arylene ether ketone), poly(vinyl alcohol), poly(ethylene-co-tetrafluoroethylene), poly(phenylene oxide), poly(aryl piperidinium), poly(ether imide), polysulfone, and cardo polyether ketone, and the conductivity of hydroxide ions is high in these polymers.89,109–121 Despite this, most studies employ commercial membranes with membrane processes and applications.122–127 Tokuyama A201, Fumasep FAA-3, NREL Gen 2, and Sustainion are all examples of commercially available membranes based on imidazolium or quaternary ammonium polymers; nevertheless, Sustainion was determined to have the highest performance for AEMWEs.61 However, due to the limitation of AEMs, there are important concerns for the long-term stability of AEMWE systems and the temperature. Accordingly, comprehensive research has been initiated to develop AEM materials with higher thermal and chemical stability in alkaline environments. Despite that, as novel AEMs materials are proposed and developed, they might improve hydrogen energy production. Scopus data shows that publications on AEMs in synthesis have steadily increased since 2010 (Fig. 3). But only a small amount of synthetic AEMs is used in water electrolysis. Since 2012, the number of publications has been slowly growing, and more studies may be done in the future. The AEMs' goal in water electrolysis is to replace commercial AEMs, make a low-cost process, speed up the movement of ions from the cathode to the anode, and get a high rate of electron transfer in the electrochemical process so that hydrogen can be made from water. Several AEMs have been developed, and their advantages and disadvantages are listed in Table 2.
![]() | ||
Fig. 3 Articles on the synthesis of different AEMs and the increasing use of AEMs in water electrolysis have been published in the scientific literature (Data source: Scopus, dated 13th July 2023). |
Polymers | Advantages | Disadvantages |
---|---|---|
Poly(2,6-dimethyl-1,4-phenylene oxide) (PPO) | Low-cost, commercially available, high mechanical and thermal stability, and rich functional properties | Weak in an alkaline environment, poor chemical permanence in high temperatures, poor ionic conductivity, immediately decompose because their backbones have ether linkages, and poor solubility in dipolar-based solvents due to their hydrophobicity |
Poly(ether sulfone) (PES) | Excellent mechanical properties, high thermal stability, dimensional stability, high transparency, ion exchangeability, and high flexibility in chemical modification | Non-crystalline, hydrophobic, quickly decomposes and is readily attacked by ester, ketones, hydroxide, aromatic hydrocarbons, and halogen |
Poly(aryl ether sulfone) (PAES) | Commercially available, thermal stability, and remarkable mechanical properties | Poor chemical permanence at high temperatures and high base environments |
Poly(arylene ether) (PAE) | Excellent film-forming property, remarkable solubility in organic diluents, excellent glass transition temperatures, high conductivity, chemical, thermal, oxidative stabilities, and high corrosion resistance | High hydrophobicity and low alkaline stability and, therefore, degrade under alkaline environments at high temperatures |
Polysulfone (PSF) | Excellent in chemical resistance, mechanical properties, low-cost, oxidative stability, high hydroxide conductivity, and thermal stability | Nucleophilic attack by OH ions due to their aryl ether groups, poor ionic conductivity, and low alkaline stability |
Polyether ether ketone (PEEK) | Highly thermoplastic, ease of processability, good mechanical strength, light-weight, self-lubrication, high chemical and abrasion resistance, high adhesive properties, highly resistant to attack by organic solvents and heat resistance | Expensive, a high-temperature process, semi-crystalline, highly hydrophobic, and easily attacked by sodium and halogen agents |
Poly(aryl ether ketone) (PAEK) | Commercially accessible, good mechanical and thermal stability, and easy chemical modification | The low ionic conductivity, poor chemical permanence in high temperatures, and high base environment |
Poly(terphenyl dimethylpiperidinium) (PTP) | Chemically stable under an alkaline environment, with high mechanical and conductive properties | High water absorption and poor mechanical characteristics |
Poly(vinyl chloride) (PVC) | Non-toxic, low-cost, excellent binder properties, high thermal and chemical stability, and good mechanical performance | Poor hydrophilic and high toughness |
Poly(vinylbenzyl chloride) (PVBC) | Low-cost, abundant chloromethyl group in structure, good solubility in polar solvents, and commercial availability | High brittleness and low conductivity, and alkaline stability |
Polybenzimidazole (PBI) | Low-cost, high oxidative, low-water uptake, high hydroxide conductivity, high chemical, mechanical, and thermal stability | Dissolve in high-boiling point polar solvents and low conductivity |
Polyvinyl alcohol (PVA) | Readily available and film-forming, biodegradable, biocompatible, cross-linking ability, flexible, thermostable, tensile strength, highly soluble in water, non-toxic, and low-cost | High water uptake, high swelling, poor drying, low ionic conductivity, and selectivity |
Polyethylene (PE) | Slight swelling, high chemical stability, hydrophobicity, high crystallinity, high mechanical properties, low cost, and ease in functionalization by radical grafting technique | Low thermal and mechanical stability, high gas permeability |
Polystyrene (PS) | Commercially available, the comfort of chemical modification, easy access, renovation, low-cost, strong alkaline, and high thermal stability | Weak mechanical stability, high rigidity, non-flexibility, poor film-forming ability, low thermal and chemical stability, limited processability, and poor chain entanglement |
Poly(vinyl pyridinium) | Excellent electrochemical properties | Low resistance in an alkaline environment |
Polyphenylene ether (PPE) | Good dimensional stability, low moisture absorption, high thermal and mechanical stability, and high dielectric strength | High cost, poor ionic conductivity, and alkaline stability |
Poly(ether imide) (PEI) | Low cost, easy access, good thermal stability, renovation, ease of chemical surface modification, high hydroxide conductivity, mechanical reliability, and chemical resistance | Unstable in alkaline medium, high cost, and low OH ion mobility |
Polyethylenimine (PEI) | Low toxic, water-soluble, available in both linear and branched structures | Poor film-forming ability |
Polyketone (PK) | Low-cost, eco-friendly, easy processability, excellent functional properties, high rigidity, flexibility, high chemical resistance, and low gas permeability | Low solubility, low thermal and mechanical stability |
Polyether ketone (PEK) | Excellent thermal and chemical stability, high glass transition temperatures, and high chemical stability in harsh environments | OH− ions easily attack it |
Chitosan (CS) | Naturally and abundantly available polymers, hydrophilic, excellent mechanical properties, thermally stable, film-forming ability, ease of quaternate biopolymer, non-toxic, and low cost | Low mechanical stability, low alkaline strength, and low conducting properties |
Cellulose | Abundance, low-cost, hydrophilic, high thermal and mechanical properties, renewability, abundant hydroxyl groups, high surface area, and biocompatibility | Poor conductivity, solubility issues in organic solvents, and low conductivity |
Polystyrene-b-poly(ethylene-co-butene)-b-polystyrene (SEBS) | Commercially available, ether-free polymer, high thermal stability, chemical stability, mechanical properties, and alkaline stability | Low solubility |
Polyvinylidene fluoride (PVDF) | Low-cost, good processability, high dimensional stability, low melting point, high mechanical stability, excellent thermoplastic, high chemical resistance, used as a binding agent | It is highly hydrophobic and unstable to attack through bases, ketones, amines, esters, and semi-crystallinity |
Polytetrafluoroethylene (PTFE) | Excellent chemical, electrical resistance, high crystallinity, high durability, high acid and base resistances, hydrophobic | High cost, inelastic, low abrasion-resistance, and low mechanical stability |
Polyolefin | Low cost, high chemical resistance (acid and alkali), ease in processability, and high hydroxide conductivity | The polyolefin's inertness attribute makes it difficult to change the structure, low alkaline stability, and high hydrophobic |
Polyfluorene | Outstanding alkaline stability and film-forming | Low mechanical stability, dimensional strength, flexibility, and conductivity |
Recent reviews of an AEM focused primarily on fuel cell applications, an electrocatalyst for AEMWE, the technical evaluation of a water electrolysis cell, alkaline water electrolysis for H2 generation, the economic analysis of H2 production through electrolyzer technologies, the design of a catalyst and electrode for seawater electrolysis, and the challenges and future opportunities of a component in AEMWE.27,29,32,45,49,50,55,57,58,150–162 In addition, there are reviews on AEMs such as molecular design and microscopy analysis of membranes for fuel cell and water electrolysis, dealing with alkaline durability of imidazolium-based AEMs, synthesis of ion-exchange membranes for various applications, synthetic approaches for alkaline-stable AEMs, Friedel–Crafts reaction-based synthesis of AEMs for fuel cell and electrolysis applications, click chemistry in AEMs for energy-based applications, the significance of polymer ion exchange membrane structures and properties for hydrogen production by water electrolysis, synthesis of silica-AEMs composite, and anion exchange polyelectrolytes for the synthesis of AEMs and ionomers.50,88,145,160,163–173 However, no comprehensive review of AEMs from synthesis to application through physiochemical analysis and evaluation of properties has been identified in recent studies. This review is distinct from other contemporary works because it thoroughly reviews the synthesis of AEMs, such as hydrocarbon and fluorine, with choices of homogeneous (organic/organic) and heterogeneous membranes (organic/inorganic) to their chemical structure of backbones and functionalization of cationic groups and side chains. Furthermore, several analytical approaches were used to explore the structure of AEMs and their important attributes, such as ionic conductivity (IC), ion exchange capacity (IEC), water uptake and swelling ratio, mechanical and alkaline stabilities, and their application in water electrolysis. The synthesized AEMs and other commercial AEMs used in hydrogen production have been discussed and tabulated regarding cell voltage, current density, and durability. Further, an electrocatalyst is also an important component in water electrolysis for hydrogen production. Accordingly, the performance of various electrocatalysts in AEMWE is also discussed. The review aims to discuss the current reports to fill the research gaps of AEMs; hence, research areas in which more exploration and improvements are highly needed. Finally, we concisely conclude the topic and confer the challenges and prospects for the imminent development of AEM for WE application. The authors are hopeful that this review will be helpful to researchers who employ AEMs in their work on water electrolysis. Fig. 4 shows the overall review in a schematic diagram of the synthesis, properties, and application of AEMs for water electrolysis.
![]() | ||
Fig. 4 A schematic representation of the synthesis, identifiable characteristics, and strength of AEMs for water electrolysis. |
![]() | ||
Fig. 5 (a) Synthesis of Q-PES-C, reproduced with permission from ref. 177, Copyright 2005, Elsevier. (b) Synthesis of QPPO, reproduced with permission from ref. 178, Copyright 2014, Elsevier. (c) Synthesis of QAPVA and its cross-linking procedure, reproduced with permission from ref. 180, Copyright 2008, Elsevier. |
In another approach, various polyelectrolytes are designed with the base polymer of PPO to combine phase-separation and a cross-linking method to improve the physicochemical properties of AEMs (e.g., CBQAPPO), and the synthesized polymer membrane side chain has two QAs, which improves the membrane performance.179 Poly(vinyl alcohol) (PVA) is a low-cost, water-soluble, and hydrophilic polymer. As a result, Xiong et al. developed QAs-grafted PVA-based AEMs in 2005.180 To obtain the solid form of quaternized PVA, KOH solution and 2,3-epoxypropyl trimethylammonium chloride (EPTMAC) were added to dissolved PVA at 65 °C and observed for 4 h (Fig. 5c). After that, quaternized PVA (QAPVA) was cross-linked with glutaraldehyde under constant stirring. The dissolved solution was cast for 6 h at 45 °C to produce the membrane using glass. At the end of the study, they obtained a membrane (cross-linked quaternized PVA) after drying it. Due to the carcinogenic properties of chloromethyl ether, Vinodh et al. synthesized AEMs using the non-toxic chloromethylation agent para-formaldehyde.175 A 48 h reaction of chloromethylated polystyrene ethylene butylene polystyrene (PSEBS) was prepared by chloromethylating it with hydrochloric acid and paraformaldehyde using zinc chloride as a catalyst. Afterward, the quaternizing agent triethyl amine (TEA) was used to prepare the quaternized chloromethylated PSEBS (QAPSEBS) under nitrogen (N2) conditions. Finally, the membrane was prepared by casting using the QAPSEBS solution, and the prepared membrane was alkalized with 1 M KOH at RT for 24 h.
Polyketone (PK) is one of the polar polymers with desirable properties such as good mechanical strength, heat resistance, chemical resistance, and stability against corrosion.181,182 Due to the advantages of PK, Ataollahi et al. developed PK-based AEMs by amination and methylation (e.g., PK-PDAPm).183 PK was synthesized from carbon monoxide and ethylene reactions in the presence of palladium (Pd). In addition, the prepared PK was modified with polyamine using the amination process, producing amine and QAs from 1,2-diaminopropane (DP) and triethylamine, respectively. Additionally, 1,1,1,3,3,3-hexafluoroisopropanol (HFIP) and ammonium hydroxide were used to dissolve the modified PK. Afterward, the solution was cast on the Petri dish to form a membrane. The membrane was then immersed in CH3I (methylation) and 1 M KOH to acquire anion exchange characteristics. Recently, a sequence of poly(arylene ether sulfone) (PAES) bearing a high density of QAs (TMA) on the main chain was synthesized using the bisphenol monomer for dimensional stability.184 The monomer 4,4′-bis (4-(4-((4-phenol)diphenyl methyl)phenoxy)phenyl)sulfone (BTP-OH) was synthesized in the presence of DCM, and the prepared polymer was dissolved using boron tribromide (BBr3). BTP-OH was further functionalized on PAES (PAES-BTP) in the presence of BTP-OH, bis(4-fluorophenyl)sulfone, bis(4-hydroxyphenyl)sulfone, and sulfolane by polycondensation, followed by chloromethylation using CMME and zinc chloride (as Lewis catalyst) to obtain chloromethylated CM-PAES-BTP. Further, it was dissolved in dimethylacetamide (DMAc) to cast on the glass to prepare the membrane (60–70 μm), followed by TMA quaternization. The quaternized membrane (QPAES-BTP) was further soaked in 1 M KOH for two days to form the OH− ions. Because chloromethylation takes seven days, this process may not be easy to follow to make instance membranes. Similarly, polyether ketone (PEK) was also synthesized by polycondensation in the presence of NMP and potassium carbonate using 4,4′-difluorobenzophenone, 4,4′-isopropylidenediphenol, and 4,4′-isopropylidenediphenol.185 Later, the PEK copolymer was chloromethylated with CMME (CM-PEK) in the presence of 1,1,2,2-tetrachloroethane. The obtained CM-PEK was dissolved in NMP, cast on glass, and dried to obtain the membrane (35–40 μm). Furthermore, the chloromethylated membrane was quaternized (QAs-PEK) using TMA, followed by QAs-PEK being dissolved in 1 M KOH to form the OH− ions for producing AEMs (QAPEK-OH). This process can be expensive and time-consuming since it involves several procedures and chemicals to synthesize copolymers.
Due to its stiff aromatic structure, polybenzimidazole (PBI) has exceptional thermal stability, chemical resistance, mechanical strength, etc. The long side-chain alkyl-tethered QAs groups of AEMs could improve the alkaline stability. Accordingly, Li et al. developed long side-chain QAs groups (LSCQAs)-grafted PBI AEMs with high chemical stability and conductivity.186 To this degree, firstly, high molecular weight PBI was prepared via the polycondensation of 3,3′-diaminobenzidine and 4,4′-oxybisbenzoicacid (Fig. 6a). Then, highly durable PBI-grafted LSCQAs AEMs were prepared by a LSCQAs group and added in dissolved PBI solution. The LSCQAs-grafted PBI were soaked in dimethyl sulfoxide (DMSO) and cast to the prepared membrane on the glass (N-PBI). This process could be a low-cost and simple method to synthesize AEMs. According to another study, LSC (5-bromopentyl)-trimethylammonium bromide (LSC-BPTMA-Br)-grafted PBI was synthesized by nucleophilic substitution reaction (N–S reaction).187 Using ultra-sonication, PBI and BPTMA-Br were mixed and dissolved in DMAc, then cast onto a Petri dish and dried at 80 °C for a day to prepare the membrane (PBI-BPTMA). In this study, the long side-chain QAs are advantageous to enhance the membrane's hydrophilicity and conductivity of ions. Other researchers used naphthalene-based PBIs (NPBIs) grafted with cationic side chains ((6-bromohexyl)trimethylammonium bromide) (NPBIs-QAs) to form a membrane, and later, bromobutane-reacted NPBI-QAs (NPBI-QAs-by) were synthesized (Fig. 6b).188 The prepared NPBI-QAs and NPBI-QAs-by solid precipitations were dissolved in DMSO, then cast onto glass to prepare the membrane (40 μm). The two kinds of membranes with Br− ions were replaced by OH− ions by soaking them in 1 M NaOH.
![]() | ||
Fig. 6 (a) Synthesis and grafting of PBI, reproduced with permission from ref. 186, Copyright 2018, Elsevier. (b) Synthesis of NPBI-QAs and grafting bromobutane on it (NPBI-QAs-by), reproduced with permission from ref. 188, Copyright 2021, Elsevier. |
Recently, a sequence of poly(arylene ether ketone) (PAEK) copolymers containing long alkyl chains (LAC) heavily quaternized carbazole-based pendants was studied by Liu et al.110 The LAC attached to the pendants improves the copolymer flexibility and increases chain entanglement, which is attributed to increasing toughness and strength of AEMs. The main goal of their work was to construct LAC between the hydrophobic backbone and the pendant hydrophilic. Briefly, the prepared PAEK holding long alkyl brominated pendants was dissolved in 1,1,2,2-tetrachloroethane (TCE) solution, followed by casting on glass to prepare the membrane. Further, treatment with TMA, followed by incubation in 1 M KOH for 72 h, resulted in OH− ions in the AEMs (PAEK-HQACz). However, the membrane manufacturing process required multiple steps with chemicals and is very time-consuming. A highly stable and conductive polymer backbone with a C–NH2 link was recently fabricated by reductive amination in conventional PAEK (QA-PE-NH2) using the Leuckart reaction.189 PAEK was derived from bisphenol A, bis(4-fluorophenyl)-methane, and potassium carbonate dissolved in the mixture of DMAc and toluene through a chemical reaction at 160 °C. The dissolved PAEK was chloromethylated (CM-PAEK) by chloromethyl octyl ethers (CMOE). The CO groups of CM-PAEK were then converted to CM-NH2 (CM-PAEM) by reductive amination through Leuckart reaction. It was then quaternized by TMA, cast on the glass to prepare the membrane (43 μm), and then soaked in 1 M KOH to form the OH− ions, and this process could be a moderate and time-saving method for membrane fabrication.
Researchers have become interested in hydroxide-based AEMs (HAEMs) since they can replace the metal catalysts used in fuel cells and electrolyzers. HAEMs mainly have cationic groups (sulfonium, ammonium, pyridinium, phosphonium, imidazolium, guanidinium, etc.) and hydroxide ions.190–192 Among those, QAs-based HAEMs have been widely used in electrochemical energy systems because of their sufficient stability in an alkaline medium. The most challenging aspect of producing HAEMs is improving the cationic groups' chemical and polymeric membrane stability by cross-linking paths. Accordingly, Vengatesan et al. synthesized AEMs with a cross-linking route, i.e., the process implies the synthesis of quaternized poly(vinyl benzyl chloride) (QPVBC) AEMs by a low-cost amination reagent using hexamethylenetetramine (HMTA).193 This work entails a single-step membrane process by casting the quaternized polymer solution. The quaternized membrane was later immersed in 0.5 M KOH to obtain the membrane into hydroxide form. A simple immersing and methylation process was used by Hossain et al. to synthesize cross-linked transparent AEMs (CQPPO).194 Brominated poly-(2,6-dimethyl-phenylene oxide) (BPPO) was initially dissolved in NMP. Then, the dissolved polymer was cast on the glass to produce the membrane. Later, the cross-linked membrane (C-TA-PPO) was obtained by immersing the membrane in a dimethylamine (DMA) solution. The tertiary amine produced by the DMA in which the amine-functional characteristics of the membrane are essential for quaternization. Next, they washed C-TA-PPO with 0.1 M KOH, followed by water to remove unreacted ammonium bromide and residual alkali. C-TA-PPO was quaternized by treatment with methyl iodide and then treated with 1 M KOH to give it ion exchange properties. During the QAs functionalization, only the BPPO membrane was exposed to the TMA solution and then soaked in 1 M KOH to convert the bromide form of the membrane to the hydroxide form.
Cross-linking is a practical and effective approach generally carried out by heat treatment.195 The cross-linked AEMs show better mechanical, thermal, and chemical stability and solve the swelling water issues.196 Characteristically, hydrogen, ionic, and covalent bondings are engaged in the process of AEMs cross-linking.197,198 Among these bondings, cross-linking by covalent bonds is considered highly stable and extensively used to prepare AEMs. However, AEMs cross-linked by covalent bonds could not be dissolved, making it challenging to recycle the waste AEMs. To these concerns, Hou et al. fabricated recyclable cross-linked BPPO-based AEMs through disulfide bonds, having QAs groups, and also prepared AEMs without cross-linking (RC-QPPO).196 Initially, dissolved BPPO was quaternized using a reaction with TMA (Fig. 7a). The treated polymer solution was further cast to prepare the membrane. Finally, QAs-BPPO AEMs developed by the membrane was soaked in 1 M NaOH to produce hydroxide ions. BPPO was dissolved in the mixture of NMP and potassium thioacetate (AcS) for cross-linking the membrane, followed by TMA. The obtained quaternized polymer precipitation was washed with HCl and dried. Further, the precipitate was dissolved in dimethylformamide (DMF), followed by casting. Finally, a disulfide bond cross-linking with AEMs was obtained by soaking in 1 M NaOH. Recycling QAs-BPPO AEMs (RC-QPPO) was performed by dissolving the membrane in the DMF and dithiothreitol (DTT) mixture. After dissolving, the solution was cast to prepare a membrane and cross-linked by the same procedure. The recyclable AEMs materials make this process very cheap.
![]() | ||
Fig. 7 (a) Synthesis of recyclable RC-QPPO, reproduced with permission from ref. 196, Copyright 2018, Elsevier. (b) Synthesis of (i) PPO-C-nQA-30 and (ii) PP-mQA copolymers, reproduced with permission from ref. 200, Copyright 2021, Elsevier. (c) Procedure for the synthesis of acSxQAPSF, reproduced with permission from ref. 201, Copyright 2020, Elsevier. |
The ion-conducting property of QAs (TMA) and ethylene oxide (EO) cross-linked with different cross-linkers were used to prepare PPO (Q-CPPO), endowing it with excellent mechanical and thermal properties.199 This cross-linked structured membrane has high ionic conductivity and chemical stability. However, this process is time-consuming and expensive because multiple cross-linkers and chemicals must be synthesized to prepare the membrane. Chu et al. recently reported the study of two sets of QAs-PPO AEMs bearing different cation side chains with and without triazole-containing linkers.200 Initially, various lengths of alkyne-functionalized cation QAs (nQA) were synthesized by amination using TMA (Fig. 7b). The PPO was then catalyzed with Cu(I) bromide by reacting with nQA (PPO-C-nQA). The quaternization of this bromoalkylated polymer (PPO-mBr) was continued in the presence of NMB to replace Br− ions with QA groups (PPO-mQA), where the letter “m” refers to the number of carbon (C) atoms between the polymer backbone and QAs. PPO-C-nQA and PPO-mQA membranes were then produced by dissolving in NMP and casting on Petri dishes, followed by vacuum drying. The membrane was further soaked in 1 M NaOH to produce OH− ions. The process also involves many chemical requirements such as expensive procedures and several steps in the synthesis of the membrane.
Recently, Han et al. synthesized QAs aggregated and crosslinked sulfonated PSF AEMs.201 Firstly, the sulfonated PSF (S-PSF) was prepared by dissolving PSF in sulfonating agents (chlorosulfonic acid and trimethylchlorosilane) and treating it with sodium methoxide to form S-PSf-Na precipitates (Fig. 7c). Further dissolution of the precipitate in DMF, followed by casting onto the glass, yielded the membrane. As the sodium ion has been replaced by hydrogen in the membrane (SPSF-Na), it has been soaked in 1 M HCl (SPSF-H). In addition, the obtained SPSF-H was crosslinked with chloromethylated PSF using n-butylamine (n-bu) to get cS-CMPSF. The aggregated form of cross-linked quaternized membrane (acS-QAPSF) was obtained by the n-bu cross-linking process, followed by the reaction of TMA. The process involves multiple steps and is expensive and time-consuming. TMA is used for membrane fabrication and is quite alkaline.
The conventional method of dispersing polymer films directly in QAs (TMA) solution for the quaternization process has drawbacks, such as irregular reactions and unwanted aggregates of TMA. Yu et al. overcame these issues by cross-linking the TMA with PSF using ultraviolet (UV) cross-linking (c-QCMPSF).202 In this study, chloromethylated PSF was prepared using trimethylchlorosilane (TMCS) and dissolved in DMAc. TMA solution and the photo-responsive cross-linker trihydroxypropane triacrylate (TMPTA) were added. After the solution was poured onto the glass, a thin film was created with a knife and treated with UV radiation for 20 min, followed by oven curing at 60 °C for a day to obtain the membrane.
The aryl ether bonds of backbone polymers (PSF, PEEK, PPO, etc.) are not strong under basic environments because hydroxide ions are easily attracted and severely attacked. Additionally, crosslinking using chemical or physical methods in the AEMs is an effective technique to improve the chemical and mechanical stabilities of the membrane. However, few cross-linkers lead to decreased membrane conductivity. Accordingly, researchers have identified a diamine that is not just a cross-linker. Still, it also acts as a QA reagent, which can help to improve the conductivity of the cross-linked AEMs. Accordingly, Gao et al. synthesized a sequence of different diamine cross-linker-based membranes (C-CQASEBS) developed from the aryl-ether free block copolymer poly(styrene-ethylene/butylene-styrene) (SEBS) used as a base polymer.203 The cross-linking process was between the chloromethylated SEBS and various diamine cross-linkers (N,N,N′,N′-tetramethyl-1,2-ethylenediamine (TMEDA), N,N,N′,N′-tetramethyl-1,4-butane diamine (TMBDA), and N,N,N′,N′-tetramethyl-1,6-hexane diamine (TMHDA)). After preparing various cross-linked membranes, the membrane was further soaked in TMA to increase the conductivity. The SEBS was the backbone polymer because of its excellent chemical stability. The synthesis process is straightforward to reproduce the membranes and could be a low-cost protocol.
The cross-linking method is limited due to the incorporation of cross-linkable moieties; it is incompatible with the polymer's main chain and is complicated to produce. Alternatively, quaternizable groups can be cross-linked by a di-tertiary amine, which provides cations and cross-links in the membrane. Thus, polymers with linear structures have better membrane properties due to their firmer and healthier chain entanglement than those with hyperbranched, dendritic, or star-shaped structures. Research on various polymer structures could lead to improvements in AEMs. AEMs fabrication through hyperbranched polymer (HBP) is less popular because of simpler chain entanglement, making it challenging to develop AEMs. Liu et al. used two synthetic protocols, including cross-linking and RAFT polymerization, to fabricate AEMs (HPS) from HBP to overcome this issue.143 The prepared hyperbranched polystyrene (PS) can be easily dissolved in NMP and cast on the glass to form a flexible membrane. After that, the membrane was cross-linked with TMHDA in NMP solution and cast on the glass to obtain a cross-linked membrane. Also, hyperbranched poly(vinyl benzyl chloride) (HPVBC) was prepared using the RAFT polymerization method, and HPVBC was treated with polyisoprene to obtain HPVBI. Later, it was dissolved in NMP to prepare the membrane, followed by quaternized (QHPVBI) using TMA and then soaked in 1 M NaOH for ion exchange. They confirm that the flexible HBP AEMs exhibit excellent properties compared to AEMs fabricated from linear polymers.
AEMs for electrolyzers need a high hydroxide conductivity. Compared to randomly based copolymers, multiblock-based copolymers can enhance the hydroxide conductivity.204 Therefore, Mandal et al. prepared a sequence of multiblock-based copolymers (poly(BuNB-b-BPNB-b-BuNB-b-BPNB)) bearing a hydrocarbon backbone to fabricate AEMs.205 The synthesis of this multiblock copolymer typically involved the polymerization of norbornene with vinyl addition using an active catalyst ((allyl)palladium(triisopropylphosphine)chloride) and an activator (lithium tetrakis(pentafluorophenyl)-borate·(2.5Et2O)). The synthesized multiblock copolymer was then dissolved in chloroform and cast on an aluminum dish to fabricate the membrane (50 μm). Further, quaternization (QAs) was performed on the membrane using TMA. After soaking in 1 M NaOH, the Q-multi-block AEMs were replaced with OH− ions to replace the bromide (Br−) ion. In another study, star-shaped block copolymers were synthesized by anionic polymerization to fabricate AEMs (QSCP).127 This simple synthesis process used polymerization to prepare star-shaped polyisoprene-b-poly(4-methylstyrene) (SCP). The SCP was brominated (SCP-Br) in this synthesis by reacting with N-bromosuccinimide (NBS). In addition, SCP-Br was dissolved in chloroform and cast onto glass to prepare the membrane. The membrane was treated with TMA for quaternization and soaked in 1 M KOH to replace the Br− ions with OH− ions. This technique can make a cheap membrane with minimal chemicals, processes, and time management. Poly(ether imide) (PEI) AEMs with high alkalinity and ion conductivity were synthesized by Oh et al. through an N–S reaction.136 4,4′-Oxydiphthalic anhydride (ODPA) and 4,4′-oxydianiline (ODA) were used to synthesize PEI in the presence of DMAc. A solid PEI was obtained by drying and heating it at 200 °C. CMME was used to chloromethylate PEI (CM-PEI) in the presence of TCE and tin chloride, which was used as a catalyst in Friedel–Crafts (F–C) reaction. In addition, CM-PEI was dissolved in TCE and cast on glass, then dried to obtain the membrane. Then, it was soaked in TMA for quaternization, followed by washing in 1 M KOH to form OH− ions (QPEI). It would be a low-cost process, but using CMME for this process would not be ecologically safe due to the toxic nature of the material.
Recently, different polyethylene (PE) membranes were developed by Espiritu et al. and treated with gamma radiation to improve their surface properties.206 This radiation method generates active sites on the polymer backbone for further functionalization. This study used low-density PE (LDPE) and linear low-density PE (LLDPE) dissolved in toluene and cast onto a Petri dish to prepare the membrane (50 μm). Then, they were treated with varying monomer concentrations (vinylbenzyl chloride (VBC)) and gamma radiation. Having obtained the membranes, they were functionalized with TMA for quaternization, and then OH− ions were generated by 1 M KOH. By polymerization, the monomers methyl methacrylate, vinylbenzyl chloride, and ethyl acrylate were mixed to produce a polymer called poly(MMA-co-VBC-co-EA) (PMVE).207 The copolymer was dissolved in DMF to obtain the membrane and cast onto the glass. Later, the membrane (50–70 μm) was immersed in TMA for quaternization and soaked in 1 M KOH for ion exchange (QPMVE). They did not use any chloromethyl agents in this membrane process because vinylbenzyl chloride has chloromethyl groups in its structure.
Recently, Zhang et al. developed polysulfone (PSf)-based alkaline-stable AEMs that are not quaternized or imidazolium-based membranes (e.g., PSf-Im-OH).208 PSf is a recognized high-performance polymer frequently used to produce ion-exchange membranes. CMME is used to prepare chloromethylated PSf (CMPSf). Later, the membrane was made by dissolving CMPSf in N,N-dimethylacetamide, followed by adding 1-methyl-imidazole (MIm) under stirring (Fig. 8a). After the solution was cast on the glass substrate, the obtained membrane was soaked in a 0.5 M NaOH solution for alkalization. A new method was proposed by Gong et al. in the synthesis of AEMs containing pendent imidazolium (Im) side chains with an ether-containing spacer between the chloromethylated PSf and hydroxyl-holding imidazolium.210 In brief, hydroxyl-bearing Im(1-(2-hydroxyethyl)-3-methylimidazolium chloride (HMIM-Cl)) was synthesized in the first step of the process (Fig. 8b). Then, HMIM-Cl reacted with chloromethylated PSF (CMPSf) by the Williamson etherification reaction to obtain pendant Im side chain-modified PSF (PIMPSF-Cl). Further, it was dissolved in 1-methyl-2-pyrrolidone (NMP), followed by solution casting on the glass to obtain the membrane (70 μm). At last, the membrane was immersed in 1 M KOH to obtain a hydroxyl ion (PIMPSF-OH). This process also reveals several steps and requires many chemicals to produce the AEMs, which could be a moderate cost.
![]() | ||
Fig. 8 (a) Synthesis route of PSf-Im-OH−, reproduced from ref. 208, Copyright 2011, Royal Society of Chemistry. (b) Chemical structure for synthesizing the pendulous imidazolium (Im)-functionalized PSf (PIMPSf-OH−), reproduced with permission from ref. 210, Copyright 2017, Elsevier. (c) Schematic depiction for the synthesis route of Im-PES, reproduced with permission from ref. 213, Copyright 2016, Elsevier. |
Other groups synthesized 1,4,5-trimethyl-2-(2,4,6-trimethoxyphenyl)imidazolium-treated polyphenylene oxide (PPO-TMIM).211,212 Initially, 1,4,5-trimethyl-2-(2,4,6-trimethoxyphenyl) imidazolium (TMIM) was synthesized from 5-dimethyl-2-(2,4,6-trimethoxyphenyl)imidazole using a chemical procedure. This membrane was prepared by casting PPO mixed with TMIM on a Petri dish. The membrane's (PPO-TMIM) hydroxide form was obtained by soaking in 1 M KOH. This lengthy, multi-step procedure for preparing membranes is time-consuming. The conventional methodologies considered are chloromethylation, and the bromination functionalization step is crucial in preparing Im-based AEMs. However, this routine is complex. Therefore, synthesizing Im-based AEMs via a simplistic and efficient functionalized practice is recommended. Poly(ether sulfone) (PES) is a thermoplastic polymer with excellent thermal stability, mechanical stability, anti-corrosion properties, etc. Due to these properties of PES, Liu et al. developed a new route for synthesizing Im-functionalized poly(ether sulfone) (Im-PES).213 To this end, double bonds PES matrix (PES-DB) was produced with a monomer bis(3-allyl-4-hydroxyphenyl)sulfone, and then the PES-DB matrix was able to react immediately with the ionic liquid 1-allyl-3-methylimidazolium chloride (AmimCl). As a next step, Im-PES was prepared by mixing NMP with PES-DB and adding 3-methylimidazolium chloride (Fig. 8c). Further, this mixture was reacted with benzoyl peroxide. Then, the transparent solution was cast on the glass to the prepared membrane, and after that, it was soaked in 0.5 M NaOH solution.
In polymer chemistry, physical or chemical cross-linking techniques are efficient practices to attain greater stability and high electrochemical performance of the membrane. Among the cross-linking methods, chemical approaches are widely used to optimize the membrane's chemical structure to achieve good conductivity and thermal, chemical, and mechanical stability. However, a chemical cross-linking method has the disadvantage of causing unwanted gelation. Accordingly, a novel Im-functionalized end-group cross-linked poly(arylene ether sulfone) (PAES) AEMs was synthesized by Lee et al.214 The Im-functionalized PAES contains several steps that begin with the nucleophilic aromatic substitution (NAS) reactions, where PAES reacts with tetramethyl moieties (M-PAES). By dissolving M-PAES in TCE and treating it with benzoyl peroxide (BPO) and N-bromosuccinimide (NBS), M-PAES was brominated (Br-PAES). After drying Br-PAES, the Im group was introduced into the Br-PAES copolymer (Im-PAES) using MIm. The 3-hydroxyphenylacetylene (HPA) salt was then prepared under the reflux method using thermal trimerization. The prepared salt form of the end-group cross-linker was further reacted with the dissolved Im-PAES in NMP to produce end-group cross-linked Im-PAES (E-Im-PAES). The dissolved E-Im-PAES was cast onto the glass plate to prepare the membrane. After introducing the alkyne end-group trimerization at 180 °C, the prepared membranes were immersed in 1 M NaOH. The process involves many protocols and several chemicals, making it lengthy and moderately costly. Another novel approach by Lin et al. synthesized 1,2-dimethyl-3-(4-vinylbenzyl)imidazolium-chloride-based AEMs (DMVIm-Cl) using simple procedures.209 Initial steps included synthesizing [DMVIm][Cl] by stirring 1,2-dimethylimidazole with 4-vinylbenzyl chloride. Afterward, the mixture of solutions was then stirred under stirring to produce a homogeneous solution consisting of styrene, acetonitrile, divinylbenzene (DVB), benzoin ethyl ether (BEE), and [DMVIm][Cl]. The solution was cast on glass and irradiated with UV light to produce the Im-based membrane. The developed membranes were soaked in 1 M KOH, and the chloride ions (Cl−) were converted into hydroxide ions (OH−). A low-cost, straightforward synthesis process could be found in this study.
Polyether ether ketone (PEEK) polymer has several advantages such as high chemical and thermal stability, low production cost, and ease of functionalization with other cationic and anionic ions. Hence, AEMs-based imidazolium-containing PEEK was investigated as an alternative material in the study by Kim et al.215 The method used to fabricate the AEMs is direct synthesis using monomers with an ammonium group and copolymerization. In brief, they synthesized an imidazolium-based monomer tetra(2-ethyl-methylimidazole-methylene)-4,4′-dihydroxy diphenylether (IDHDPE). Then, they were functionalized with polymer PEEK, and further, it was converted to imidazolium iodide functional groups using CH3I (Im-PEEK). The imidazolium-containing polymer solution was cast on the glass to develop the membrane. This process is straightforward and avoids the toxic conditions related to CMME. The membrane was further soaked in 1 M KOH to form the hydroxide form.
Recently, Chen et al. utilized polybenzimidazole (PBI)-based AEMs containing cobaltocenium for the first time.216 A polymer membrane was prepared with various cobaltocenium cations in the microwave reactor, including cobaltocenium, 1,1′-dimethylcobaltocenium and 1,1′-dicarboxycobaltocenium. Therefore, they studied the relationship between these three membranes' structure and working performance. The membranes were immersed in 1 M KOH to prevent damage from other ions. In AEMs, cobaltocenium cations provide outstanding oxidation resistance and chemical stabilities.217,218 This study reveals several steps of synthesis and expensive chemicals involved in these processes. Son et al. recently synthesized a PEEK sequence with various imidazolium (Im) groups using many Im moieties containing monomers via a condensation–polymerization reaction.219 Initially, 4,4′-difluorobenzophenone and the synthesized monomer (0 to 40 mol%) were used to prepare PEEK with several groups of di-imidazole (PEEK-DI-IME). After this synthesis, PEEK with several groups of di-imidazolium was designed by adding iodomethane (CH3I) to prepare PEEK-DI-IME, followed by dissolving them in DMAc to convert the imidazole to imidazolium (PEEK-DI-IMM) precipitates. Further, DMAc solution was used to dissolve PEEK-DI-IMM and casting it onto the glass to obtain the membrane. Finally, 1 M KOH was used to soak the PEEK-DI-IMM membrane to replace the iodide (I−) ions with OH− in the membrane. The synthesized membrane has monomers bearing many imidazolium moieties, has some benefits, shows high stability, and has no gelation or alcohol during the process. The process could be a moderately expensive and very long procedure due to the synthesis of monomers and preparation of various imidazolium content with the polymer, followed by the time-consuming membrane protocol.
![]() | ||
Fig. 9 (a) Chemical structure for synthesizing PSf-DABDA, reproduced with permission from ref. 220, Copyright 2019, Elsevier. (b) Diagram of transformation from poly(methoxy-triptycene ether sulfone) to pyrazolium cross-linked poly(triptycene ether sulfone) through demethylation, film casting, and ion exchange process, reproduced with permission from ref. 221, Copyright 2019, American Chemical Society. (c) The function of various cationic homopolymers, and their chemical structures, reproduced from ref. 225, Copyright 2018, John Wiley & Sons, Inc. |
In dry conditions, AEMs are rigid and weak. They also weaken and soften at high hydration. To overcome these issues, Vandiver et al. synthesized large and thin polyethylene-b-poly(vinylbenzyl trimethylammonium) AEMs (PE-b-PVBTMA) by the anionic polymerization method.222 Initially, the polymer (polyethylene-b-poly(vinylbenzyl bromide)) was prepared using a chemical procedure.223 After this polymer synthesis, it was dissolved in xylene under continuous stirring to produce a membrane. They used an automatic thin-film applicator to obtain a consistent thickness of the membrane. After casting the membrane, the membrane was immersed in TMA solution to functionalize the vinylbenzyl trimethylammonium. After that, bromide ions were developed on membranes to prevent OH− ions from being converted. Membrane fabrication is a multistep process, which could be expensive. AEMs could be produced under favorable environments by applying this highly efficient immersion-crosslinking technique. Afsar et al. recently synthesized cationic-functionalized BPPO (QBPPO-P) using the N–S reaction.224 To do this, they used 4-(dimethylamino)pyridine (DMAP) as a QA precursor, which was functionalized on BPPO, and low-concentration (0.1 M) HCl was used to produce chloride ions for substituting the bromide ions in the membrane. The process made a favorable environment for membrane preparation using 0.1 M HCl compared to 1 M KOH.
In a recent study, Sun et al. investigated the alkaline permanence of AEMs under the influence of alkaline media, using a sequence of organic cations such as imidazolium, pyridinium, quaternary ammonium, phosphonium, benzimidazolium, and pyrrolidinium synthesized with their corresponding polymers.225 The polymer prepared from the block structure may have a higher level of hydroxide movement than the polymer from a random structure (Fig. 9c). Microphase separation can be created in the block-structured copolymer compared to random-structured copolymers. Microphase separation could also lead to the formation of ion channels in the membrane, which would increase the ion conductivity.205 Due to the advantages of the block copolymer, recently, a sequence of bis-imidazolium (B-Im)-functionalized polynorbornene AEMs with various contents was synthesized by Zhang et al.226 They first synthesized ionic liquids (B-Im functionalized norbornene monomer). Secondly, the block copolymer structure was obtained by the ionic liquids' copolymerization reaction with 5-norbornene-2-methylene glycidyl ether. Thirdly, the block copolymer-structured membrane was prepared using the above copolymerized block copolymer. Finally, AEMs were obtained by soaking in 1 M NaOH to replace Br− ions with OH− ions. Due to the synthesis of several ionic liquid compounds, the approach is moderately expensive. The piperidinium cations and an aryl ether-free backbones main chains were considered to be promising for alkaline stability. Therefore, aryl-ether free backbones and a series of AEMs (49 μm) with long alkoxy-containing bis-piperidinium based side chains along with hydrophobic alkyl chains (e.g., PISBr-NON) were synthesized by Zhang et al.227 This study synthesized 1,2-bis(2-(2-methylpiperidine)ethoxy)ethane (NON) for use as QAs for quaternization.
To minimize the use of poisonous chemicals, chloromethyl agents CMME or CMOE, and quaternary ammonium agents (TMA) in membrane fabrication, Qaisrani et al. developed a simple and environment-friendly method of cross-linking and quaternizing chloromethylated PSF to prepare a membrane.228 In this synthesis, cross-linking and quaternization are achieved in a single step during chloromethylation of PSf by adding the dual functional cross-linker DACEE (refer to ref. 228 for the full form of DACEE), which is derived from 1,4-diazabicyclo(2,2,2)octane (DABCO) (Fig. 10a). A similar procedure was followed for another dual-functional cross-linker, DADBH. The chloromethylation agent used was trimethylsilyl chloride, which is non-toxic. DACEE-functionalized PSf solution was cast on the glass to develop the membrane (PSF-DACEE), and it was further soaked in 1 M NaOH to produce hydroxide ions. The exact process was also conducted for the DADBH-functionalized PSf to prepare the membrane. The synthesis is advantageous over standard cross-linking procedures due to the cheap, environment friendly, and easy processes. Poly(aryl piperidinium) (PAP)-based AEMs have been developed via polycondensation through superacid-mediation using N-methyl-4-piperidone (NMP), and biphenyl helps in the alkaline stability of the membrane.121 As a result, 1,6-dibromohexane (DBH) and 2-bromoethanol (BE) react with PAP to form cationic groups (piperidine) and cross-linked forms (Fig. 10b). A cross-linked structure can be helpful for the design of membranes with high ionic conductivity, mechanical properties, and alkaline stability. Then, cross-linked PAP solution was cast onto a glass surface to form the membrane and soaked in 1 M NaOH to produce the OH− ion (PAP-OH). This method of synthesis and crosslinking does not require chloromethylation or quaternization and is simple, eco-friendly, and low-cost.
![]() | ||
Fig. 10 (a) Chemical structure for synthesizing the PSf-DACEE, reproduced with permission from ref. 228, Copyright 2019, American Chemical Society. (b) Synthesis of PAP-OH, reproduced with permission from ref. 121, Copyright 2021, Elsevier. (c) The synthesis route of PPO-OH− membrane, reproduced with permission from ref. 230, Copyright 2021, Elsevier. |
The heavily arranged cationic groups in block copolymers help improve hydroxide conductivity. The ionic parts and polymer backbones tend to be separated by a smaller spacer in conventional AEMs. Therefore, the backbones must be hydrated and attacked by several hydroxide ions. To solve these problems, Wang et al. created highly arranged bis-cation (1,3-bis(trimethylammoniumbromide-methyl)-5-(prop-2-ynyloxy)benzene (TABB)) types connected to the polymer backbones of PPO AEMs by flexible lengths of spacers to improve the hydroxide conductivity and minimize the hydration of the membrane (e.g., NPPO-2QA).229 However, preparing this type of AEM could be time-consuming and expensive since multiple steps and chemicals are used. Another work demonstrates a simple and low-cost method to synthesize different ratios of PPO AEMs.230 The initial preparation of brominated PPO (BPPO) involved dissolving the PPO in a chlorobenzene solution and reacting with the mixture of N-bromosuccinimide (NBS) and 2,2′-azobis(2-methylpropionitrile) under the reflux method (Fig. 10c). After that, the various ratio of BPPO and DABDA was added in NMP and heated at 85 °C for 2 h, and different solution ratios were cast on the glass and dried at mild temperatures to obtain the membrane (PPO/DABDA). Recently, Liu et al. developed guanidinium-functionalized PPO with enhanced alkaline stability by the reaction of brominated PPO (BPPO) with 2-benzyl-1,1,3,3-tetramethylguanidine (BTMG) used as QAs.141 Then, guanidinium-BPPO was dissolved in dimethylformamide to cast on the glass and dried, followed by the membrane being soaked in 1 M NaOH to form the OH− ions to obtain AEMs (BPPO-G-OH).
The in situ crosslinking of a polymer solution by thermal treatment is a low-cost method of producing self-crosslinks. It shows advantages such as low water absorption/swelling and enhanced resistance to chemical degradation. Wu et al. did not use additional cross-linkers to achieve QAs cation with PPO via in situ cross-linking.195 Using the Menshutkin reaction, the copolymer BPPO-(dimethylamino)ethyl methacrylate (DMAEMA) was synthesized in the presence of NMP. The developed BPPO-DMAEMA was dissolved in NMP to be cast on the glasses to form the membrane. After this, the membranes were soaked in 1 M NaCl to obtain the chloride ion. This process and cross-linking to make side chains with flexibility and pendant alkene groups shows that cross-linking the membranes can be done cheaply and easily without external cross-linkers. The recent work by Wu et al. focused on improving the performance of PPO AEMs (60–80 μm) through efficient Cu(I)-catalyzed azide-alkyne (CuAAC) reaction chemistry.231 The side-chain of the precursor with a terminal alkynyl group and Di-QAs improves the conductivity of hydroxyl groups and the alkaline stability of the membrane (e.g., DQ-PPO-x-OH). This type of AEMs fabrication is costly since it requires various chemicals to synthesize different compounds, making it a multi-step and time-consuming process. Liu et al. recently prepared a series of PPO AEMs with sterically-hindered QAs cations via the CuAAC reaction.232 This study analyzes structural matters for a sequence of side-chain-type PPO with pendant QAs cations. They prepared three types of QAs using the Menshutkin reaction, and then three kinds of alkyne-functionalized ammonium monomers were synthesized by reacting 6-iodo-1-hexyne with three amines separately. Through the CuAAC reaction, three types of quaternized copolymers PPO were synthesized using the monomers. Quaternized copolymers were then dissolved in NMP and cast onto the glass to obtain the membrane (PPO-CH) with a thickness of 50 ± 5 μm. The membrane was immersed in 1 M NaOH to convert the I− ions into OH− ions. Using several chemicals to generate the various cations and monomers for the lengthy and laborious membrane construction process demonstrates that this procedure is costly.
Low-cost SEBS-based AEMs (Q-SEBS) were recently developed by Yu et al.233 The chloromethyl and bromohexanoyl groups were grafted onto the SEBS copolymer through chloromethylation (chloromethyl methyl ether) and F–C reaction synthesis. After grafting, the obtained polymer precipitation was dissolved in chloroform and cast onto a Petri dish, and the resulting membrane was then separately quaternized with Mpy, 2-methyl-1-pyrroline (Mpyl), MIm, and N-methylpiperidine (MPip). After this, the quaternized membrane was soaked in 1 M KOH to produce the OH− ions. Recently, poly[(terphenyl piperidinium)-co-(oxindole terphenylylene)] (PTP) was synthesized by superacid-catalyzed copolymerization, followed by their quaternization (Q-PTP) using CH3I.234 Initially, chemicals such as isatin, N-methyl-4-piperidone, and p-terphenyl were used to prepare the PTP copolymer using polymerization. Later, it was quaternized by dissolving in the mixture of CH3I and NMP. The iodide (I−) form of the PTP copolymer solution was cast on the glass to prepare the PTP membrane (45 μm). The prepared membrane was soaked in 1 NaOH under to replace the I− ions with OH− ions. The process was a moderately expensive and time-consuming process due to many chemicals and multistep processes. Poly(aryl piperidinium) with long aromatic fragments was synthesized using superacid-catalyzed polycondensation (SCP) reaction, followed by quaternization to prepare the membrane (4 μm) with excellent properties.235 The synthesized membrane exhibits outstanding solubility in both DMSO and NMP. However, combining copolymers via this method is expensive due to the extensive use of chemicals, the reaction under low temperatures (−15 to 0 °C), and the time-consuming and multistep procedures. The one-step, low-cost process of aliphatic polymer PVC-based AEMs (e.g., PVC45) was recently developed by Liu et al.129 PVC was dissolved in DMAc at 60 °C with stirring. Afterward, the dissolved solution was cast onto the glass and directed at 60 °C overnight to prepare the membrane. The quaternization degree was determined by soaking the prepared membrane in triethylenetetramine (TETA) at various times. The TETA served to cross-link and provided a rich functional network. Finally, the membrane was soaked in 0.5 M NaCl to replace the Cl− ions with OH− ions, and this type of AEM fabrication is a time-saving and straightforward process. As a result of their higher valency to promote ionic conductivity, metal cations are highly attractive and can replace organic cations in AEMs.
Recently, a metallopolymer AEM was prepared via a one-step process by reacting pyridine-containing monomers 2-Ar, nickel chloride, and dicyclopentadiene (DCPD) and the catalyst in methanol–chloroform at RT.236 After the solution was transferred to a Teflon-based mold, solvent evaporation was followed by drying to produce thin AEMs. This study only had one concern: the preparation of monomers requires a lot of chemicals, which is time-consuming. In another work, polytetraarylphosphonium (pTAP)-based AEMs were developed by Arunachalam et al.237 In this synthesis, anhydrous pyridine and bisphenol were dissolved in DCM in a round-bottom flask under a nitrogen environment, then, they were cooled to 0 °C. Afterward, the above reaction was done with DCM-dissolved triflic anhydride solution and cooled in the flask. The solution was cast onto the glass to prepare the membrane (60 μm). A day was spent soaking the membrane in 1 M KOH to produce OH− ions (pTAP-OH). The polymer synthesis involves many washing processes with DMC, and using an ice bath leads to moderate cost. Recently, poly(dimethoxybenzene-co-methyl-4-formylbenzoate) (PDMB) AEMs with high molecular weight have been developed by Zhu et al. due to their chemical and structural stability.144 By the F–C polycondensation of 1,4-dimethoxybenzene (DMB) and methyl-4-formylbenzoate (MFB), they produced PDMB copolymer. Tetrahydrofuran (THF) and lithium aluminum hydride (LAH) were used to hydroxylate the prepared PDMB (PDMB-OH). Then, it was chloromethylated (PDMB-C) by reacting with thionyl chloride and quaternized with MPip. The quaternized PDMB was then dissolved in DMF, and the dissolved solution was cast on the glass to form the membrane (PDMP-Pi), which was soaked in 1 M KOH solution to produce OH− ions. Due to the use of many chemicals in the multistep process to synthesize the polymer for membrane fabrication, the process could be somewhat expensive. Recently, a simple photo-polymerization (UV radiation) method was used to prepare mono and di-QAs, pyrrolidinium (Py), and Im-functional monomers, and their corresponding polyolefin-based AEMs were studied.238 Electrospinning can control the membrane size, enhance ion transport, and permit large-scale production. Zhou et al. developed polyketone AEMs based on the Paal–Knorr reaction, followed by quaternization.133 A one-pot method was used to synthesize the quaternized functional polyketone (QfPK). The membrane (70–80 μm) was fabricated via casting and electrospun from the resulting solution. By hot-pressing, polyketone AEMs with fibers were produced. However, this process involves synthesizing polyelectrolytes with many chemicals, and electrospinning may be relatively expensive.
The naphthalene monomer is a hydrophobic compound that can be functionalized into the main chain of polymers to enhance the folding performance and ion transport. In this regard, Zhu et al. studied PVP-based AEMs (e.g., NAPEK-PVP-6-Q4) with 40–45 μm thicknesses.239 They designed two block polymers, one of which is naphthalene introduced as the hydrophobic component, and the other is bisphenol A, a general monomer applied as another hydrophobic component. They studied the effects of the hydrophobic parts introduced in the polymer chain on the performance of AEMs. Recently, the structural connections between the different cationic structures (CH3I, N,N-diisopropylethylamine, and 1,4-dibromobutane) functionalized poly(terphenylene)-based AEMs has been investigated, and F–C polycondensation has been used to synthesize the membranes with a thickness of 50 μm.240 CH3I-quaternized poly(terphenylene)-based AEMs (m-TPN-PiQA) performed better than the others. For AEMs, a cation with a double-cyclic structure based on N-spirocyclic may provide good alkaline stability. Thus, in a recent study by Zhang et al., they synthesized a novel type of N-spirocyclic cation (QAs) such as CMBO-ASD (refer to ref. 241 for the full form of CMBO-ASD)-functionalized PSF AEMs (e.g., PSF-xCy-ASD), which were highly stable under alkaline conditions among other types of QAs.241 They synthesized a novel QA, CMBO-ASD, then reacted it with chloromethylated PSF for quaternization. Further, quaternized PSf was dissolved in DMSO, cast on the glass to prepare the membrane (46 μm), and then soaked in 1 M KOH to form OH− ions. The procedure involved in synthesizing and producing hydrocarbon AEMs by functionalizing various QAs is detailed in Table 3.
AEMs | Copolymer | Process | BA | MA | CL | QAs | DA | CS | IEA | Cost | Ref. |
---|---|---|---|---|---|---|---|---|---|---|---|
a AEMs = anion exchange membranes; BA = brominating agent; MA = methylation agent; CL = cross-linker; QAs = quaternization agent; DA = dissolving agent; CS = casting substrate; IEA = ion exchange agent. | |||||||||||
Q-PES-C | PES | Polymerization | — | CME | TMA | DMF | Glass | 1 M NaOH | Low | 177 | |
QPPO | PPO | Polymerization | — | CMME | TMA | NMP | Plexiglas | 2 M KOH | Low | 178 | |
CBQAPPO | BPPO | Free radical | NBS | — | — | TMHDA + VBC | NMP | Glass | KOH | Low | 179 |
QAPVA | PVA | Stirrer | — | — | GA | EPTMAC | Water | Glass | KOH | Very low | 180 |
QAPSEBS | PSEBS | Polymerization | — | Paraformaldehyde and conc. HCl | — | TEA | THF | Glass | 1 M KOH | Very low | 175 |
PK-PDAPm (100 μm) | PK | Solvothermal | — | CH3I | — | DP, TEA | HFIP | Petri dish | 1 M KOH | Low | 183 |
QPAES-BTP (60–70 μm) | PES-BTP | Polycondensation | — | CMME | — | TMA | DMAc | Glass | 1 M KOH | Moderate | 184 |
QAPEK-OH (35–40 μm) | PEK | Polycondensation | — | CMME | — | TEA | NMP | Glass | 1 M KOH | Moderate | 185 |
N-PBI (10 μm) | PBI | Polymerization | — | — | — | LSC cation | DMSO | Glass | 1 M NaOH | Very low | 186 |
PBI-BPTMA (60 μm) | PBI | Polymerization | — | — | — | BPTMA-Br | DMAc | Petri dish | 1 M KOH | Very low | 187 |
NPBI-QAs-by (40 μm) | NPBI | Polycondensation | — | — | — | BHQA | DMSO | Glass | 1 M NaOH | Moderate | 188 |
PAEK-HQACz | PAEK-HBrCz | Polycondensation | — | — | — | TMA | TCE | Glass | 1 M KOH | High | 110 |
QA-PE-NH2 (43 μm) | PAEK | Polycondensation | — | CMOE | — | TMA | NMP | Glass | 1 M KOH | Moderate | 189 |
QPVBC (100 μm) | PVBC | — | — | — | HMTA | — | NMP | Glass | 0.5 M KOH | Very low | 193 |
CQPPO (50 μm) | BPPO | Commercially obtained | — | — | DMA | Methyl iodide | — | Glass | 1 M KOH | Moderate | 194 |
RC-QPPO (50–70 μm) | BPPO | Commercially obtained | — | — | AcS | TMA | DMF | Petri dish | 1 M NaOH | Low | 196 |
Q-CPPO (40–50 μm) | PPO | Commercially obtained | — | — | EO-based | TMA | NMP | Petri dish | 1 M KOH | High | 199 |
PPO-C-nQA (50 μm) | BPPO | Click chemistry | NBS | — | 1,2,3-Triazole | TMA | NMP | Petri dish | 1 M NaOH | High | 200 |
acS-QAPSF | SPSF | — | — | CMME | n-bu | TMA | DMF | Glass | 1 M KOH | High | 201 |
c-QCMPSF | PSF | Commercially obtained | — | TMCS | TMPTA | TMA | DMAc | Glass | 0.5 M Na2SO4 | Moderate | 202 |
C-CQASEBS | SEBS | Commercially obtained | — | BCMB | THMDA, TMEDA, TMBDA | TMA | TCE | Glass | 0.01 M NaNO3 | Low | 203 |
HPS | PS and PVBC | Raft polymerization | — | — | TMHDA | — | NMP | Glass | — | Low | 143 |
QHPVBI | PVBC and polyisoprene | Raft polymerization | — | — | — | TMA | NMP | Glass | 1 M NaOH | Low | 143 |
Q-multi-block AEM (50 μm) | Tetrablock copolymer | Polymerization | — | — | — | TMA | Chloroform | Aluminum dish | 1 M NaOH | Moderate | 205 |
QSCP | SCP | Polymerization | NBS | — | — | TMA | Chloroform | Glass | 1 M KOH | Low | 127 |
QPEI (30–33 μm) | PEI | N–S reaction | — | CMME | — | TMA | TCE | Glass dish | 1 M KOH | Low | 136 |
LDPE (30–130 μm) | PE film | Radiation grafting | — | TMA | Toluene | Petri dish | 1 M KOH | Moderate | 206 | ||
QPMVE (50–70 μm) | PMVE | Polymerization | — | — | — | TMA | DMF | Glass | 1 M KOH | Very low | 207 |
PSf-Im-OH 52 μm | PSF | Commercially obtained | — | CMME | MIm | DMAc | Glass | 0.5 M NaOH | Low | 208 | |
PIMPSF-OH (70 μm) | PSF | Commercially obtained | — | CMOE | — | HMIM-Cl | NMP | Glass | 1 M KOH | Moderate | 210 |
PPO-TMIM | PPO | — | — | — | — | TMIM | NMP | Petri dish | 1 M NaOH | High | 211 |
Im-PES (52 μm) | PES-DB | Polycondensation | — | — | — | AmimCl | NMP | Glass | 0.5 M NaOH | Low | 213 |
E-Im-PAES (50 μm) | PAES-M | NAS reaction | NBS | — | HPA | MIm | DMF | Glass | 1 M NaOH | Moderate | 214 |
SAN-[DMVIm][Cl] (100 μm) | [DMVIm][Cl] | Polymerization | — | — | — | — | DVB + BEE | Glass | 1 M KOH | Low | 209 |
Im-PEEK (50 μm) | PEEK | Polycondensation | — | — | — | CH3I | DMAc | Glass | 1 M KOH | High | 215 |
PEEK-DI-IME | PEEK | Polycondensation | — | — | — | CH3I | DMAc | Petri dish | 1 M KOH | Moderate | 219 |
QPSF-DABDA | PSF | Commercially obtained | — | (CH3)3SiCl | — | DABDA | NMP | Glass | 1 M NaOH | Low | 220 |
PE-b-PVBTMA | PE-b-PVBBr | Polycondensation | NBS | — | — | TMA | Xylene | Glass | — | Moderate | 222 |
QBPPO-P | BPPO | Commercially obtained | — | — | — | DMAP | NMP | Glass | 0.1 M HCl | Low | 224 |
PISBr–NON (49 μm) | PISBr | Polymerization | — | — | — | NON | NMP | Glass | 2 M NaOH | High | 227 |
PSF-DACEE (50–60 μm) | PSF | Commercially obtained | — | (CH3)3SiCl | — | DACEE | DSMO | Glass | 1 M NaOH | Moderate | 228 |
PAP–OH (75 μm) | PAP | Polymerization | — | — | DH, BE | N-Methyl-4-piperidone | NMP | Glass | 1 M NaOH | Very low | 121 |
NPPO-2QA | NPPO | Polymerization | NBS | — | — | TABB | NMP | Glass | 1 M NaOH | Low | 229 |
PPO/DABDA (36–40 μm) | BPPO | Polycondensation | NBS | — | — | DABDA | NMP | Glass | 1 M KOH | Low | 230 |
BPPO-G-OH (50 μm) | BPPO | Polycondensation | NBS | — | — | Guanidinium | DMF | Glass | 1 M NaOH | Moderate | 141 |
BPPO-DMAEMA | BPPO | Commercially obtained | — | — | — | DMAEMA | NMP | Glass | 1 M NaCl | Low | 195 |
DQ-PPO-x-OH (60–80 μm) | DQ-PPO-x-I | Polycondensation | — | — | — | — | NMP | Glass | 1 M NaOH | High | 231 |
PPO-CH (50 μm) | PPO-n-X | CuAAC reaction | — | — | — | Different QAs | NMP | Glass | 1 M NaOH | Moderate | 232 |
Q-SEBS | LSC-SEBS | Friedel–Crafts | — | CMME | — | Im, Mpyl, Mpy | Chloroform | Petri dish | 1 M KOH | Moderate | 233 |
Q-PTP (45 μm) | PTP-x | Polycondensation | — | — | — | Methyl iodide | NMP | Glass | 1 M NaOH | Moderate | 234 |
PVC45 (75 μm) | PVC | Commercially obtained | — | — | — | TETA | DMAc | Glass | 0.5 M NaCl | Low | 129 |
pTAP-OH (60 μm) | pTAP | Polymerization | — | — | — | — | — | Glass | 1 M KOH | Low | 237 |
PDMB-Pi (30 μm) | PDMP | Polycondensation | — | Thionyl chloride | — | N-Methyl piperidine | DMF | Glass | 1 M KOH | Moderate | 144 |
QfPK (70–80 μm) | PK | Commercially obtained | — | — | — | Methyl iodide | — | Glass | 1 M KOH | Moderate | 133 |
NAPEK-PVP-6-Q4 | Naphthalene | ATRP reaction | NBS | — | NQQN | — | NMP | Glass | 1 M NaOH | High | 239 |
m-TPNPiQA | m-Triphenyl | Friedel–Crafts | — | — | — | N-Cyclic QAs | DMSO | Glass | 2 M NaCl | Low | 240 |
PSF-xCy-ASD | PSF | Polymerization | — | CMOE | — | CMBO-ASD | DMSO | Glass | 1 M KOH | Moderate | 241 |
Incorporating inorganic nanoparticles into polymer matrixes enables the development of ionic conductive channels through interactions between nanoparticles and polymers. Accordingly, the large surface area, high thermal conductivity, and good electrical and mechanical properties of graphene oxide (GO) make it a viable material for preparing composite AEMs. By oxidizing graphite, it yields a 2D-single-layered oxidized sheet containing functional groups such as epoxide (–O–), hydroxyl (–OH), carbonyl (–CO), and carboxylic (–COOH). Thus, Goel et al. studied functionalized GO as a cross-linker to link with PPO AEMs (e.g., PPO/AGO).242 In the first step, PPO was brominated by dissolving it in chlorobenzene, then N-bromosuccinimide (NBS) and 2,2′-azobisisobutyronitrile (AIBN) were added under continuous stirring to get the brown precipitate (BPPO). Second, GO was synthesized using a modified Hummers' method and then functionalized with 3-aminopropyl trimethoxysilane (APTMS) (AGO) through reflux (Fig. 11a). Thirdly, the crosslinking of BPPO and AGO was achieved by mixing both (BPPO and GO) in NMP solution. Using a cross-linked polymer solution, the solution was cast onto the glass and dried under a vacuum to obtain the membrane (130–150 μm). In addition, the membrane was quaternized by TMA and then soaked in 1 M KOH to replace Br− ions with OH− ions. Due to the use of chemicals to synthesize GO and the membrane process, the process could be moderately expensive. Because of its symmetry constraints, the 6-azoniaspiro[5.5]undecane (ASU) has high alkali resistance among QA cations. The incorporation of ASU cation into AEMs is a promising synthetic strategy. Recently, Long et al. developed layered ASU-functionalized GO and PPO-structured membranes (e.g., ASU-GO/PIPPO) on a silica wafer.243 The first step in this study was synthesizing GO and ASU separately using the modified Hummers' method and a simple chemical reaction, respectively, followed by GO functionalization with ASU (ASU-GO) using the reflux method; then, the ASU-GO was allowed to freeze-drying (Fig. 11b). In the presence of piperidine (using N-methyl piperidine) and ASU (PIPPO) under DMF solution, the bromomethylated PPO was functionalized. The methanol-dispersed ASU-GO ink and the ethanol-dissolved PIPPO were each cast one at a time using the spray technique on a silicon wafer to generate the composite layered membrane.
![]() | ||
Fig. 11 (a) Reaction depiction for aminopropylsilane-treated graphene oxide (AGO) and chemical structure for the synthesis of PPO/AGO membrane cross-linking, reproduced with permission from ref. 242, Copyright 2021, Elsevier. (b) Synthetic process of GO/PIPPO and ASU-GO-functionalized PIPPO, reproduced with permission from ref. 243, Copyright 2020, Elsevier. (c) Synthesis of poly(VSi-co-QVBC)/PVA/fGO membrane, reproduced with permission from ref. 94, Copyright 2018, Elsevier. (d) Synthesis of DGO, reproduced from ref. 245, Copyright 2016, Royal Society of Chemistry. |
GO's poor dispersibility can hinder its usage for various purposes; therefore, modifying GO to develop the composites with enhanced dispersibility to another composite with polymer to prepare AEMs with better structure is vital. Accordingly, Sharma et al. created composite AEMs using silica-functionalized GO (fGO) composite with quaternized PEI (QPEI/f GO/PVA).244 Firstly, Hummers' method was modified to synthesize GO and functionalize with 3-aminopropyltriethoxysilane (APTEOS) under ultra-sonication. Secondly, bromomethane was used to quaternize PEI (QPEI) in the presence of DMSO. Thirdly, the polymers PVA, QPEI, and fGO were dispersed in DMSO under continuous stirring at 90 °C. The prepared solution was cast on the glass to prepare the membrane. Moreover, the membrane was cross-linked with formaldehyde, which was difficult to dissolve the membrane.
The copolymer triethoxyvinylsilane (VSi) with quaternized vinyl benzyl chloride (QVBC) (using N-methylmorpholine) was synthesized recently by polymerization and then functionalization with PVA using the sol–gel method.94 Then, the modification of the functionalized PVA was achieved by adding amide-functionalized graphene nanoribbons (GNR) to DMSO (Fig. 11c). The composite membrane (poly(VSi-co-QVBC)/PVA/fGO) was cast on the polyethylene plate, and the membrane was crosslinked with formaldehyde. It was the role of fGNR in this work to improve the membranes' physical and chemical properties. GO is a suitable inorganic particle for the surface modification of AEMs due to its hydrophilicity and negative charge. Li et al. recently modified commercially available AEMs (ASE, NEOSEPTA, Astom Crop, Japan) with GO (GO@PDA-ASE) using electrodeposition.246 For sealing the GO layer, the modified AEMs were further soaked in the polydopamine (PDA) solution obtained from dopamine hydrochloride in the presence of Tris–HCl solution. The membrane's anti-fouling and stability improved with GO and PDA coatings.
Similarly, AEMs were synthesized using dopamine-modified GO blended with chloromethylated PSF (DGO-CMPSF) by Luo et al.245 Dopamine was then modified with GO (DGO) using a mixture of dopamine hydrochloride and GO (Fig. 11d). Later, they synthesized chloromethylated PSF by treating dissolved PSF (in DMAc) with chloromethyl ether (CMPSF). The obtained CMPSF was mixed with DGO under stirring in the presence of DMAc, followed by cross-linking with N,N,N′,N′-tetramethylethylenediamine (TMEDA). The solution was cast to prepare a membrane (83–90 μm). The membrane was quaternized with TMA and then treated with 1 M KOH to generate OH− ions. In another study, the commercial AEMs Ionics AR204 with a thickness of 500 μm was modified with PDA, followed by the co-modification of nanoparticles (e.g., PDA/GO or TiO2/AEM) using a simple immersion method studied by Xu et al.247 and using this kind of commercial AEMs (Ionics AR204), and further modification could be a low-cost compared to chemically synthesize and modify AEMs. Similarly, Motealleh et al. modified a Sustainion® X37-50 AEMs with zirconia nanoparticles (Sustainion® X37-50/ZrO2) to improve the membrane's mechanical properties using the sol–gel process.248
In a recent study, Emmanuel et al. synthesized PVA-blended imidazolium-functionalized anion exchange silica precursor (PVA/Im-AESP) using an alkylation process, and Im-AESP obtained the corresponding silica-based AEMs mixed with PVA and the addition of tetraethyl orthosilicate (TEOS) via the sol–gel method.249 In detail, Im-AESP was prepared by mixing the silane (3-chloropropyltriethoxysilane) with MIm under the refluxing method. After obtaining Im-AESP, it was added to the PVA solution at 60 °C under stirring and then HCl was added to form a gel. Later, the above mixture was added to TEOS, and the solution was cast onto a Teflon plate to prepare the membrane with 172–191 μm thickness. An easy and low-cost approach for preparing AEMs is demonstrated in this procedure.
Layered double hydroxides (LDHs) are outstanding anionic conductors with higher ion exchange capacity (IEC) than other inorganic particles and have superior alkaline stability. They could be an excellent inorganic material for the composite with the fabrication of AEMs. Therefore, recently, PSF was composited with Mg–Al LDH to synthesize AEMs.134 TMA-quaternized PSF was dissolved in DMF, followed by casting on the glass to fabricate the PSF membrane. In the case of QPSF/LDH, LDH was added into the dissolved QPSF solution, followed by ultra-sonication and cast on the Petri dish to prepare the nanocomposite membrane (PSF/Mg–Al LDH) with a thickness of 35 μm. The membrane was soaked in 1 M KOH and KHCO3 to form the OH− ion. Chitosan (CS) is an alkaline polymer with good film-forming properties. Its structure contains abundant hydroxyl (OH−) and amine (–NH2) groups; this allows it to be quaternized without using chloromethylation, which is considered a toxic process. In light of these advantages of CS, Zhao et al. recently synthesized low-cost quaternized CS/PVA AEMs modified with MgAl (QCS/PVA-LDH).250 Chitosan was quaternized (QCS) by glycidyltrimethylammonium chloride (GTMAC) under acetic solution; then, it was soaked in 1 M KOH to replace the Cl ions with OH ions. Blended LDH/QCS/PVA was prepared by dissolving QCS and PVA in an acetic solution at 90 °C, followed by adding LDH under stirring. Afterward, GA was added to crosslink the matrix at RT, then the membrane was cast onto the glass. This study synthesized Mg–Al LDH by the solvothermal method to obtain a structured flower morphology and prevent the agglomeration of the particles.
Zn2+ ions were incorporated into the PSf to prepare the composite Zn2+/PSf AEMs to enhance the ionic conductivity/ionic channels of PSf.251 The chloromethylating and imidazolium-functionalized PSf (PSf-ImCl) solution was mixed with the dissolved Zn2+ ions (under NMB) solution and then cast on the glass to form the composite membrane (Zn2+/PSf). The composite membrane was then soaked in 1 M KOH to obtain AEMs (PSF-ImOH/Zn2+), and these membrane processes could be simple, low-cost, and time-saving. The copolymers, terpolymers, and main polymers blending of membrane fabrication provide the required potential to enhance the properties of the synthesized membranes. Accordingly, the procured acrylonitrile–butadiene–styrene (ABS) and polystyrene (PS) were dissolved in THF, and then ball-milled resin particles and activated carbon (AC) were added under continuous stirring.252 The mixture solution was cast onto the glass and dried to obtain the composite membrane (ABS/PS/AC), and then the membrane (80–120 μm) was soaked in NaCl solution for further use. The pore-filling strategy can significantly improve the membrane's wettability and enhance the interfacial adhesive strength between the catalytic electrode and membrane. Accordingly, the pore-filling method was used to develop the PTFE/Ni–Fe LDH membrane structure to enhance the alkaline stability.253 A simple process in this work shows that the procured PTFE substrate was allowed into the ethanol solution for the wettability, then dipped 2–3 times into (Ni–Fe) LDH precursor solution, followed by KOH and water. The wettable membrane was dried to obtain a PTFE/Ni–Fe LDH composite membrane.
![]() | ||
Fig. 12 (a) Illustration for the synthesis of QHPEEK, reproduced with permission from ref. 254, Copyright 2019, Elsevier. (b) Schematic depiction for the synthesis of VImPPO and their semi-interpenetrating polymer network (sIPN) membrane, reproduced with permission from ref. 128, Copyright 2021, Elsevier. (c) Schematic illustration of the synthesis route of xPDM-yPEGm membranes, reproduced with permission from ref. 255, Copyright 2021, Elsevier. |
Another flexible and mechanically strengthened structure was demonstrated using a rigid photo-crosslinked PPO and a flexible PVA composite.128 The brominated PPO (BPPO) was synthesized by dissolving it in NBS and AIBN solvents under the reflux method (Fig. 12b). The obtained BPPO was further dissolved in a mixture of NMB and 1-vinylimidazole (VIm) solvents to produce VIm-functionalized BPPO (VImPPO). Following its dissolution in DMSO, 1,8-octanedithiol (cross-linker) and 2-benzyl-2-(dimethylamino)-40-morpholinobutyrophenone (photoinitiator) were mixed. Subsequently, PVA was added to the above solution mixture, dispersed on glass under UV light, and then heated at 80 °C for 24 h to obtain a membrane (VImPPO/PVA) with a 50–60 μm thickness. The membranes were soaked in 1 M NaOH to produce OH−. Since not many chemicals are used, and the process takes less than 50 h, the time-saving procedure is moderately expensive due to the UV light used.
Polyethylene glycol (PEG) is one of the oligomers, and it has some significant features such as high hydrophilic, typical molecular weight, as well as excellent chemical and thermal stabilities. Due to PEG's importance, Wan et al. recently developed PEG blended with functionalized PPO composite AEMs (e.g., xPDM–yPEGm) to improve ion transportation and protect the piperidine ammoniums from the attack of OH− ions.255 In this process, firstly, brominated BPPO was functionalized with 4,4-diethoxybutan-1-amine (DEB) and Mpi under the solvent NMB. Later, PEG was dissolved in NMB and mixed with the above solution to blend. The blended membrane (PEG/PPO) was obtained by casting. Then, the membrane (40 μm) was soaked in 1 M KOH to the OH− form (Fig. 12c).
In contrast to covalent cross-linking, cation–dipole interactions (C–D) provide advantages such as high mechanical strength and reduced swelling. C–D interactions could increase the OH− conductivity in the membranes. Accordingly, poly(ethylene glycol) (PEG) is a polar homopolymer with ample dipoles that enhance ion conduction. Due to the advantages of PEG and C–D interactions, Zhang et al. synthesized and studied C–D interactions between pendent QAs cation-functionalized PPO-grafted PEG (QPPO/PEG).256 In the presence of NBS, brominated PPO was performed. The azidation of BPPO (BrNPPO) was achieved by reacting with NaN3, and then the resulting BrNPPO was quaternized using TMA (QPPO). PEG-grafted QPPO (P-QPPO) residues were collected by mixing QPPO with PEG in the presence of NMB. The precipitates were dissolved in NMB and cast on the glass to obtain the membrane. After soaking in 1 M NaOH, the composite (P-QPPO) membrane formed OH− ions.
Researchers have synthesized a composite of PVA nanofibers crosslinked with glutaraldehyde (GA) and poly[2-20-(m-phenylene)-5-50-bibenzimidazole] (ABPBI) AEMs (PVAf/ABPBI).131 Initially, PVA nanofibers prepared by electrospinning were crosslinked with GA. Afterward, the cross-linked PVA membrane was composited with ABPBI polymer by immersing the film in dissolved ABPBI solution under a reflux method. Later, the membrane was soaked in 3.1 M KOH to form OH− ions for 7 days, longer than other membrane processes for ion formation. The drawback of this study is that the fabrication of the PVA nanofiber membranes using electrospinning requires experts and a 7 day time-consuming process. In another study, quaternized and cross-linked poly(vinylbenzyl chloride) (PVBC) was blended with cross-linked PVA (QPVBC-PVA) to produce mechanically robust and excellent ion conductive AEMs.257 PVBC is quaternized with TMA and mixed with dissolved PVA (subjected to DMSO). Then, the solution was cast on polytetrafluoroethylene (PTFE), and the prepared membrane was soaked in GA solution to crosslink the PVA. Lastly, 1 M KOH was used to soak the cross-linked membrane to form OH− ions and obtain the QPVBC-PVA AEMs. Despite polyethyleneimine (PEI) being highly dense in amino groups and able to cross-link with other polymers, its mechanical strength is high after it is composited with PVA because of the strong hydrogen bonding between PEI and PVA. Accordingly, Xiao et al. developed a low-cost GA cross-linked PVA/branched PEI (PVA/BPEI) membrane by casting.258 This process involved dissolving PVA and BPEI in water and mixing them under stirring, adding GA for cross-linking and casting the mixture on a Teflon plate to prepare a PVA/BPEI membrane (100 μm). The composite membrane was quaternized using benzyl chloride in methanol and then soaked in 1 M KOH to exchange the ions. Poly(diallyldimethylammoniumchloride) (PDDA) is a quaternized copolymer, water-soluble, and highly conductive. The significant impact of PDDA led to its composite with PVA (PVA/PDDA) via thermal and cross-linking methods.259 The membranes are then soaked in 2 M KOH to exchange the ions, and this could be a low-cost and straightforward process. Nevertheless, thermal and chemical cross-linking methods that require many steps are time-consuming.
In recent years, imidazolium ionic liquids (IILs) have gained considerable attention due to their valuable structural modifications, ease of synthesis, and high ionic conductivities similar to other QAs. As a result, Park et al. synthesized typical IILs, 1-butyl-3-vinylimidazolium bromide (C4VIBr), which was functionalized with styrene (PIS) and blended with PVA for membrane (e.g., PISPVA) fabrication.101 The first step of the process consisted of synthesizing C4VIBr using 1-vinylimidazole and 1-bromobutane in the presence of methanol using the reflux method. In the process of polymerization, the obtained C4VIBr was functionalized with styrene (PIS) using DMF (dissolving agent) and AIBIN (initiating agent). The composite membrane was obtained by mixing PIS with PVA in DMSO and then casting it on the glass. The prepared membrane was soaked in 1 M KOH to produce AEMs (55 μm), showing a low-cost and straightforward process.
Similarly, Kim et al. studied the dual-functional (cross-linker and quaternization) characteristics of (3-acrylamidopropyl)trimethylammonium chloride (APTA)-integrated poly(ethylene glycol)diacrylate (PEGDA) (PEGDA-APTA) by photopolymerization (ultraviolet rays).260 The photopolymerized membrane was soaked in 1 M NaHCO3 to replace the Cl− ions with HCO3− ions. In another study, quaternized poly(4-vinylbenzylchloride-styrene) composite membranes (QP(VBC-St)) were developed using the polymerization method.261 The mixture of 4-vinylbenzyl chloride (VBC), styrene (St), and AIBN solutions was first added to the autoclave and heated at 90 °C under an oil bath to prepare the precipitation (P(VBC-St)). In a subsequent step, P(VBC-St) was dissolved in NMP, followed by TEA for quaternization (QP(VBC-St)). To prepare the membrane (70 μm), the solution was cast onto a Petri dish. Afterward, it was soaked in 1 M KOH to replace the Cl− ions with OH− ions, this process demonstrates that it is a simple and low-cost process, with only one concern: the time required to synthesize and fabricate the membrane. In the same way, Ye et al. prepared QP(VBC-St) AEM using the same method.262 Additionally, they doped some radical inhibitors to improve the chemical properties of the membrane during the membrane process. Recently, commercial membrane polyethylene (LDPE) with a thickness of 60 μm was modified with 4-vinylbenzyl chloride (VBC) under UV irradiation, followed by amine activation.132 The DABCO-functionalized membrane (LDPE-g-VBC/QA) was soaked in 1 M KOH to form OH− ions for further use.
Poly(vinyl benzyl chloride) (PVBC) naturally contains ether-free alkyl chains and chloromethyl classes in its structure and, therefore, avoids using hazardous chloromethyl ether agents for the chloromethylation process. Thus, PVBC-pyrrolidinium (MPy) and PSF (PVBC/MPy/PSF-c) were recently composited using different cross-linkers to improve the membrane structure and alkaline stability.135 Briefly, PVBC and Mpy were mixed and dissolved in DMAc. Then, the dissolved chromomethylated PSF and one of the bifunctional tetra amines (diamines) cross-linkers were added to the PVBC-Mpy solution. Further, the solution was cast on the glass to prepare the cross-linked PVBC-PSF membrane. The same procedure was followed for other seven types of bifunctional tetraamine cross-linkers to prepare seven types of PVBC-PSF membranes. Among them, the cross-linkers such as 1,4-bis(2-methylimidazol-1-yl)butane, N,N,N′,N′-tetramethyl-1,6-hexanediamine, and 4,4′-trimethylene-bis(1-methylpiperidine) used to prepare cross-linked PVBC-PSF may show better performance for the water electrolysis application. The diamines-based cross-linkers have difunctional monomers that contain tertiary amine classes. They are generally involved in crosslinking the membranes because of their ease of availability and rapid reaction.
In another study, only N,N,N′,N′-tetramethyl-1,6-hexanediamine (TMHDA) as a cross-linker was used to cross-link the PVBC-Mpy-PSF membrane (PVBMPy-c-PSF).263 PVBC has proven challenging to fabricate AEMs due to its brittleness. Therefore, it requires further polymer support to achieve flexibility. Fig. 13a shows a schematic representation of the cross-linked structure of PVBMPy-CL-PSF. Poly(vinyl acetal) (PVAc) was used as the supporting polymer due to its flexibility, film-forming ability, and cross-linker. PVAc was synthesized by the acetalization reaction using PVA and dissolved in DMSO under heating at 90 °C with 4-dimethylaminopyridine (DMABA) used as QAs and HCl.264 The synthesized PVAc and PVBC were dissolved separately in NMP. A mixed dissolved solution was cast onto the glass and dried to make the composite membrane (PVAc-c-PVBC). MIm and CH3I were used to soak the composite membrane for the Cl− ion form, and then 1 M KOH was used to form the OH− ions (Fig. 13b). The process showed a cheap and straightforward protocol to synthesize AEMs.
![]() | ||
Fig. 13 (a) Schematic depiction of crosslinked PVBMPy-CL-PSF, reproduced with permission from ref. 263, Copyright 2021, Elsevier. (b) The schematic process of PVAc-c-PVBC/OH membranes, reproduced with permission from ref. 264, Copyright 2014, Elsevier. (c) Schematic diagram of sandwiched-porous PBI AEM, reproduced from ref. 266, Copyright 2015, Royal Society of Chemistry. (d) Schematic illustration of quaternization of cellulose nanocrystal (CNC), reproduced with permission from ref. 267, Copyright 2018, American Chemical Society. |
Wang et al. synthesized crosslinked SEBS-poly(biphenyl piperidine) (PBP) composite AEMs (PBP-c-SEBS) by dissolving chloromethylated SEBS in chloroform and dissolving PBP in DMF, mixing them both, and heating at 45 °C for 60 h.265 The chloromethylation agent 1,4-dichloromethoxybutane (BCMB) was used in this investigation. Finally, the mixed solution was cast on the glass and dried, followed by TMA quaternization. Afterward, 1 M KOH was used to soak the composite membrane to form OH− ions. This type of composite process is straightforward to fabricate the AEMs. Diminishing the hydrogen permeability and enhancing the membrane's mechanical properties could be developed by a sandwiched type membrane. Accordingly, Zeng et al. developed a sandwiched porous (sp-PBI) membrane prepared by the method of pore-forming.266 PBI solution was first dissolved in dibutyl phthalate (DBP), then cast on glass, and cured at various temperatures (90, 120, and 160 °C) to prepare the porous PBI (sp-PBI) membrane (Fig. 13c). Then, the PBI solution was sprayed on the dried p-PBI and heated using a hot plate to prepare a sandwiched PBI (sp-PBI) membrane. It could be a cheap and time-saving process.
A renewable polymer material, cellulose has been widely used as a membrane in the separation process. Additionally, cellulose has more hydroxyl groups and can be functionalized with cations for ionic conduction in AEMs fabrication. Therefore, Cheng et al. recently synthesized cellulose crystals modified with QAs and functionalized on quaternized PPO to fabricate AEMs (QCNC/QPPO).267 Cellulose was quaternized using NMB and extra CH3I was added to convert the amine to QAs-cellulose (QC). Fig. 13d shows a schematic illustration of the quaternization of CNC (QCNC). On the other hand, brominated PPO was quaternized by NMB and extra added TMA. The composite membrane was developed by mixing QC and QPPO in the presence of DMSO and then cast on the glass. The membrane was soaked in 1 M NaOH to form the OH− ions. Bacterial cellulose membrane (BCM) is a biological membrane with abundant hydroxyl functional groups favoring the surface modification processes with other polymers for AEMs fabrication. Therefore, Zou et al. studied the purified BCM composite with poly(diallyl dimethyl ammonium chloride) (PDDA) using the impregnation method to produce AEMs (BCM/PDDA).268 The PDDA-modified BCM was cross-linked with GA in the presence of acetone. Later, the composite membrane was soaked in 3 M KOH for ion exchange. The process could be simple and easy, but the BCM must be purified before use.
There has been an increase in the use of polymer fillers recently because their high ammonium content throughout the structure results in an increase in OH− transport compared to inorganic fillers. Therefore, Zhang et al. synthesized poly(vinyl imidazole) (PVI) spheres quaternized with different cations followed by chitosan (QPVI/CS) composites for the film-forming and enhance the mechanical properties to make quaternized polymer sphere AEM.138 PVI particles were prepared by polymerizing vinyl imidazole (having excellent reactive activity and high N content) and other reagents in the presence of acetonitrile (Fig. 14a). PVI particles were quaternized (QPVI) by various QAs agents (allyl chloride, chlorobutane, benzyl chloride, chlorodecane, and ethyl chloroformate) in alcohol under a reflux method. After, they were treated with 0.5 M KOH to form the OH− ions. The composite chitosan with respective QPVIs was prepared by mixing acetic acids at 80 °C for 5 h, followed by GA cross-linking under continuous stirring. Casting the cross-linked blended solution on the glass plate and drying the membrane (60–65 μm) was used to prepare it. Lastly, all membranes were soaked in 0.5 M KOH to form OH− ions.
![]() | ||
Fig. 14 (a) (i and ii) Preparation of QPSs and the fabrication of its membrane, reproduced with permission from ref. 138, Copyright 2016, American Chemical Society. (b) Schematic chemical structure of the fabricated TPPVBN30 membrane, reproduced with permission from ref. 269, Copyright 2013, Elsevier. (c) The chemical structure for crosslinking acylated PPO (XAc-PPO) synthesis, reproduced from ref. 270, Copyright 2020, MDPI. |
On the other hand, the reinforcement technique enhances the membrane stability. The porous PTFE is considered a supporting polymer substrate due to its excellent mechanical strength, chemical and thermal stability, swelling stability, cheapness, and significant thinness. Reinforcement can reduce the swelling of the membranes in a water medium and efficiently improve the stability of the membranes. The PTFE-based reinforcement technique is one of the highly effective approaches to enhance the membrane's mechanical strength. Generally, PVBC becomes highly brittle after casting and is highly soluble in hot water. Thus, it needs an additional supporting polymer substrate to enhance the mechanical strength and a cross-linker for improving the chemical stability. Accordingly, Zhao et al. studied mechanically strengthened PVBC, which was developed by reinforcing it on the PTFE membrane and cross-linked by diethylamine (DEA) to enhance the chemical stability of the membrane (e.g., TPPVBN30).269 Initially, the cross-linked PVBV was quaternized by TMA. The quaternized PVBC was dissolved in DMF to cast on the PTFE substrate to obtain the membrane, and finally, it was soaked in 1 M KOH to exchange the ions. Fig. 14b shows the schematic molecular structure of the TPPVBN30 AEM.
Pore-filled membrane (P-F) is an excellent method to improve the mechanical properties using porous substrates and enhance the wettability of the membrane. Due to phase separation, P-F membranes can be fabricated using a hydrophobic substrate with chemical stability. A thin and uniform structure is needed to prepare the P-F membrane for better application. In this way, Son et al. developed a thin, uniformly structured cross-linked acylated PPO on a polyethylene substrate to prepare a P-F membrane (Q-Ac-PPO/PE) using a centrifugal machine.270 The synthesis of acylated PPO (Ac-PPO) initially used aluminum chloride and 6-bromohexanoyl chloride in the presence of 1,2-dichloroethane (Fig. 14c). Furthermore, an Ac-PPO solution dissolved in tetrahydrofuran (THF) was cross-linked with TMHDA, and the cross-linked solution was pore-filled on the PE substrate using centrifugal machines. The ion-conducting membrane was made by further soaking the thin and uniformly structured P-F membrane (30 μm) in TMA. Lastly, the P-F membrane was immersed in 1 M KOH to form the OH− ions.
Another work that used solvent-free techniques to fabricate the BPPO/VBC/DVB AEMs on reinforcing PETEX substrate was studied by Wu et al.271 The procured purified BPPO and benzoyl peroxide was dissolved in the mixture of 4-vinylbenzyl chloride and divinylbenzene (DVB). The solution was cast onto the PETEX substrate and sandwiched between PVC sheets. The peeled membrane from PVC sheets shows 80 μm; the flexible membrane was quaternized by TMA and soaked in 2 M NaOH to form the hydroxyl ions on the membrane. In another work, PBI nanofiber was synthesized by electrospinning using procured PBI and LiCl in the presence of DMAc.272 Later, synthesized bromomethylated mTPBr was dissolved in NaH to cast on the glass. The prepared PBI nanofiber was placed on the mTPBr solution and dried to obtain the reinforced (PBI-mTPBr) membrane (50 μm), followed by quaternization using TMA for 7 days. The healthier covalent bonds can be formed between the PBI nanofiber and mTPBr polymer and then react with the PBI nanofiber amines. After, 1 M KOH was used to soak the membrane to exchange the ions. Table 4 details the synthesis and fabrication of a range of heterogeneous and homogeneous-based hydrocarbon AEMs for easy understanding.
AEMs | Copolymer/monomer | Process | BA | CMA | CL | QAs | DA | CS | IE | Cost | Ref. |
---|---|---|---|---|---|---|---|---|---|---|---|
a AEMs = anion exchange membranes; BA = brominating agent; CMA = chloromethylation agent; CL = cross-linker; QAs = quaternization agent; DA = dissolving agent; CS = cast substrate; IE = ion exchange. | |||||||||||
PPO/AGO | PPO | Cross-linking and Menshutkin reaction | NBS | — | — | TMA | NMP | Glass | 1 M KOH | Moderate | 242 |
ASU-GO/PIPPO | PPO | Cross-linking | NBS | — | — | N-Methyl piperidine | Ethanol | Silicon wafer | 1 M KOH | High | 243 |
QPEI/f GO/PVA | PEI | Cross-linking | — | — | — | Bromoethane | DMSO | Glass | 0.1 M NaCl | Moderate | 244 |
Poly (VSi-co-QVBC)/PVA/fGO | Poly (VSi-co-QVBC) | Polymerization | — | — | — | N-Methyl morpholine | DMSO | Polyethylene | — | Moderate | 94 |
GO@PDA-ASE | Commercial AEM used | Electrodeposition | — | — | — | — | — | — | — | Low | 246 |
DGO/CMPSF | PSF | Polymerization | — | CME | TMEDA | TMA | DMAc | — | 1 M KOH | High | 245 |
PDA/GO or TiO2/AEM | Commercial AEM used | Dip coating | — | — | — | — | — | — | 0.01 N NaNO3 | Low | 247 |
Sustainion® X37-50/ZrO2 | Commercial AEM used | Sol–gel | — | — | — | — | — | — | 1 M KOH | Low | 248 |
PVA/Im-AESP | Silica precursor | Reflux | — | — | — | MIm | — | Teflon plate | 1 M NaCl | Low | 249 |
PSF/MgAl LDH | PSF | Chloromethylation | — | (CH3)3SiCl | — | TMA | DMF | Petri dish | 1 M KOH and KHCO3 | Low | 134 |
QCS/PVA-LDH | Chitosan | Solution-blending method | — | — | GA | GTMAC | Acetic aqueous + hot water | Glass | 1 M KOH | Low | 250 |
PSF-ImOH/Zn2+ | PSF | Soft template | — | — | — | Imidazolium | NMP | Glass | 1 M KOH | Low | 251 |
ABS/PS/AC | ABS/PS | Phase-inversion | — | — | — | — | THF | Glass | 2 M NaCl | Low | 252 |
PTFE/Ni–Fe | PTFE | Pore-filling | — | — | — | — | — | — | 1 M KOH | Low | 253 |
Q-PEEK-PBI | PEEK/PBI | Menshutkin reaction | CMOE | — | MPi | DMSO | Glass | 1 M KOH | High | 254 | |
VImPPO/PVA | BPPO/PVA | Blending | NBS | — | 1,8-Octanedithiol | VIm | DMSO | Glass | 1 M NaOH | Moderate | 128 |
PEG/PPO | BPPO | Blending | NBS | — | — | DEB and MPi | NMP | Glass | 1 M KOH | Low | 255 |
QPPO/PEG | BPPO | Blending | NBS | — | — | TMA | NMP | Glass | 1 M NaOH | Low | 256 |
PVAf/ABPBI | PBI | Impregnation | NBS | — | GA | TMA | NMP | Glass | 3.1 M KOH | Moderate | 131 |
QPVBC-PVA | VBC | Blending | — | — | PVA | TM | DMSO | PTFE | 1 M KOH | Low | 257 |
PVA/BPEI | PEI | Blending | — | — | GA | Benzyl chloride | Water | Teflon | 1 M KOH | Moderate | 258 |
PVA/PDDA | PDDA | Blending | — | — | GA | — | Water | Petri dish | 2 M KOH | Low | 259 |
PISPVA | Styrene | Polymerization | — | — | — | — | DMSO | Glass | 1 M KOH | Moderate | 101 |
PEGDA-APTA | PEGDA | Polymerization | — | — | — | — | — | — | 1 M NaOH | Low | 260 |
QP(VBC-St) | P(VBC-St) | Polymerization | — | — | — | TEA | NMP | Petri dish | 1 M KOH | Low | 261 |
QP(VBC-St) | P(VBC-St) | Autoclave | — | — | — | TEA | NMP | Flat dish | 1 M KOH | Low | 262 |
LDPE-g-VBC/QA | PE and VBC | Blending | — | — | — | Dabco | Diphenyl ether/benzophenone | Petri dish | 1 M KOH | Moderate | 132 |
PVBC/MPy/PSF-c | PVBC and PSF | Polymerization | — | — | Various tertiary amines | — | DMAc | Petri dish | 1 M KOH | Moderate | 135 |
PVBMPy-c-PSF | PVBC and PSF | Polymerization | — | — | TMHDA | — | DMAc | Petri dish | 1 M KOH | Low | 263 |
PVAc-c-PVBC | PVA and PVBC | Polymerization | — | — | — | DMABA | NMP | Glass | 1 M KOH | Low | 264 |
PBP-c-SEBS | PBP and SEBS | Polymerization | — | BCMB | TMA | — | DMF | Glass | 1 M KOH | High | 265 |
sp-PBI | PBI | Reflux | — | — | — | — | DBP | Glass | 6 M KOH | High | 266 |
QCNC/QPPO | Cellulose and PPO | Polymerization | NBS | — | — | TMA | DMSO | Glass | 1 M NaOH | Low | 267 |
BCM/PDDA | Cellulose and PDDA | Impregnation | — | — | GA | — | — | — | 3 M KOH | Low | 268 |
QPVI/CS | Vinyl imidazole and CS | Polymerization | — | — | GA | Different QAs used | Acetic acid | Glass | 0.5 M KOH | Moderate | 138 |
TPPVBN30 | VBC | Polymerization | — | — | DEA | TMA | DMF + ethanol | Glass | 1 M KOH | Low | 269 |
Q-Ac-PPO/PE | PPO | Polymerization | — | — | TMHDA | TMA | THF | PE | 1 M KOH | Low | 270 |
BPPO/VBC/DVB | BPPO and VBC | Blending | — | — | — | TMA | DVB | PETEX | 2 M NaOH | Moderate | 271 |
PBI/mTPBr | PBI | Electrospinning | — | — | — | TMA | THF | Glass | 1 M KOH | Low | 272 |
![]() | ||
Fig. 15 (a) Radiation grafting of amino (DMAEMA) and succeeding protonation to synthesize the AEM (ETFF-g-PMAOEDMAC), reproduced with permission from ref. 273, Copyright 2007, Elsevier. (b) Preparation of pyridinium-functionalized fluorinated OH−-form membranes, reproduced with permission from ref. 275, Copyright 2015, Elsevier. (c) Quaternized Nafion and Aquivion perfluorinated membranes, reproduced from ref. 276, Copyright 2020, MDPI. |
Recently, Fang et al. synthesized pyridinium-functionalized fluorinated AEMs.275 The pyridinium salt of the pyridine (4-vinylpyridine (4-VP)) has good chemical and thermal stability compared to conventional QA systems. Thus, they synthesized the copolymer named poly(HFMA-co-4VP-co-BMA) (PHVB) by the copolymerization method with 4-vinylpyridine (4VP), hexafluorobutyl methacrylate (HFMA), and butyl methacrylate (BMA) in the presence of DMF (Fig. 15b). After dissolving the copolymers in DMF, they were cast on the glass to prepare the membrane, and then soaked in 1 M HCl, followed by 1 M NaOH. Time-savings and minimal chemicals could make this a low-cost and straightforward process to manufacture fluorocarbon-based polymers.
Perfluorinated ionomers known as Nafion membranes are categorized for their excellent chemical, thermal, and mechanical properties. Due to these advantages, Jung et al. used Nafion with an additional form of sulfonyl fluoride (Nafion-SO2-F) for the AEMs fabrication.277 The membrane was reacted with 1,4-dimethylpiperazine (DMP) to enable the QAs cation site. Then, the DMP-functionalized Nafion (Nafion-DMP) was treated with 1 M KOH for ion exchange. For a comparison study, trimethylchlorosilane (TMCS) was employed to chloromethylate PSF (CMPSF). After it had been dissolved in DMF and reacted with DMP (PSF-DMP), a membrane was created using solution-casting. Lastly, ion exchange in the PSF-DMP membrane was produced using 1 M KOH. In another work, Lee et al. introduce the QAs (TMA) into the Nafion-SO2F and Aquivion-SO2F using the Menshutkin reaction under the autoclave method.276 The quaternized perfluorinated polymers were dissolved in NMP to cast on the glass to prepare membrane (50 μm), followed by the membrane soaked in 1 M NaCl to exchange the ions (Fig. 15c). Using the N–S reaction, fluorinated poly(aryl ether oxadiazole) (FPAEO) was synthesized using 2,5-bis(2,3,4,5,6-pentafluorophenyl)-1,3,4-oxadiazole, 4,4′-(hexafluoroisopropylidene)diphenol, and 4,4′-isopropylidenebis(2,6-dimethylphenol) and the prepared polymer was crosslinked with TMHDA.278 This work used NBS to brominate FPAEO, and N-methyl imidazole was used to quaternized FPAEO (FPAEO-Im) and obtain a linear structured AEM (L-FPAEO-MIM). The FPAEO-Im solution was cross-linked with TMHDA (C-FPAEO-MIM), followed by casting on glass and drying to obtain the membrane, and ion exchange on the membrane was done by soaking in 0.5 M NaCl (Fig. 16a). The study used TMHDA to act as a cross-linking agent and enhance the dimensional stability, and imidazole has five-ring heteroaryl molecules that enhance the chemical stability (Fig. 16b). The imidazole is more significant than the QAs group. For this study, the preparation of fluorinated polymer at low temperature (0 °C) is the only concern.
![]() | ||
Fig. 16 (a) Schematic chemical structure for the synthesis of linear FPAEO-MIM (L-FPAEO-MIM) and (b) cross-linking process of FPAEO-MIM (C-FPAEO-MIM), reproduced with permission from ref. 278, Copyright 2013, Elsevier. |
Jiang et al. studied a stretchable bication cross-linker, which was then incorporated into the poly(arylene alkylene) side chain to fabricate cross-link structured AEMs (m-XPTPA-2N).279 Initially, the ionomer (7-bromo-1,1,1-trifluoroheptan-2-one and bromopentyl-tethered poly(meta-terphenylene alkylene)) was prepared using a simple procedure, followed by dissolving in NMP (Fig. 17a). After that, the solution was cast onto the glass, and after drying, the membrane with ions (Br− and I−) was washed with 2 M NaCl to replace the existing ions with Cl− ions. A di-tertiary amine, N,N-dimethyl-6-(N′,N′,N′-trimethylammonium) was then used to quaternized the polymer solution, followed by the addition of bi-cation cross-linker N,N,N′,N′-tetramethyl-1,6-hexanediamine. The polymer solution was poured onto the glass to prepare a thin membrane (30 μm). The membrane processes of this type show a moderate cost, a lengthy protocol, and is time-consuming. One polymer backbone known as polyfluorene (PF) has high alkaline stability because it does not have aryl ether bonds in its structure, and the long alkyl-functionalized cation groups are thermally stable compared to benzyl-substituted cation groups under an alkaline environment. Thus, Lee et al. synthesized different fluorine-based ionomers (PFB, PFF, and PFBFF) with long-alkyl functionalized cations on the side chain of the polymer (PF) backbones.280 They used TMA for quaternization. Quaternized PFs (Q-PFs) ionomers in the bromide form were dissolved in DMF, cast on the glass, and dried. The membranes were ion-exchanged by dissolving in 1 M NaOH to obtain the membranes. Even though this procedure utilizes the synthesis of numerous ionomers accordingly, quaternization uses high amounts of chemicals and is intensive with time.
![]() | ||
Fig. 17 (a) The chemical structure for the synthesis of cross-linked poly(meta-terphenylene alkylene) ionomers (m-XPTPA-2N), reproduced with permission from ref. 279, Copyright 2021, Elsevier. (b) The chemical structure for synthesizing the ionomer (PPO-22-3QA8F), reproduced from ref. 282, Copyright 2019, Royal Society of Chemistry. (c) Scheme for the synthesis of TQ-PDBA membrane, reproduced from ref. 283, Copyright 2020, Royal Society of Chemistry. |
Polyolefin, or polypropylene (PP), is highly durable in alkaline environments. These polyolefin-based AEMs are produced with the hot-pressing method. However, their use has been restricted to water electrolysis because of their poor solubility in organic media. Furthermore, hot-pressing polyolefin based on QAs results in thick membranes and the high-temperature operation of functionalized polyolefin, which may cause chemical degradation. Therefore, Zhang et al. recently developed fluoropolyolefin-based AEMs by tetraalkylammonium-modified fluoropolyolefin using the direct Ziegler–Natta (Z–N)-catalyzed copolymerization method.281 A fluorinated monomer called 1-(but-3-en-1-yl)-4-fluorobenzene (BFB) was synthesized and then copolymerized with 11-bromo-1-undecene (BUD) to produce PBFB–Br. Inserting the fluorine atom in the polyolefin enhances the solubility of the as-synthesized copolymers in an organic solvent, enabling the quaternarization of the polyolefin with TMA (PBFB-QAs). The obtained PBFB-QAs were dissolved in DMF to cast on the glass and dried (60 °C), followed by soaking the membrane (50 μm) in 1 M NaOH to exchange the ions.
Li et al. investigated the perfluorinated side chain tethered to the backbone of PPO polymers (e.g., PPO-22-3QA8F).282 In their study, they examined how the fluorination and tri-QAs side chain influenced the structure of the polymer (Fig. 17b). The fluorinated PPO membrane (20–30 μm) was developed by solution casting, and 2 M NaCl was used to submerge the membrane and create a Cl− form, followed by 1 M NaOH to convert the Cl− form into OH− for future usage. This study involves synthesizing many monomers and ionomers in a multi-step procedure, which can be expensive. Introducing a fluorinated moiety into the polymer backbone can improve the membranes' functioning. Thus, Gao et al. studied the hydrophobic fluorine polymer with specific functional sites as the main chain.283 They synthesized an ionic liquid bearing a long, flexible alkyl-spaced cation as the side chain. Later, partial diallyl bisphenol A poly(arylene ether) (PDBA) polymer was synthesized in the presence of decafluorobiphenyl, diallyl bisphenol A poly(arylene ether), and 4,40-(hexafuoroisopropylidene)diphenol with DMAc. The synthesized PDBA was brominated (PDBA-Br) using NBS and benzoyl peroxide in the 1,1,2,2-tetrachloroethane medium. Then, PDBA–Br was crosslinked with ionic liquid (NQQN) in the presence of NMP, followed by the solution cast on the glass to obtain the membrane (30–50 μm). The ion exchange of the ionic liquid cross-linked PDBA membrane was done in 1 M NaOH to replace the Br ion with OH− ions. Through this process, fluorinated AEMs (e.g., TQ-PDBA) can be synthesized at a low cost and in a short amount of time (Fig. 17c). Another group synthesized novel cross-linked AEMs with fluorine (hexafluorobenzene)-included backbone and hydrophobic multi-cation side chains (1-(N,N-dimethylamino)-6,12-(N,N,N-trimethylammonium)dodecane bromide).284 This work introduced the CC bond into the AEMs to form an end-group cross-linked structure. However, fabricating this type of end crosslinked structured AEMs (e.g., PHFB-VBC-DQ) requires a complicated synthesis of multi-cations and a multi-step procedure that results in high chemical consumption and is time-consuming.
Recently, high free volume with fluorinated AEMs (e.g., NCBP-Im) was synthesized by Zhang et al.285 AEMs have a bend-twist type block copolymer structure containing imidazolium ions, and this type of structure enables more OH− ion transport and relatively high conductivity (Fig. 18). This study synthesized two oligomers, such as fluorine-terminated PAEs (FPAE) and hydroxyl-covered poly(phthalazinone ether)s (HPPE). Both oligomers were coupled via a chemical reaction in the presence of DMAc, potassium carbonate, and toluene at 40 °C for 5 h to obtain a block copolymer precipitate called NCPB, followed by chloromethylation using CMME (cm-NCPB). The cm-NCPB was quaternized using 1,2-dimethylimidazole (DMIm) in the presence of NMB. The prepared membrane (40 μm) was soaked in 1 M NaOH for ion exchange. This type of AEM fabrication requires many chemicals and is a lengthy process, which is likely costly.
![]() | ||
Fig. 18 The chemical structure for the synthesis of NCBP-Im, reproduced with permission from ref. 285, Copyright 2018, Elsevier. |
In another study, an octopus-structured side chain was grafted on poly(arylene piperidinium) AEMs.288 In this type of membrane, the “octopus head” is considered a side chain comprised of a β-cyclodextrin, a rigid structure, and “arms” are considered to be long piperidinium ionic liquids, which are flexible. In this, specific ionic liquid (1-bromohexyl-N-methylpiperidinium) (bpIL) and the polymer poly(biphenyl piperidinium) (PBP) were synthesized separately. During PBP preparation, 1,1,1-trifluoroacetone, trifluoromethanesulfonic acid, and 1,1,1-trifluoroacetic acid was added for fluorination. After fluorination, PBP was treated with mono-6-I-β-CD (PBPCD), then dissolved in NMP with PBPCD and sodium hydride. After adding bpIL slowly to the above solution, the mixture was called PBPCL. In membrane fabrication, PBPCL was dissolved in NMP and cast onto glass. The obtained membrane was soaked in 1 M NaOH to form OH− ions. Due to many chemicals and two or three steps, this process could be moderately expensive and time-consuming.
The introduction of pendant aliphatic groups/chains of QAs on fluorinated polymers enhanced the properties of AEMs. Thus, Mahmoud et al. studied the effect of aliphatic ammonium groups on perfluoroalkyl (PAF) membranes.289 The perfluoroalkylene monomer was used to synthesize PAF copolymer, which was then chloromethylated by CMME (CMPAF). Separately, various QAs TMA (trimethylammonium), DMHA (dimethylhexylamine), DMIm, and DMBA (dimethylbutylamine) were functionalized on CMPAF to obtain quaternized PAF (QPAF). In addition, QPAF was dissolved in NMP and cast on the glass to get the membranes (45–100 m), which were then soaked in 1 M KOH for ion exchange. According to this study, aliphatic cation DMBA-functionalized PAFs have excellent properties compared to other QAs-functionalized PAFs. Side-chain (S-C)-based AEMs have gained much attention because of their high conductivity. The S-C-based copolymer offers a more efficient ion-conducting channel than the main-chain (M-C) type polymer. Thus, Wang et al. recently studied a functionalized claw-type based side chain on a partially-fluorinated-poly(arylene ether) (F-PAE) membrane using nucleophilic substitution polycondensation to improve the alkaline and conductivity stability.290 Decafluorobiphenyl and 4,40-(hexafluoroisopropylidene)diphenol were used to synthesize F-PAE using toluene and DMAc, and later, multiple sites were introduced into F-PAE using 4-triphenylmethylphenol (BPM), DMAC, and K2CO3. Further, the active sites-incorporated F-PAE was brominated (F-PAE-Br) using 6-bromohexanoyl chloride (BHC) in the presence of chloroform, and later, they added trifluoroacetic acid and triethyl silane. Then, it was quaternized by MPip in the presence of NMP to cast on the glass and dried to obtain the membrane (e.g., FPAE-3B-X-PD (30 μm)), followed by the membrane being soaked in 1 M KOH.
Aromatic main chains with aryl-ether groups are chemically unstable in alkaline environments. To address this problem, Ren et al. synthesized polyfluorene (PBF) AEMs using an acid-catalyzed X polycondensation-based aryl-ether free polymer backbone, highly adaptable alkyl side chains, and cationic groups.291 Initially, 9,9-bis(6-bromohexyl)-9H-fluorene (BHF) was synthesized by the fluorine alkylation through C–H alkylation (Fig. 19a). Later, polyfluorene (PF) homopolymer and PBF copolymers were synthesized using BHF by the simple chemical method and F–C polycondensation, respectively. A nucleophilic substitution reaction with the quaternization agent MPi converted the polymer side chains containing bromohexyl (PBF–Br) groups to cationic, resulting in high alkaline stability. The quaternized polymer (dissolved in NMP) was cast onto the glass and dried. Ion exchange was accomplished by immersing the PBF membrane (50 μm) in 1 M NaOH to obtain the PBF-OH membrane.
![]() | ||
Fig. 19 (a) Synthetic route for preparing ether-bond-free polyfluorene (PBF–OH), reproduced with permission from ref. 291, Copyright 2019, American Chemical Society. (b) The chemical structure for synthesizing QPAF4-Cx-pip membranes, reproduced from ref. 292, Copyright 2019, Royal Society of Chemistry. (c) The synthesis route of the PFOTFPh-Cx-TMA membrane, reproduced from ref. 100, Copyright 2020, Royal Society of Chemistry. |
Koronka et al. synthesized partially fluorinated perfluoroalkyl and fluorine-based copolymers.292 In this synthesis, the authors developed the combined alkyl spacer and long-lasting hetero-cyclo aliphatic quaternized ammonium groups to improve the performance of AEMs (e.g., QPAF4-C3-pip (40–60 μm)), and Fig. 19b showed the chemical structure for the synthesis of QPAF4-Cx-pip membranes. However, this process is expensive due to the synthesis of various fluorene-based monomers, followed by polymerization, which involves the use of many chemicals and is time-consuming. Generally, the flexibility and mechanical strength of the membrane are enhanced by the chain entanglement. This chain entanglement can be increased with the increase in the molecular weight of the polymer. Therefore, Miyanishi et al. developed poly(fluorene-alt-tetrafluorophenylene) AEMs (e.g., PFOTFPh-Cx) with high molecular weight synthesized by the C–H activation technique.100 The synthesized copolymer has no ether link in this study, and the backbones are entirely aromatic (Fig. 19c). The continuous inflexible aromatic structure of copolymers affects the membrane's flexibility. Thus, Liu et al. synthesized a sequence of PAEK copolymers bearing a long alkyl and a solidly quaternized carbazole pendant to make the membrane (e.g., PAEK-HQACz) highly flexible.110 Adding a long alkyl chain onto the pendants enhances the flexibility of polymers, boosts the chain entanglement, and increases the membrane's strength. In this work, long alkyl tetra-benzyl carbazole-mediated PAEK copolymer was synthesized using monomers (4-methoxy)phenylhydroquinone and 9-(6-bromohexyl)-3,6-bis(3,5-dimethylphenyl)carbazole, followed by brominated (PAEK-Br) by NBS and benzoyl peroxide. The dissolved PAEK-Br using 1,1,2,2-tetrachloroethane (TCE) was cast on glass, dried, and then soaked in TMA for quaternization. The quaternized membrane was then soaked in 1 M KOH to exchange the ions in the membrane; thus, the total membrane process involved several steps and was very labor-intensive.
Poly(aryl piperidinium) (PAP) is a high molecular weight polymer, rigid, ether-free, hydrophobic backbone with excellent thermal and chemical stabilities. Compared to other QAs, guanidinium salt has excellent structural stability, thermal stability, chemical stability, and hydroxide conductivity. Therefore, Wang et al. synthesized high dimensional stability, ether-free, and hydrophobic guanidinium-functionalized PAP to fabricate AEMs (e.g., G-PPTPT-n:m).293 The monomer (poly(piperidine trifluoro-phenylethyl p-terphenyl)) was initially synthesized using piperidine-4,4-diolhydrochloride, 2,2,2-trifluoroacetophenone, and p-terphenyl in the presence of DCM, followed by the addition of trifluoromethanesulfonic acid and trifluoroacetic acid for the incorporation of fluoride. Later, the guanidinium salt-functionalized poly(piperidine trifluoro-phenylethyl p-terphenyl) copolymer was synthesized using the above monomer under a nitrogen atmosphere. The copolymer was quaternized using N,N-diisopropylethylamine (DIPEA), and it was dissolved in NMP to cast on the glass and dried. The membrane was soaked in 1 M NaOH to exchange the ions. However, this process needs a nitrogen environment and low temperature to synthesize monomers and copolymers. Recently, a highly flexible cross-linker bication was prepared into a side chain of the alkali-stable polymer backbone poly(arylene alkylene) to improve the properties of membranes, such as mechanical and oxidative stability.279 First, a polymer was synthesized with bromopentyl-tethered poly(meta-terphenylene alkylene) (e.g., m-PTPABr). Then, this polymer was dissolved in NMP, adding N,N-dimethyl-6-(N′,N′,N′-trimethylammonium)hexane iodine for the quaternization. Also, partial quaternization on the polymer was carried out using TMHDA under continuous stirring to cast on the glass and then dried to form the bication cross-linked m-PTPABr (e.g., m-XPTPA40-2N). Finally, the membrane (30 μm) was soaked in 1 M NaOH to exchange the ions, and the process of this type of AEMs could be inexpensive and straightforward.
At higher pH, piperidinium cations are stable against hydroxide ions attack, and such a blend of a rigid polymer backbone with an alkali-stable ion could provide excellent AEMs for various applications. Accordingly, Novalin et al. developed a highly stable poly(terphenyl piperidinium-co-trifluoroacetophenone) (PAPQ83).294 In this study, CH3I was used as a quaternization agent. The developed membrane (PAPQ83) thickness was 50 μm, followed by 1 M NaOH was used to soak the membrane to exchange the ions (Fig. 20a). Recently, the hydrophilic nature of hexyl QAs and the highly hydrophobic nature of fluorine-switched pendants were hosted into the commercially available side chains of SEBS (e.g., HQA-Fn-SEBS).139 In a simple process, bromohexyl fluorobenzoyl was functionalized with SEBS in the presence of chloroform, and the solution was cast on the glass and dried to prepare the membrane (30–40 μm). Further, the membrane was quaternized using TMA, and then the dried membrane was soaked in 1 M KOH, followed by a bicarbonate solution to exchange the ions (Fig. 20b). N-Cyclic-based cations known as piperidinium have gained considerable attention due to their lower costs and superb chemical stability. Accordingly, a partially fluorinated and cluster form of piperidinium quaternized PAP was synthesized to fabricate AEMs (e.g., 3QPAP).295 The fluoride components such as 2,2,2-trifluoro-1-phenylethanone (TFPE), trifluoroacetic acid (TFA), and trifluoromethanesulfonic acid (TFMSA) were used for the fluorination. Then, PAP was quaternized (QPAP) using 1-(6-bromohexyl)-1-methylpiperidinium bromide (BrPD) and N,N-diisopropylethylamine (DIPEA) in the presence of DMSO. Further, QPAP was dissolved in DMSO to cast on the glass, and the prepared membrane was soaked in 1 M NaOH to exchange the ion. The prepared AEMs may be chemically stable; however, the process requires multiple steps, has high chemical costs, and is time consuming.
![]() | ||
Fig. 20 (a) The chemical structure for the synthetic process of PAPQ83, reproduced from ref. 294, Copyright 2021, Elsevier. (b) Synthesis route of hexyl quaternary ammonium-tethered fluorobenzoyl-embedded SEBS (HQA-Fn-SEBS), reproduced with permission from ref. 139, Copyright 2022, Elsevier. (c) The schematic chemical structure for synthesizing the P(4PA-co-2PA) ionomers, reproduced with permission from ref. 300, Copyright 2022, Elsevier. |
Side-chain cations such as N,N-dimethylpiperidinium and 6-azoniaspiro[5.5]undecane showed outstanding resistance to degradation in higher alkaline environments.296,297 Therefore, Pan et al. synthesized two partially fluorinated models of poly(arylene piperidine) AEMs by polycondensation and then functionalized mono- and di-cation side-chains, N,N-dimethylpiperidinium (e.g., qPBPip) and 6-azoniaspiro[5.5]undecane (e.g., qPBmPip), respectively.297 Partial fluorination on the polymer backbones used fluorine precursors such as 2,2,2-trifluoroacetophenone and trifluoromethanesulfonic acid. The synthesized brominated poly(arylene piperidine) was quaternized in the NMP/DMSO solution mixture. Later, the quaternized polymer was dissolved in DMSO to cast on the glass and dried to prepare the membrane. This study used 1 M NaBr solution to soak the membrane for 7 days for ion exchange. The AEMs bearing dicationic side chains in the polymer backbone could be helpful in high-level ion conductivity. In another work, a side-chain-grafted poly(biphenyl N-methyl piperidinium) AEMs containing bis-cation and a long methylene spacer were synthesized using polymerization (e.g., QPBPipXAc).298 Biphenyl, 1-methyl-4-piperidone (MPPD), and TFAc were used to prepare the copolymer (poly(biphenyl N-methyl piperidinium)) in the presence of trifluoroacetic acid, trifluoromethanesulfonic acid, and dichloromethane. The synthesized copolymer was quaternized in 1-(6-bromohexyl)-1-methylpiperidinium bromide (6-Br-MPD), potassium carbonate, and DMSO. The obtained membrane was soaked in 1 M NaCl for the ion exchange. This type of AEMs fabrication could be a low-cost, time-saving, and straightforward procedure.
Ethylene tetrafluoroethylene (ETFE) is a commercially available copolymer with a chemical structure containing polytetrafluoroethylene and polyethylene components, which means it has physicochemical properties of both hydrocarbons and fluorocarbons, contributing to its excellent thermal and chemical properties. Further, this partially fluorinated membrane can also be fabricated through radiation grafting techniques, which use the inherent properties of the polymer to improve the mechanical and durability properties. Biancolli et al. prepared ETFE membranes and irradiated them with electron beams in different atmospheres (air and nitrogen).299 In the presence of 1-octyl-2-pyrrolidone (a surfactant), the irradiated ETFE was grafted with VBC. The grafted membrane was quaternized with MPip (ETFE-MPRD) and then soaked in 1 M NaCl to exchange the ions. Irradiation may be effective but is costly compared to synthesizing AEMs using polymerization.
Recently, different monomers and copolymers were synthesized to develop the highly stable poly(p-quaterphenylene alkylene) polymer structure for the fabrication of AEMs.300p-Quaterphenylene was copolymerized into poly(arylene alkylene) ionomers and pendant pentyltrimethylammonium units. The polymerization of p-quaterphenyl used with trifluoromethyl ketone (7-bromo-1,1,1-trifluoroheptan-2-one (BTFH)) endow the polymer with limited solubility; therefore, the copolymerization of the polymer with biphenyl was used accordingly to enhance the solubility. The p-quaterphenylene group is used in this study to increase the strength of the polymer. The obtained quaternized bromomethyl tethered copolymers (Q-P(4PA-co-2PA)) ionomer using TMA was dissolved in NMP to cast on the glass. The prepared membrane (30 μm) was soaked in 2 M NaCl to exchange the ions (Fig. 20c). It is relatively expensive to fabricate AEMs in this work due to the synthesis of different monomers and copolymers with high chemical consumption. Recently, Zhou et al. studied a quaternized poly(fluorene piperidine) (PBBP), which is prepared through acid-catalyzed (A-C) polycondensation reactions.301 They used 1,1,1-trifluoroacetone (TFA), 9,9-dimethylfluorene (DF), and MPPD to prepare a PBBP via A-C polycondensation. The PBBP was crosslinked by various cross-linking agents such as TMHDA, APTES, and BPPO to prepare different AEMs correspondingly. 1 M NaOH was used to exchange the ion for all membranes. In this process, the preparation of AEMs involves various cross-linkings that could be costly. Poly(olefin) is one of the polymer backbones containing no aryl ether bonds and is chemically stable in alkaline environments. Comparing fluorinated poly(olefin) membranes to non-fluorinated membranes, poly(olefin)-based membranes demonstrate significant dimensional stability.302 Moreover, the fluorinated poly(olefin) polymer (e.g., F20C9N) membranes (70 μm) show high conductivity due to the fluorinated monomer, which allows for improved hydroxide separation and transport.
Pendant ammonium and long alkyl groups have provided significant chemical stability to the membrane. Therefore, recently, the effect of various ammonium head groups (dimethylhexyl, heterocyclic N-methylpiperidinium, dimethylbutyl, and dibutylmethyl) on quaternized partially fluorinated bis(3-chlorophenyl)perfluorohexane (PAF) AEMs (40–50 μm) was investigated by Mahmoud et al.303 They synthesized monomer (2-(3-(2,5-dichlorophenyl)propoxy)tetrahydro-2H-pyran) to synthesize copolymer PAF, followed by bromination. The brominated PAF was quaternized using aliphatic amines (QPAF-C3), and the obtained QPAF-C3 was dissolved in NMP to cast on the glass to prepare the membrane, followed by soaking in 1 M KOH for ion exchange. Various ammonium head groups introduced in partially fluorinated PAF could be chemically stable. Still, this synthesis process is rather time-consuming and expensive due to the need for monomers and copolymers synthesized with many chemicals. Similarly, Miyake et al. reported five types of ammonium groups (MIm, trimethylamine (TMA), N-butyldimethylamine, pyridine, and DMIm) quaternized on partially fluorinated poly(arylene ether) (PAE) polymer for the alkaline testing stability of the membranes.304 This study prepared block copolymer PAE by polymerizing hydrophobic oligomers and functionalizing their hydroxyl groups. Decafluorobiphenyl (DFBP) and hexafluorobisphenol A (HFBPA) were used for partial fluorination on PAE. In this study, CMME used to chloromethylate PAE (CMPAE) is not environmentally safe. The authors synthesized different cation ammonium groups using a simple reaction and then quaternized CMPAE (QPAE). In this investigation, the polymer was dissolved in 1,1,2,2-tetrachloroethane (TCE) or DMAc before being cast onto a glass plate to create a membrane that was 50 μm thick, and the fabrication process of these partially fluorinated AEMs could be simple and low-cost.
Poly(arylene ether) bearing ammonium units-substituted fluorine molecules show high conductivity and chemical and thermal stability. Thus, it was prepared with poly(phenylene ether) ionomer to achieve high chemical stability.306 This ionomer was prepared as a dumpy block ionomer to exploit the ionic matter. The monomers (2,5-dichloro-4-phenoxybenzophenone (2,5-DCPBP) and 4-chloro-4′-fluorobenzophenone (CFBP)) and trimer were synthesized to prepare the copolymer. The copolymer (poly(p-phenylene) and poly(arylene ether) (PPAE)) were synthesized by nickel-catalyzed polymerization method. Later, CMME was used to chloromethylated the co-polymer (Ch-PPAE). Then, Ch-PPAE was dissolved in DMAc and cross-linked with 4,4′-diaminobenzophenone (DBP), followed by casting on the glass to prepare the membrane, followed by quaternization using TMA and exchange the ion on the membrane (e.g., QCPPAE) using 1 M NaOH.
The polymer intrinsic microporosity (PIM) with a vastly rigid and uneven backbone structure has a higher free volume. The PIM block offers a free volume in the membrane structure, thus facilitating hydroxide ion transport. Accordingly, Gong et al. developed a sequence of composite membranes based on QAs-polysulfone (QAs-PSF) and a free volume of PIM (PSF-PIM) to fabricate AEMs.307 The composite membrane demonstrated a higher conductivity than the bare PSF membrane. Using bisphenol A (BA), potassium carbonate, and 4,4′-difluorodiphenylsulfone (DFDS) in the presence of DMAc, they synthesized fluoro-terminated block co-polymer PSF (F-PSF), which was maintained at 140 °C in an N2 atmosphere. In addition, OH-terminated PIM (OH-PIM) was synthesized using tetrafluorophenedionitrile, potassium carbonate, and 5,5′,6,6′-tetrahydroxy-3,3,3′,3′-tetramethylspirobisimdane in DMF, and the same reaction environment was followed at 70 °C. In a subsequent step, the synthesized OH-PIM and F-PSF were mixed with potassium carbonate in the presence of DMAc to form the composite PIM-PSF. The composite was chloromethylated using CMOE, then quaternized using N,N-dimethyl-butylamine. The quaternized blended polymer solution was cast on the glass to prepare membrane Q-DMBA-PIM-b-PSF and soaked in 1 M NaOH. This type of membrane could be an effective solution, but it involves lengthy and time-consuming processes, and using chloromethylation agent CMOE is toxic. Different homogeneous and heterogeneous-based fluorinated AEMs are shown in Table 5, along with their synthesis and fabrication cost analyses.
AEMs | Copolymer/monomer | Process | BA | CL | QAs | DA | CS | IEA | Cost | Ref. |
---|---|---|---|---|---|---|---|---|---|---|
a AEMs = anion exchange membranes; BA = brominating agent; CL = cross-linker; QAs = quaternization agent; DA = dissolving agent; CS = casting substrate; IEA = ion exchange agent. | ||||||||||
ETFE-g-PDMAEMA | Commercial membrane used | γ-Ray irradiation | — | — | DMAEMA | — | — | — | High | 273 |
ETFE-g-VBC | Commercial membrane used | γ-Ray irradiation | DABCO | TMA | — | — | 1 M KOH | High | 274 | |
PHVB | HFMA and BMA | Polymerization | — | — | 4-VP | DMF | Glass | 1 M NaOH | Low | 275 |
Nafion-DMP | Nafion | Polymerization | — | — | DMP | DMF | Glass | 1 M KOH | Low | 277 |
PSF-DMP | PSF | Polymerization | — | — | DMP | DMF | Glass | 1 M KOH | Low | 277 |
Nafion-SO2F | Commercial membrane used | Menshutkin reaction | — | — | TMA | NMP | Glass | 0.5/1 M NaOH | Low | 276 |
Aquivion-SO2F | Commercial membrane used | Menshutkin reaction | — | — | TMA | NMP | Glass | 0.5/1 M NaOH | Low | 276 |
C-FPAEO-MIM | FPAEO | N–S reaction | NBS | TMHDA | MIm | DMF | Glass | — | High | 278 |
m-XPTPA-2N | m-PTPABr | Polymerization | — | — | Various QAs | NMP | Glass | 1 M NaOH | High | 279 |
Q-PFs | PF | Polymerization | — | — | TMA | DMF | Teflon | 1 M NaOH | High | 280 |
PBFB-QAs | BFB–Br | Polymerization | BUD | — | TMA | DMF | Glass | 1 M NaOH | High | 281 |
PPO-22-3QA8F | PPO | Polymerization | NBS | — | PMDETA | DMF | Glass | 1 M NaOH | High | 282 |
TQ-PDBA | PDBA | Polymerization | NBS/BPO | NQQN | — | NMP | Glass | 1 M NaOH | High | 283 |
PHFB-VBC-DQ | PHFB | Polymerization | NBS | — | N,N-DQA | DMSO | Glass | 2 M KOH | High | 284 |
NCBP-Im | FPAE and HPPE | Polymerization | — | — | DMIm | NMP | Glass | 1 M NaOH | High | 285 |
msQPBI | PBI | Polymerization | — | — | N-MIm | DMSO | Glass | 0.01 M KOH | High | 286 |
XMePh-Z | PES | Polymerization | NBS | — | TMA | DMF | Petri dish | 1 M NaOH | High | 287 |
PBPCL | PBP | Polymerization | — | bpIL | — | NMP | — | 1 M NaOH | Moderate | 288 |
QPAF | PAF | Polymerization | — | — | TMA | NMP | Glass | 1 M KOH | Moderate | 289 |
FPAE-3B-X-PD | FPAE | Polycondensation | BHC | — | MPi | DMF | Glass | 1 M KOH | High | 290 |
PBF | BHF | Polycondensation | DBH | — | MPi | NMP | Glass | 1 M NaOH | Moderate | 291 |
QPAF4-C3-pip | PAF-pip | Polymerization | — | — | Piperidinium | DMAc | Glass | 1 M KOH | High | 292 |
PFOTFPh-Cx | — | C–H activation | — | — | TMA | DMSO | — | 1 M NaOH | High | 100 |
PAEK-HQACz | PAEK | Polymerization | NBS | — | TMA | TCE | Glass | 1 M KOH | High | 110 |
G-PPTPT-n:m | PPTPT | F–C reaction | — | — | DIPEA | NMP | Glass | 1 M NaOH | Low | 293 |
m-XPTPA40-2N | m-PTPA | Polymerization | — | — | TMHDA | NMP | Glass | 1 M NaOH | High | 279 |
PAPQ83 | PAP | F–C reaction | — | — | CH3I | DMSO | Glass | 1 M NaOH | Low | 294 |
HQA-Fn-SEBS | Commercial membrane used | Solution-casting | — | — | TMA | CHCl3 | Glass | 1 M KOH | Low | 139 |
3QPAP | PAP | F–C reaction | — | — | BrPD/DIPEA | DMSO | Glass | 1 M NaOH | Moderate | 295 |
qPBPip or qPBmPip | PBPip or PBmPip | Polycondensation | — | Various cationic side-chains | DIPEA | DMSO/NMP | Petri dish | 1 M NaOH | High | 297 |
QPBPipXAc | PBPipXAc | Polymerization | DBH | MPPD | 6-Br-MPD | DMSO | Glass | 1 M NaCl | High | 298 |
ETFE-MPRD | Commercial membrane used | Irradiation | — | VBC | MPip | — | — | 1 M NaCl | High | 299 |
Q-P(4PA-co-2PA) | Quaterphenylene alkylene and biphenylene alkylene | Polymerization | — | BTFH | TMA | NMP | Glass | 1 M NaOH | High | 300 |
PBBP | TFA and DF | Polycondensation | — | TMHDA or APTES or BPPO | CH3I | NMP | — | 1 M NaOH | Moderate | 301 |
F20C9N | 4-(4-Fluorophenyl)-1-butene | Polymerization | BUD | — | TMA | Tetrahydrofuran | — | 1 M NaOH | Moderate | 302 |
QPAF-C3 | PAF | Polymerization | Carbon tetrabromide | — | Aliphatic amine | NMP | Glass | 1 M KOH | High | 303 |
QPAE | PE, DFBP and HFBPA | Polymerization | — | — | TMA | TCE or DMAc | Glass | 1 M KOH | Moderate | 304 |
PBI–DIm–Si | DAB and 6FDA | Polymerization | DBH | CPTMOS | DIm | DEA & water | Petri dish | 1 M KOH | High | 305 |
PBI/IL-GO | BCPH and DB | Polymerization | — | IL-NH2/GO | MIm | DMSO | PTFE | 2 M KOH | Low | 137 |
QCPPAE | 2,5-DCPBP and CFBP | Polymerization | — | DBP | TMA | DMAc | Glass | 1 M NaOH/1 M Na2SO4 | High | 306 |
Q-DMBA-PIM-b-PSF | BA and DFDS | Polymerization | — | — | DMBA | DMAc | Glass | 1 M NaOH | High | 307 |
Most conventional membrane synthesis protocols involve multi-step synthesis, which raises the overall cost. AEMs can also be fabricated from commercially available membranes using radiation grafting, an emerging method of synthesizing their functionalized forms. Accordingly, polyethylene with a thickness of 15 μm of commercially available AEMs was obtained from Goodfellow, UK, which was further irradiated and then grafted to achieve a sub-30 μm membrane by Wang et al.311 Specifically, the electron beam was generated at 4.5 MeV using a Dynamatron unit (Fig. 21a). This activates the membrane by producing peroxy groups covalently attached to the polymer chains. Dry ice was used after irradiation to store the membrane at −40 °C. The irradiated membranes were placed in glass vessels and further soaked in a mixture of vinylbenzyl chloride (VBC) and 1-octyl-2-pyrrolidone solutions, followed by N2 purging for 2 h with the vessel sealed tightly. Afterward, the grafting procedure was performed by heating at 40 °C for 6 h. Then, amination was achieved by soaking the membrane in TMA solution for 24 h and washing it with high-purity water. In the end, chloride (Cl−) ions were formed by the membrane being soaked in NaCl solution, and finally, the final grafted membrane containing the chloride ions was stored in pure water until use. It was found that the modified membrane was highly robust. However, using high-end e-beam sources and after irradiation, the membrane stored at shallow temperatures (−40 °C) remains a costly and time-consuming process. Similarly, Mahmoud et al. studied a commercial ethylene-tetrafluoroethylene (ETFE) membrane irradiated using 60Co radiation as a source.312 After irradiation, the membrane was treated with vinylbenzyl chloride under an argon atmosphere, soaked in TMA for 24 h, washed with water, and dried to obtain irradiated and grafted AEMs in the Cl− ion form. As a result of using high-end radiation sources to irradiate the membrane, processed commercial membranes appear to be moderately expensive.
![]() | ||
Fig. 21 (a) Schematic representation of the synthesis direction of the LDPE-AEM, reproduced from ref. 311, Copyright 2018, Royal Society of Chemistry. (b) Synthesis of PES-S-OH membrane, reproduced with permission from ref. 313, Copyright 2016, Elsevier. (c) Spacer-containing quaternized PPO AEMs' preparation directions, reproduced with permission from ref. 113, Copyright 2021, Elsevier. |
Similarly, Marino et al. explored the bromide form of bare membrane FAA-3 obtained from Fumatech GmbH. Anion exchange was accomplished by treating the membrane with 1 M Na-(Cl, F, OH) and NH4I for 48 h.314 Recently, Hossain et al. developed sulfonium-based AEMs from commercially available poly(ether sulfone) (PES) by initially introducing the chloromethyl groups, followed by functionalization with dimethyl sulfide.313 The corresponding AEMs were obtained by anion exchange with 1 M KOH solution concentration, and the chloromethylation reaction was carried out with chloromethyl methyl ether and thionyl chloride (Fig. 21b). The catalyst ZnCl2 was used as the Lewis acid. The impact of side-chain structures and spacer on membrane performance was studied recently by Li et al., who developed a series of PPO AEMs with different side-chain structures and spacer.113Fig. 21c depicts the preparation method of several side-chain structures on quaternized PPO AEM and the presence of a spacer group. In another work, alkaline doped polymers possess high hydroxide conductivity and chemical stability in an alkaline solution. Therefore, a commercially available polymer solution diluted with DMAc was recently cast on glasses and dried at different temperatures (90, 120, 170, and 190 °C).315 The dried membranes were then doped with KOH to study the membrane properties further. This type of polymer solution obtained from the market to prepare AEMs could be cheaper than processing commercial AEMs with several steps.
Most organic polymer membrane synthesis involving highly toxic organic solvents may harm the environment and humans. On the other hand, chitosan's naturally occurring hydroxyl and amine groups enable it to be quaternized without chloromethylation. In addition, poly(vinyl benzyl chloride) (PVBC) polymer contains naturally occurring chloromethyl classes and avoids the use of toxic chloromethyl ether agents. Therefore, future research focusing on these polymer types would be highly advantageous and cost-effective and reduce the use of hazardous organic solvents in membrane preparation. A hybrid method is typically used to produce composite membranes with effective ion channels. The uniform dispersion of inorganic particles may be difficult in organic/inorganic composite membranes, but this composite type would be advantageous for anti-fouling. In addition, future research should analyze the properties of inorganic particles composited with organic polymers for application purposes. Due to the membranes' physical, thermal, and chemical properties, fluorine incorporation in the hydrocarbon may be advantageous. However, the fluorination-based synthesis process is expensive due to the increased number of synthesis stages and the use of costly compounds. Compared to hydrocarbon membranes, fluorinated membranes are covered in fewer articles in the scientific literature. Therefore, future research on fluorinated polymers should focus on low-cost processes, and an effective fluorinated membrane is urgently required. Reinforced membrane fabrication is an excellent method for enhancing the membrane's structural stability; however, the reinforced membrane's thickness must be regulated, and the porous substrate must be coated uniformly. There are commercially available membranes, but they require additional processes such as surface modifications, quaternization, and ion exchange. Using commercial membranes increases the cost of this procedure. Ultimately, it was determined that numerous polymer synthesis processes involve multiple stages and could be time-consuming and costly. Therefore, it is recommended to synthesize a polymer with few straightforward steps as this is the most cost-effective method.
![]() | ||
Fig. 23 SAXS profiles of (a) the iodide form of PQP-100 AEM, reproduced with permission from ref. 235, Copyright 2022, Elsevier. (b) PPO-based AEMs, reproduced with permission from ref. 113, Copyright 2021, Elsevier. (c) PBCL (0.69–1.48)/IL(1.55) membranes, reproduced with permission from ref. 288, Copyright 2021, Elsevier. (d) Iodide form of QAFPK-1-n-E/C (n = 8, 6, and 5), reproduced with permission from ref. 133, Copyright 2021, American Chemical Society. (e) Iodide form of n-propyl (nP), isopropyl (iP), or cyclohexyl (CH)-functionalized PPO membranes, reproduced with permission from ref. 232, Copyright 2021, Elsevier. (f) Bare and dual-alkyl side chains-grafted QAs-PPO, reproduced with permission from ref. 316, Copyright 2021, American Chemical Society. |
Recently, Li et al. studied the PPO AEMs performance by adjusting the side-chain structures and the impacts of alkoxyl or alkyl extenders and spacers.113 They analyzed the various AEMs (DQEC, DQ, SDQ, and SDQEC) by SAXS that did not appear as apparent diffraction peaks due to not having a microphase separation. Among AEMs, SDQEO (alkoxyl extender and dual QAs were connected to the main chain along with a spacer) shows diffraction peaks due to having the presence of MSM (Fig. 23b). For this morphology, SDQEO could be a high ion conductivity. The piperidinium treated with aryl ether-free-based AEMs (PTP-75, 85, and 90) shows no significant ionic peaks due to the indiscriminate distribution of piperidinium groups on the main polymer backbones.234 Another study developed the microphase separation by the poly(biphenyl piperidinium) (PBP)-based AEMs (PBPCL-0.69, 1.29, 1.48 and PBPIL-55) structure containing the hydrophilic octopus-like side chain and the hydrophobic fluorinated moieties.288 The PBP containing a combination of side chains β-cyclodextrin (β-CD) and piperidinium ionic liquids (PBPCL) shows larger d-spacing and better MSM than PBP containing only piperidinium ionic liquids (PBPIL). According to this study, AEMs analyzed by SAXS have higher d-spacing, indicating a larger ion cluster and significant MSM (Fig. 23c).
Similarly, electrospinning and hot pressing methods obtained well-fabricated MSM structure for QAs-functionalized polyketone-based AEMs (QAFPF-1-n-E), and the calculated d-space values were obtained in the range of 9.1–10.8 nm.133 However, in the case of QAs-functionalized polyketone (QAFPK-1-n-C) produced by casting, followed by the stripping method, low d-values (5.7–6.3 nm) were obtained, indicating that it could be a less microphase separated structure, and these higher and lower d-values specify that the ionic cluster of QAFPK-1-n-E is better than that of QAFPK-1-n-C (Fig. 23d). AEMs developed with aryl-ether-free backbones, long flexible alkoxy-containing bis-piperidinium side chains, and hydrophobic alkyl chains spacer (PISBr-NON) show characteristic diffraction peak at 1.6 nm−1, indicating that this membrane has a good microphase separated structure.227
QAs-containing PPO-isopropyl AEMs showed a broad scattering peak (ionomer peak) at 1.9 nm−1 due to the growth of ionomer aggregation.232 No ionomer peak was observed for PPO containing n-propyl and cyclohexyl AEMs (Fig. 23e). The case of hydrophilic dual alkoxy side chains-grafted QAs-PPO shows more substantial diffraction peaks when compared to dual-alkyl side chains-grafted QAs-PPO and bare QAs-PPO.316 Additionally, the d-spacing values of alkoxy side chains-grafted QAs-PPO are higher (11.9–13.8 nm) than others, indicating a well-long-scale ordered MSM (Fig. 23f). The reason could be that the grafting hydrophilic side chains (alkoxy) can stimulate MSM and the growth of ion conduction paths. Recently, PEG with molecular weight of 2000 with various percentages (1–5%) was composited with Q-PPO, showing that the broad diffraction peak was increased as the PEG weight percentage increased.255 These diffraction peaks indicate that all AEMs have MSM; thus, PEG 2000 could effectively induce the MSM in the membranes while compositing with PPO (Fig. 24a). The rigid backbone polymer structure and using a high dosage of guanidinium salt in the main polymer backbones do not give rise to significant ionomer peaks.293 Thus, the mechanical properties of the polymer backbone and the dosages of salt used for quaternization are essential in forming MSM.
![]() | ||
Fig. 24 SAXS spectrum of (a) various percentages (1–5%) of the PEG (MW 2000) composite with QPPO, reproduced with permission from ref. 255, Copyright 2021, Elsevier. (b) Uncross-linked (VImPPO-PVA) and crosslinked VImPPO-PVA AEMs, reproduced with permission from ref. 128, Copyright 2021, Elsevier. (c) Quaternized NAPEK-PVP-x membranes, reproduced with permission from ref. 239, Copyright 2021, Elsevier. (d) SAXS scattering profile of NPBI-QAs and NPBI-QAs-by, reproduced with permission from ref. 188, Copyright 2021, Elsevier. (e) SAXS spectrum of QA-PE-CO and QA-PE-NH2 membranes, reproduced with permission from ref. 189, Copyright 2021, American Chemical Society. (f) SAXS profile of TQ-PDBA-x, reproduced from ref. 283, Copyright 2020, Royal Society of Chemistry. |
Similarly, the ionomer peak (q = 0.18 nm−1) was observed for the photoinitiator-mediated crosslinked (1,8-octanedithiol)1-vinylimidazole-functionalized PPO composite with PVA (crosslinked VImPPO-PVA), indicating that this AEM has moderate MSM.128 Additionally, the d-spacing value found for these AEMs was 34.9 nm. However, these crosslinked AEMs do not show a long-level ordered MSM based on the obtained SAXS peak (Fig. 24b). In contrast, no more ionomer peak was found in the uncrosslinked VImPPO-PVA. Therefore, the 8-octanedithiol crosslinker could have the effect of producing an MSM in AEMs. Another study shows three types of crosslinked naphthalene-poly(vinylbenzyl chloride)-based block polymer AEMs with the ratio of 1:
6, 1
:
8, and 1
:
10, named NAPEK-PVP-6-Q4, NAPEK-PVP-8-Q4, and NAPEK-PVP-10-Q4, respectively.239 Among AEMs, a more substantial scattering diffraction peak (q) at 0.52 nm−1 with a d-spacing value of 12.08 nm was obtained for NAPEK-PVP-10-Q4 AEMs (Fig. 24c). It is revealed that the membrane containing naphthalene and a high ratio of hydrophilic polymer (1
:
10) backbone stimulates the ion clusters for membrane performance compared to low ratios of naphthalene and PVP (1
:
6 and 1
:
8) used to prepare AEM. The N-alkyl substitution on naphthalene-based PBI-QAs (e.g., NPBI-QAx-By) AEMs shows better microphase separation, confirmed through scattering ionomer peaks (2.7–2.9 nm−1) with corresponding d spacing values (2.15–2.31 nm).188 At the same time, no scattering peaks were observed when QAs separated from the naphthalene-based PBI polymer through an alkyl linker (e.g., NPBI-QAs). It could be the reason for the inter-chain hydrogen bond between the polymer backbones (NPBI) in the alkyl linker procedure to separate the QAs from the polymer, hindering the ionic aggregation (Fig. 24d). However, the N-alkyl substitution reaction greatly promotes the QAs on NPBI, and this procedure avoids the inter-chain hydrogen bond between the polymers. Thus, microphase separation was enhanced by the ionic aggregation through N-alkyl substitution reaction through QAs-functionalized NPBI AEMs.
Another study shows that C–NH2 groups/linkages in the main polymer backbone generate hydrogen bonds in the AEMs, which could enhance the conductivity and mechanical strength. Accordingly, QAs-poly(arylene ether ketone) functionalized with C–NH2 (QA-PE-NH2) shows more substantial ionomer peaks (higher ion clusters) when compared to QA-PE-functionalized CO (QA-PE-CO).189 The higher ion clusters in QA-PE-NH2 AEMs were found to be hydrophilic networks (hydrogen bonds) that are spread throughout the membranes and could stimulate the growth of ion clusters to larger clusters (Fig. 24e). The effect of fluorine in SEBS membranes functionalized with hexyl QAs was studied for the MSM.139 The fluorine content in the polymer backbone clearly showed high scattering peaks (high ionomer aggregation), confirming the long-scale ordered MSM in the membranes compared to hexyl QAs-functionalized SEBS. The increasing hydrophilic regions and d-spacing values with increased ionic group content were found in the membranes (3Q-PAP), attributed to the aggregate ionomer for the conductivity.295 Additionally, the d-spacing is minimal, and no scattering peak was detected for the low ionic content of AEMs. A highly enriched MSM/ion cluster was found in ionic liquid-crosslinked AEMs with fluorine-based polymer main backbones and long flexible hydrophilic multi-cation as side chains (e.g., TQ-PDBA-x).283 The obtained various scattering peaks “q” position (1.029–0.729 nm−1) and “d” spacing values (6.1 to 8.6 nm) suggest that the invented AEMs have microphase separation and relative ion domain, which has a size that can help the formation of the ion cluster for the enhancement of AEMs performance for ion conductivity (Fig. 24f).
![]() | ||
Fig. 25 AFM images of (a) PBP-67, PTP-83, and PQP-100 AEMs, reproduced with permission from ref. 235, Copyright 2022, Elsevier. (b) Surface smoothness of PPO-CH, PPO-iP, and PPO-nP membranes, reproduced with permission from ref. 232, Copyright 2021, Elsevier. (c) PPO AEMs with tethered SDQEO, SDQ, SDQEC, DQEO, DQ, and DQEC side-chains, reproduced with permission from ref. 113, Copyright 2021, Elsevier. |
Adjusting the degree of quaternization could affect obtaining a well-arranged microphase morphology. Accordingly, the poly(vinyl chloride) (PVC) membrane was quaternized with triethylenetetramine (TETA) with the degree of quaternization (4, 8, and 14 h) observed under moderate conditions.129 At this point, a minimum of 8 h is needed for the quaternization of PVC to obtain a well-separated microphase morphology. Therefore, PVCs treated for 8 and 14 h quaternization time with TETA show better MSM than that treated for 1 h quaternization time, as confirmed through the AFM images. Another researcher observed that no MSM formation was found in the study of adjusting the viscosity of piperidinium-grafted aryl ether-free AEMs.234 It could be the reason for the unintentional distribution of the hydrophilic nature of piperidinium clusters on the backbones. Therefore, the adjusting viscosity does not benefit producing MSM in membranes.
Similarly, well-phase separated structured hydrophilic–hydrophobic morphology was observed for the linear (e.g., AEM-LCP) and star-shaped polymer (e.g., AEM-SCP) backbone polyisoprene-based AEMs.127Fig. 26a shows various scale-resolution AFM images of the hydrophilic (black region) and hydrophobic (white region) separation of the AEM-SCP and AEM-LCP. Polyketone-based nanofiber AEMs (e.g., QAFPK) show good mobility and conductivity, enabling high interactions between cationic functional domains (hydrophilic) and hydrophobic polymer backbone that allow a healthy MSM structure. The authors claimed that the high conductivity polymer possesses well MSM.133 Higher hydration could increase the smoothness on the surface of the PPO membrane, and the polymer chain (alkoxy) mobility tends to enhance the MSM.316 At this point, the smaller hydrophilic clusters (dark regions) were observed for the membranes, e.g., QPPO17 and 16C25-10C25. On the other hand, alkoxy with extra side chains grafted on the polymer backbones enhanced the hydrophilic cluster formation for the 16C25-3O25 membrane. In contrast, the extender and other side chains led to a larger hydrophilic cluster observed in the 3O25-16C25 membrane (Fig. 26b). Therefore, alkoxy groups and extender with side chains could enhance the higher MSM. Well-preserved MSM structures are visible in hydrophilic (dark) and hydrophobic (bright) sections, such as the polymer backbone (VImPPO) and cross-linkers (1,8-octanedithiol), respectively, in a mechanically strengthened photoinitiator mediated on crosslinked VImPPO-PVA (e.g., sIPN-91).128 However, the VImPPO and PVA composite membrane (e.g., n-64) did not exhibit any such active hydrophobic and hydrophilic areas in the absence of a crosslinker (Fig. 26c). Therefore, the crosslinker could be a crucial component in creating membrane microphase separation.
![]() | ||
Fig. 26 AFM images of (a) AEM-SCP and AEM-LCP membranes with different scale resolutions, reproduced with permission from ref. 127, Copyright 2021, Elsevier. (b) QPPO17, 16C25-10C25, 16C25-3O25, and 3O25-16C25 membranes, reproduced with permission from ref. 316, Copyright 2021, American Chemical Society. (c) AFM images of the crosslinked VImPPO-PVA membrane (sIPN-91) and un-crosslinked VImPPO-PVA membrane (n-64), reproduced with permission from ref. 128, Copyright 2021, Elsevier. |
Recently, Zhou et al. studied MSM on polyfluorene piperidine (PBBP)-based AEMs. To this end, PBBP was crosslinked with DBMHDA, BAPTES, and BPPO in molar ratios of 0.1, 0.2, and 0.3 and then quaternized. PBBP was also studied without crosslinking for comparison sake.301 Unlike BAPTES and BPPO crosslinking, effective MSM was found for a 0.3 molar ratio of DBMHDA-crosslinked PBBP quaternized membrane. AEM based solely on PBBP, however, did not exhibit any MSM. Hexyl QAs functionalized (HQAs) with mono and penta-based fluorobenzoyl modifications on SEBS AEMs recently yielded nicely organized MSM.139 On HQAs-functionalized SEBS, however, MSM was not seen to perform any better. β-Cyclodextrin has a surface with many hydroxyl groups, making it ideal for roping side chains. The hydroxyl groups on the edge of the β-cyclodextrin may aid in causing the ionic groups to aggregate and provide an effective microphase separation (MPS). Therefore, the combination of ionic liquid and β-cyclodextrin-functionalized poly(arylene piperidinium) (PBP) AEM led to higher MSM as compared to only ionic liquid-functionalized PBP.288
It was discovered that adding a fluorinated moiety to the polymer enhanced the membranes' abilities to separate microphases, have minimal swelling, and maintain chemical stability, among other properties. The main chain uses a hydrophobic fluorine-contained polymer with unique functional sites. In the meantime, a multi-cation hydrophilic ionic liquid was used as the side chain to enhance the dimensional stability and ionic conductivity. Accordingly, diallyl bisphenol A poly(arylene ether) partially crosslinked with long flexible multi-cation ionic liquid-prepared AEMs shows high MSM.283 Friedel–Crafts polycondensation with coupling procedure may be used to produce aryl-ether-free QAs polyaromatics and stop the degradation of the polymer. Thus, ether-bond-free polyfluorene (PBF)-based AEMs encompassing stretchy alkyl pendant chains ended with piperidinium cations groups show high chemical stability, high ionic conductivity, and well-ordered MSM structure.291 Self-assembly and microphase separation are caused in the membrane due to hydrophilic differences between the polymer backbone and the long stretchable ionic side chains. Additionally, the formed hydrophilic channels can improve ion conduction, thus adversely impacting the characteristics of AEMs. Therefore, side-chain-type poly(biphenyl-N-methyl piperidinium) (PBPip) AEMs containing bis-cation filaments with an extended stretchy methylene spacer shows high microphase separation, i.e., a separate phase separation morphology in the membrane may be created by including bendable side chains further into the polymer and increasing the concentration of ionic groups on the side chains.298
![]() | ||
Fig. 27 1H NMR spectra of (a) NPBI-QA25 and NPBI-QA25-B75 polymers, reproduced with permission from ref. 188, Copyright 2021, Elsevier. (b) Copolymer PTPip-90, PTP-90 dissolved DMSO-d6/TFA, and PTP-90 dissolved DMSO-d6, reproduced with permission from ref. 234, Copyright 2021, Elsevier. (c) SEBS, and SEBSC polymers, reproduced with permission from ref. 233, Copyright 2021, Elsevier. (d) PSF and CMPSF, reproduced with permission from ref. 202, Copyright 2021, Elsevier. |
Organic functionalization was confirmed for the SEBS polymer, CMME-functionalized SEBS (SEBSC1), and lengthy side chain-functionalized SEBSC (SEBSC6) polymers using deuterated chloroform.233 Accordingly, –CH–, –CH2– groups, and protons from benzene rings were credited to the SEBS polymer (Fig. 27c). Proton peaks from the chloromethyl groups were confirmed at 4.50 ppm, confirming the CMME functionalization with SEBS. Similarly, using deuterated chloroform, CMME functionalization on PSF (CMPSF) confirmed the peak form at 4.64 ppm because of the –CH2– group in chloromethyl methyl from CMME.202 Therefore, it can be assured that –CH2– groups in chloromethyl methyl formed 1H NMR peaks ranging from 4.50 to 4.64 ppm (Fig. 27d).
Further, the PDMB copolymers of –OCH3, –COOCH3, and methyl groups were confirmed by peaks at 3.3, 3.8, and 6.1 ppm, respectively, and in this study, deuterated chloroform was used for dissolving the polymer.144 After chloromethylation (PDMB-Cl), the peak at 4.55 ppm was due to the methyl protons in benzyl chloride from CMME. Later, it was functionalized with piperidinium (PDMB-Pi), which confirms that the proton peaks at 1.4, 1.6, and 1.8 ppm belonged to piperidinium groups in the PDMB membrane. In another study, electrospun modified quaternized polyketone (Q-PK)-based AEM confirmed each step's organic functional process using 1H NMR, and this study also used deuterated chloroform for dissolving the polymer.133 After the Paal–Knorr reaction on PK (FPK), the peak at 5.4 and 4.99 ppm was due to the H2 atoms of the pyrrole structure and the H2 of the secondary carbon associated with pyrrole nitrogen. Later, quaternized FPK was confirmed by the peak at 2.03 ppm due to the methyl proton of QAs. The peak shift from 5.47 to 5.5 ppm was a proton moving from FPK to Q-FPK, which confirmed the QAs functionalization. Further, the new peak at 3.03 ppm was due to the amines of the side chains, and later, it disappeared due to quaternization.
Recently, quaternized poly(4-vinylbenzyl chloride-styrene) (QP(VBC-St)) AEM was assured through 1H NMR, and deuterated DMSO was used for the polymer solubility.261 As a result, the protons of the major backbone alkyl group confirmed a peak at about 2.2 ppm. Moreover, the protons of the chloromethyl group (CH2Cl) were confirmed at 4.6 ppm due to the chloromethylation of the main backbone. Upon quaternization, the peak/chemical shift at 4.4 ppm showed that the methylene protons were linked to the QAs and the main backbone's benzene ring. Yang et al. developed 1-vinylimidazole (VIm)-functionalized BPPO (VImPPO) AEM, and their organic construction was confirmed using 1H NMR. This study used deuterated chloroform for dissolving the BPPO and deuterated DMSO for dissolving the VImPPO.128 Because of bromoethylation, the proton of the brominated –CH2– group was detected at 4.3 ppm, while BPPO was responsible for the methyl proton signal at 2.0 ppm. The imidazolium ring was observed at 7.8, 9.37, and 7.7 ppm following VIm functionalization with BPPO, confirming the functionalization of VIm.
In another study, BPTMA-Br side chain-functionalized PBI (PBI-x%BTMA) membrane was studied, its organic functionalization was confirmed using 1H NMR, and the membrane was dissolved using deuterated DMSO.187 Accordingly, the protons of the benzimidazole and benzene ring of the main backbone were confirmed at 13.2 and 7.5–9.5 ppm, respectively, and the BPTMA-Br cation was identified at peaks ranging from 1.4 to 3.6 ppm (Fig. 28a). The peaks at 3.0 ppm and between 1.0 and 2.1 ppm after BPTMA functionalization with PBI were caused by the methyl and methylene groups in BPTMA. Further, Simari et al. identified the functional organic groups of chloromethylated polysulfone (CMPSU) and noted the chloromethyl and aromatic proton peaks at 4.64 and 7.5 ppm, respectively.134Fig. 28b shows the 1H NMR spectra of CM-PSU. Wang et al. used deuterated DMSO and chloroform to solubilize the synthesized polymers.265 Accordingly, benzyl hydrogen and the benzene ring were identified at 4.51 and 7.46 ppm, respectively, due to the chloromethylation of SEBS, where CH3I reacted with PBP polymer to change the chemical shift in the tertiary and QAs nitrogen, observed at 2.7, 3.5, and 2.2 ppm.
![]() | ||
Fig. 28 1H NMR spectra of (a) BPTMA-Br, PBI-(27–48%)BPTMA, and PBI polymers, reproduced with permission from ref. 187, Copyright 2021, Elsevier. (b) CM-PSU, reproduced with permission from ref. 134, Copyright 2022, Elsevier. (c) 20PDM-2%PEG2000 polymer, reproduced with permission from ref. 255, Copyright 2021, Elsevier. (d) PSF-ImCl/Zn-x, reproduced with permission from ref. 251, Copyright 2021, Elsevier. |
A different study analyzed incorporating PEG into side chain-functionalized Q-PPO (20PDM-x%PEG2000) for the organic structure and used deuterated DMSO to dissolve the polymer.255 To this end, the MPi functionalization of piperidinium groups on BPPO was identified at 3.1 and 2.98 ppm (Fig. 28c). The incorporation of PEG into PPO was identified as a peak shift at 3.5 ppm due to the protons of –CH2CH2O– from PEG. Recently, Zn2+-incorporated imidazolium-functionalized polysulfone (PSF-ImCl/Zn-x) confirms their organic functionalization using 1H NMR, and this study used deuterated DMSO.251 The protons of the imidazolium ring were identified at 7.81 and 7.7 ppm, and these peaks were shifted to 7.65 and 7.56 ppm to incorporate Zn2+ ions into the polymer backbone. Additionally, evidence shows that the imidazolium ring's electron losses caused a higher chemical shift in the 1H NMR spectra (Fig. 28d).
Compared to QAs cations, N-spirocyclic cations often have double-cyclic chemical structures and higher alkali stability. For improved results, polymer scientists concentrate on producing N-spirocyclic-based cations on membranes. As a result, an investigation was conducted on a newly developed hydrophilic N-spirocyclic-based cation and hydrophobic alkyl side chain-functionalized chloromethylated polysulfone.241 The 1H NMR signal at 4.64 ppm was owing to the protons of CH2Cl, which served as a confirmation that polysulfone was cholomethylated. Later, at peaks 2.67, 0.86, and 2.48 ppm, hydrophobic alkyl side chains (n-octylamine and 1,6-hexamethylenediamine) were verified. Additionally, the proton peaks from 2.2 to 1.9 ppm were confirmed for hydrophilic N-spirocyclic cation-functionalized chloromethylated polysulfone, while 3.3 to 2.9 ppm corresponded to 8-hydroxyl-5-azonia-spiro[4.5]decane cations. After quaternization, the methylene protons of –NH2 (2.67 ppm) were shifted to 2.70 ppm. Similarly, the bromomethyl-tethered copolymers (P(4PA-co-2PA)–Br) were analyzed, and the protons of aromatic phenylenes in the p-quaterphenylene of the main backbone were observed at 7.73 ppm.300 The formation of quaternization was evidenced by the fading of the Br–methyl group at 3.33 ppm and the advent of a new peak at 3.07 ppm, corresponding to the methyl groups from trimethylalkyl ammonium.
Lin et al. recently developed membranes made of poly(aryl piperidine) that had been quaternized with clustered cations (3QPAP), and deuterated DMSO was used to make the polymer more soluble.295 Before quaternization, PAP's primary backbone benzene rings were present from 7.0 to 8.0 ppm. The two signals at 2.7 and 3.1 ppm are attributed to methyl groups, while the signal at 8.4 ppm is attributed to the polymer's –NH groups. The aromatic protons signals remained constant after being quaternized with clustered cations (3QPAP), whereas the –NH peak vanished, and additional new peaks were seen due to the quaternization. To analyze their organic structural confirmation, bication cross-linked and quaternized poly(meta-terphenylene alkylene)s (m-PTPA) were utilized, and deuterated CHCl3 was used to determine the polymer's solubility.279 After quaternization and the appearance of additional peaks at 2.9 and 3.0 ppm, which were attributed to the methyl protons of dialkyl and trimethyl alkyl QAs cation group, respectively, the bromomethylene proton of the brominated membrane (m-PTPA) at 3.33 ppm completely vanished. The indications of methylene and methyl protons in the spectrum further supported the conclusion that m-PTPABr was partly quaternized.
Several different weight ratios of the cation 4-dimethylaminopyridine were functionalized with a BPPO membrane, and this was done so that their elemental composition could be determined.224 As a result, the main peaks of Br 3d and N 1s, which represent BPPO and 4-dimethylaminopyridine, respectively, were seen, demonstrating the potential integration of these compounds into PPO (Fig. 29a). After quaternization, the intensity peak of the cation-functionalized BrPPO decreased because of the structural modifications made from benzyl bromide to QAs. The QAs of C4N+ were discovered to be at 400.8 eV, confirming the successful quaternization of the nitrogen element and incorporation of 4-dimethylaminopyridine into the polymer backbone. In the same manner, procured BPPO was crosslinked with tertiary amino(dimethylamine) and quaternized with methyl iodide, and for comparison, uncrosslinked PPO was quaternized with trimethylamine (QPPO).194 XPS showed that the N 1s binding energy differed for QAs in QPPO and QAs in dimethylamine (DMA)-crosslinked QPPO (Fig. 29b). It is evident that the binding energies at the N 1s peak are 401 and 403 eV, which correspond to the nitrogen in the tertiary amine and the nitrogen in crosslinked QAs, respectively. Additionally, the binding energy of DMA-crosslinked BPPO was measured following methyl iodide functionalization, and it was discovered that the N 1s peak at 399.5 eV was caused by tertiary amino, which was converted into QAs. TMA was used to quaternize PPO, whose N 1s peak appeared at 401.5 eV. Therefore, the crosslinking and quaternization of PPO were confirmed by such distinctive peaks at specific binding energies of functional groups.
![]() | ||
Fig. 29 XPS spectra of (a) the BPPO membrane, reproduced with permission from ref. 224, Copyright 2018, Elsevier. (b) High-resolution N 1s spectrum of QPPO, C-TA-PPO, and CQPPO, reproduced with permission from ref. 194, Copyright 2018, Elsevier. (c) Bare PBI and alkaline (6 M KOH)-doped PBI, reproduced with permission from ref. 315, Copyright 2015, Elsevier. (d) PVBC, PVBC-MPy, CMPSF, crosslinked PVBC-MPy-CMPSF (M-2 to M-6#) membranes, reproduced with permission from ref. 135, Copyright 2021, Elsevier. |
To prove that an alkali was functionalized on PBI, the chemical composition of the alkaline-processed PBI polymer before and after was studied by Zeng et al., and the K peaks of alkali were observed at 376.8, 291.8, and 32.8 eV.315 The single and double bond nitrogen (N) in the benzimidazole were represented as the N 1s peak; however, this N 1s was shifted after alkali treatment on PBI, and this peak shift indicated that the benzimidazole segment was deprotonation and obtained a negative charge as a result of delocalization with electrons (Fig. 29c). Similarly, Li et al. demonstrated the structural functionality of the poly(vinyl benzyl methylpyrrolidinium)-CMPSF (PVBC-MPy-CMPSF)-crosslinked dual-functional tertiary amines (diamine) AEMs through XPS.135 N 1s peaks were seen at 400 eV as a result of crosslinkers as well as the insertion of chloromethyl groups in the polymer backbone (PVBC) and MPy-functionalized polysulfone. At 200 and 201.9 eV, respectively, the chloromethyl group, Cl 2p of chloromethylated polysulfone (PSF) and PVBC, was identified (Fig. 29d). Due to the functionalization of PVBC with MPy, the two binding energy peaks of PVBC were moved to a low energy state. After functionalization with MPy, it was found that PVBC's N 1s peak, which results from nitrogen in the pyrrolidinium, is at 402.1 eV. It was proven that diamine was functioning on the PVBC membrane due to the changes in the binding energy values from higher to lower and the appearance of the N 1s peak.
In a different study, Chen et al. investigated the chemical structure composition of Zn2+ ions-incorporated imidazolium-functionalized polysulfones.251 Zn2+ ions were responsible for the peak at 1020 eV, whereas the peak (N 1s) at 401 eV resulted from nitrogen atoms in the imidazolium rings. Ion exchange between Ac− and Cl− ions is another possible reason for these peaks. AEM with dopamine functionalization, which is highly hydrophilic and has negative charges, was produced, and the structural composition of AEM confirms that it was quaternized since it reveals nitrogen elements in the bare AEM owing to QAs.246 The GO-modified AEM survey spectrum shows more oxygen content due to the high oxygen content of GO; however, the nitrogen content was lowered due to GO restricting the penetration depth of the rays falling on base QAs-containing AEM. Due to the presence of nitrogen in dopamine, the dopamine/GO-AEM (dopamine coated on a modified AEM) and the dopamine-AEM membrane exhibit significant nitrogen content. The functionalization of dopamine on the GO-modified AEM was therefore confirmed by increasing the nitrogen in the membrane, and a high nitrogen content in the membrane may be advantageous for high conductivity.
The surface area of cellulose nanomaterial crystals (CNCs) is adequate, lacking mechanical properties and hydroxyl groups. Therefore, they need additional surface modification to improve their mechanical and surface qualities. Accordingly, the structural composition was investigated for the quaternized CNCs (QCNCs), and it was discovered that these QCNCs had N 1s and Si 2p peaks at 399.1 and 101.4 eV due to the nitrogen and silica concentrations in QAs, respectively, while these peaks are absent in CNCs.267 Zeng et al. employed XPS to investigate rGO and PBI composite interaction.266 The N 1s peaks for rGO-modified PBI were identified at 398.0 and 399.9 eV due to double and single-bond nitrogen in the benzimidazole ring of PBI, respectively. The intensity of N 1s peaks was also diminished due to GO's adsorption on PBI. As a result, the presence of N 1s peaks in the PBI affected by rGO proves that GO externally modified the PBI.
Using CH3I to quaternate the membrane, Wang et al. generated multi-cation 1,16-dibromo-5,11-(N,N-dimethylammonium)hexadecane (DBDMAH) ionic liquid-crosslinked poly(aryl piperidinium) AEMs.317 They also prepared quaternized poly(aryl piperidinium) AEMs without cross-linking. The presence of the N 1s peak with a higher intensity than in uncross-linked AEM led researchers to conclude that the DBDMAH-crosslinked AEM had a high nitrogen content. Therefore, the functionalization of multi-cations may increase the nitrogen concentration and, as a result, have the potential to increase the membrane conductivity. Hao et al. developed an AEM based on poly(fluorenyl ether), and they quaternized the membrane using trimethylamine (e.g., QA-OMPFEK).318 In this research, the author used homemade QAs (trimethylamine) for the quaternization of the membrane. The compositional structure of the generated AEM was discovered to be O 1s (531.6 eV), C 1s (286.6 eV), and N 1s (403.8 eV) owing to the polymer's main backbone and QAs, and Br (68.9 and 182.4 eV) appearance was caused by the bromide functionalization of the main backbone (Fig. 30a). The existence of such peaks confirmed the functionalization of the poly(fluorenyl ether) membrane by bromination and quaternization. The hydroxide ion can be increased by constructing composite inorganic–organic materials with graphitic-based nanomaterials and QAs-modified polymer backbones to boost the conductivity. As a result, quaternized polysulfone and graphitic carbon nitride (g-C3N4) were combined (g-C3N4-QPS).319 The presence of N, C, and O elements demonstrate the surface properties of g-C3N4. At the same time, the author only confirms the elemental composition of bulk and exfoliated g-C3N4 using XPS (Fig. 30b). Additionally, functionalized g-C3N4 was found to have stronger and higher peaks than bulk g-C3N4 as a result of oxidation.
![]() | ||
Fig. 30 XPS spectra (a) before and after the degradation of the QA-OMPEEK membrane in vanadium ion, reproduced with permission from ref. 318, Copyright 2022, Elsevier. (b) Bulk g-C3N4 and functionalized g-C3N4 (f-g-C3N4) membranes, reproduced from ref. 319, Copyright 2022, Royal Society of Chemistry. (c) Before and after treatment of PAP100-dimethyl and PAP100-POSSI in KOH, reproduced with permission from ref. 320, Copyright 2022, Elsevier. |
Porous silicon nanoparticles known as polyhedral oligomeric silsesquioxane (POSS) have a pore diameter of about 1.8 nm and many functional groups. POSS enhanced the electrolyte membranes' conductivity, mechanical, and dimensional stability. The poly(aryl piperidinum) (PAP100) membrane was therefore composited with imidazole-functionalized POSS (PAP100-POSSI), followed by quaternization using CH3I, and ion exchange was investigated to increase the conductivity and other features.320 In this experiment, greater peaks in the composition of the composite membrane included O (548 eV), N (402 eV), C (285.1 eV), and Si (100 eV) and the appearance of N and C was attributed to the PAP membrane while Si was attributed to the POSS membrane (Fig. 30c). PAP, however, could have lower conductivity and fewer other properties because of the appearance of their weaker peaks (O, N, and C). The POSSI-modified PAP (PAP100-POSSI) polymer composite AEM was confirmed by such stronger peaks, which might indicate significant physical, chemical, and conductivity properties. Compared to bare PAP-100-dimethyl membrane, the N 1s peak for PAP100-POSSI remains stronger after the alkaline treatment. After alkaline treatment, the dimethyl-piperdinium group has a ring opening for the PAP100-dimethyl membrane, and the group was exhibited at 400.0 eV. However, following the alkaline treatment, the imidazole group in POSSI at 400.0 eV was unstable.
Furthermore, XPS was used to study the structural analysis of block copolymer-based PES and quaternized with high-density QAs (MPy) and crosslinked with a lengthy rigid side chain.321 In this study, block copolymer PES that had been MPy treated showed that C–N and N 1s at 399.2 and 401.7 eV were obtained, respectively, by the polymer backbone and QAs (MPy). Only C–Br peaks at 70.9 eV were visible in the brominated PES polymer, while an additional peak at the Br ion at 67.1 eV was visible after functionalization with MPy. Specifically, Br ion and N 1s at certain binding energies on the block copolymer PES confirmed the bromination treatment and MPy quaternization. Similarly, the quaternized N-methylmorpholine (MM)-treated randomly organized copolymer polyether sulfone (rPES-MM) AEM showed nitrogen (401.7 eV) from QAs and C–N (399.2 eV) segment from the backbone.322 Before quaternization, brominated rPES (Br-rPES) also displayed ionic bromine and C–Br at 70.9 and 67.1 eV, respectively, confirming that the bromination and quaternization processes were carried out on rPES.
Recently, PVA, which had been crosslinked with glutaraldehyde and further composited with crown ether-functionalized amine-SBA and the membrane, was then activated by KOH.323 Si, C, N, and O XPS peaks were obtained for the composite membrane, and following activation, a K peak could be seen. However, no significant structural changes were seen in the composite membrane before and after KOH activation. Similarly, the N 1s peaks were detected in the binding energy ranges from 403 to 398 eV for the combination of (2-bromoethyl)trimethylammonium bromide (BTAB) and hexylbromide (HB)-functionalized poly(biphenyl pyridine)-based AEM.324 As a result, the N 1s peak at 399.8 eV was caused by pyridine, and the pyridinium and trimethylammoniums of N 1s peak at 402.1 and 401.5 eV was caused by the main backbone and BTAB, respectively. In the XPS spectra, the bromination and quaternization functionalization was formally supported by the formation of the C–Br, ionic bromine, and nitrogen peaks, as well as the stronger and higher N 1s peak, which suggested that the functionalized cation's high nitrogen content might be beneficial for high membrane conductivity.
![]() | ||
Fig. 31 FT-IR spectra of (a) azidated PPO (APPO) and di-quaternized PPO (DQ-PPO) membranes, reproduced with permission from ref. 231, Copyright 2021, Elsevier. (b) Isopropyl, n-propyl, and cyclohexyl-based QAs-functionalized PPO copolymers, reproduced with permission from ref. 232, Copyright 2021, Elsevier. (c) Poly(vinyl chloride) (PVC), crosslinked-QPVC-1N, crosslinked-QPVC-2N, and crosslinked-QPVC-3N membranes, reproduced with permission from ref. 129, Copyright 2021, American Chemical Society. (d) IR spectra of quaternized chloromethylated polysulfone (QCMPSF) and TMA-quaternized crosslinked chloromethylated polysulfone (c-QCMPSF-x) membranes, reproduced with permission from ref. 202, Copyright 2021, Elsevier. (e) –OH, –Cl, –Pi-functionalized PDMB copolymers, reproduced with permission from ref. 144, Copyright 2021, American Chemical Society. |
Compared to quaternary nitrogen-based cations, phosphonium cations are more stable in alkali conditions. As a result, the pTAP membrane, a phosphonium-based polymer, was developed and mixed with polysulfone for electrochemical use.237 The composite form of the pTAP/polysulfone membrane included the functional group of pTAP and polysulfone in the IR spectra. The phosphorus compound of phosphonium was connected to the phenyl groups of polysulfone (1000, 1147, and 1441 cm−1) present in the blended membrane. Due to membrane activation by KOH, the hydroxyl group (3421 cm−1) in the blended membrane was detected. The cations such as Mpy, MIm, MPip, and 2-methyl-1-pyrroline (Mpyl) functionalized in the chloromethylated SEBS were investigated, and their functionalization was confirmed using infrared (IR) spectroscopy.233 As a result, the aliphatic groups of SEBSs were located at 2918 cm−1 (–CH3), 2849 cm−1 (–CH2–), and 1453 cm−1 (–CH–), and the peak C–Cl (800 cm−1) occurred as a result of chloromethylation. Finally, it was determined that the CN and C–N vibrations of Im or MPyl and Pip or MPy were the sources of the cations tethering on chloromethylated SEBS at 1503 and 1200 cm−1.
Recently, TMA was used to quaternize polysulfone with crosslinked AEMs and their functionalization was examined using IR.202 The peaks for the benzene ring and the sulfoxide group of the polymer backbone emerged at 1485 cm−1 and 1073 cm−1, respectively. Later, the C–Cl group emerged at 559 cm−1 upon chloromethylation, confirming the chloromethylation of polysulfone. Following quaternization, the covalent interaction between the cations and the polymer backbone identified two additional peaks at 3363 (NR4+) and 1631 cm−1 (C–N). The presence of –OH groups from the crosslinker at 1731 cm−1 confirmed the structure of TMPTA-crosslinked Q-polysulfone (Fig. 31d). Similarly, in another study, the absence of peaks at 1722 cm−1 (–COOC) and 3400 cm−1 (–OH), respectively, indicating the hydroxylation and chloromethylation processes of PDMB.144 Later, a new peak caused by pyridinium appeared at 1627 cm−1 (–C–N), confirming that the cation (pyridinium) was functionalized on hydroxylated and chloromethylated PDMB (Fig. 31e). Zhang et al. investigated the chemical structural variations of the synthesized quaternized bis-piperidinium side chains-based AEM using IR.227 To this end, the N–C bond and CH3 functional groups were still present in the QAs group, as evidenced by the appearance of peaks at 1488 and 1465 cm−1, respectively, in the alkaline resistance test before and after. These appearance of these peaks indicates that AEM has alkaline stable qualities under difficult circumstances. Recently, the fading of the C–Br (618 cm−1) peak and the production of a new peak at 3335 cm−1, which was caused by the hydroxyl group (–OH), both indicated that the bromomethyl group in the star-shaped copolymer had been transformed into a quaternary ammonium group.127
Similarly, chloromethylated PAEK was aminated (CMPEAM) using TMA by a reductive mechanism.189 The IR spectra of the compound show peaks at 3470 to 3350 cm−1, indicating the presence of C–NH2 in the polymer, while the peak CO of polymer backbone was lost at 1660 cm−1 as a result of the conversion of the C
O group to C–NH2 in the polymer backbone. These results confirm that amination (quaternization) was done on chloromethylated PAEK (Fig. 32a). IR demonstrated absorption peaks at 2918 and 1103 cm−1, respectively, confirming dual side chains-functionalized PPO, such as alkyl and alkoxy chains.316 Similarly, the functional groups of brominated PPO (BPPO), which was then quaternized using DABDA, were examined.230 The peak C–Br was due to bromination and the bond C–C (1447 and 1596 cm−1), whereas the peak C–O–C (1184 cm−1) was due to the crosslinker DABDA, and the covalent link between DABDA and the polymer backbone appeared to be related to the highest C–N bond at 1017 cm−1 (Fig. 32b). Thus, these absorption peaks supported the functionalization of PPO with crosslinkers and bromination.
![]() | ||
Fig. 32 IR spectra of (a) QA-PE-CO and –NH2 polymers, reproduced with permission from ref. 189, Copyright 2021, American Chemical Society. (b) BPPO and PPO/DABDA-40 polymers, reproduced with permission from ref. 230, Copyright 2021, Elsevier. (c) BPPO and cross-linked PPO/AGO-8 membranes, reproduced with permission from ref. 242, Copyright 2021, Elsevier. (d) VImPPO and PVA (sIPN-91AEM) before and after crosslinking, reproduced with permission from ref. 128, Copyright 2021, Elsevier. (e) PVBMPy-CL-(10–30%)PSF membranes, reproduced with permission from ref. 263, Copyright 2021, Elsevier. (f) Before and after quaternized chitosan, reproduced with permission from ref. 250, Copyright 2021, Elsevier. |
Goel et al. studied aminopropylsilane-treated GO (AGO) that was functionalized with BPPO, and they used infrared spectroscopy to determine the chemical composition of the samples.242 As a result, this study demonstrates that a C–Br peak formed at 590 cm−1. When AGO functionalization was examined, the peak C–Br vanished, confirming the crosslinking of AGO with BPPO. Following quaternization, QAs were responsible for the peaks between 3200 and 3700 cm−1, and a CN bond at 1568 cm−1 may have resulted from a covalent interaction between QAs and AGO-PPO (Fig. 32c). The VImPPO and PVA polymers were made using a cross-linker and a light initiator.128 Thus, IR was utilized to compare the crosslinking of composite polymers before and after. Weaker CC and
C–H groups caused the peaks at 1640 and 940 cm−1 compared to the uncrosslinked composite polymer (Fig. 32d). The functional groups of the lengthy side chain BTMA-Br functionalized on PBI were examined, and they revealed a peak at 1630 cm−1 (C
N group), which was related to the PBI (benzimidazole ring).187 Possible causes of C–Br in BPTMA-Br include the typical peak at 740 cm−1, which later faded following functionalization with PBI. The absence of C–Br showed that BPTMA-Br functionalized the PBI.
A composite PVBC polymer treated with MPy and chloromethylated polysulfone was produced, and its structural aspects have been examined.263 IR data indicates that the peak at 675 cm−1 may represent a PVBC of C–Cl vibration. After the functionalization of MPy with PVBC, the peak C–Cl vanished (Fig. 32e). As a result of water molecules being absorbed on the membrane, hydroxyl groups are responsible for the hydrophilic MPy-treated PVBC, showing a peak at 3346 cm−1. The peak indicates the composite membrane on which the cation was functionalized. The weight ratio of polysulfone in the composite membrane made with PVBC-MPy may be adjusted, but the authors found no change.
Recently, Zeng et al. studied GA-crosslinked PVA combined with QPVBC.257 The peak at 1488 cm−1 in the IR spectrum was assigned to an aromatic ring (CC) by the polymer backbone, and the bond C–N at 890 cm−1 was detected by QAs. These C–N and C
C peaks in the composite membrane prove that QPVBC was constructed using cross-linked PVA. Furthermore, the composite PVA-QPVBC was formed, as evidenced by the peak at PVA's C–O–C bond in the composite membranes' IR spectra. Zhao et al. utilized quaternized chitosan/PVA composite AEM for the fuel cell application; however, they first confirmed the functional group of quaternized chitosan.250 As a result, it was revealed that hydroxyl and amine peaks were at 3432 cm−1, which is due to water absorption and amine in chitosan, respectively (Fig. 32f). After quaternization, a couple of significant peaks (3018 and 1484 cm−1) was because of the–CH bond of the QAs from the crosslinker (GTMAC). The epoxy group of GTMAC and the amine group of chitosan engaged in a ring-opening reaction, which resulted in the peak –NH at 1640 cm−1 being decreased as well. However, this study did not do an IR analysis of the Q-chitosan/PVA combination, and it was confirmed that chitosan was quaternized when the –CH group formed and the –NH group decreased.
In a recent study, IR verified the presence of PEEK functionalized with quaternized piperidinium.254 A C–O–C bond was discovered at 1226 and 1151 cm−1, which may be the polymer's backbone. The presence of a wide signal at 3400 cm−1 brought on by hydrophilic piperidinium's water absorption on the membrane was subsequently utilized to prove that PEEK had been functionalized with piperidinium. Glutaraldehyde-crosslinked PVA-branched polyethyleneimine composite membrane was developed by Xiao et al. and used for fuel cell applications.258 They examined the crosslinked composite membrane's structural composition before the application. The polyethyleneimine's amine groups were shown by the wide peaks seen at 3367 cm−1. The peaks at 1443 and 1605 cm−1 were attributed to the C–O groups in PVA, while the interaction between the amine in polyethyleneimine and the –CHO in glutaraldehyde was thought to be caused by CN, respectively. Similarly, a peak developing at C–O–C was used to establish the interaction between the OH groups in PVA and the –CHO groups in glutaraldehyde (1100 cm−1). Finally, the composite membrane's ion exchange (Cl− to OH−) was proven by red shifting a wide peak from 3367 to 3282 cm−1.
Kang et al. employed commercially available FAA-3 and Orion TM1™ AEM for water electrolysis application.325 Due to the hydrophilic nature of the membranes, both membranes exhibit a large peak at about 3600 cm−1 due to OH. Both membranes' diphenyl ether component and aromatic bond (C–C) were also responsible for the peaks at 1302 cm−1 and 1476 cm−1. AEM was created by functionalizing poly(arylpiperidinium) (PAP) with the hydrophobic guanidinium cation.293 The PAP-based backbone underwent IR both before and after the guanidinium cation. Guanidinium salt caused the CN bond to appear to this extent at 1538 cm−1. The appearance of such a C
N peak proved that the PAP polymer was functionalized with cations.
Quaternized perfluorinated polymers such as Nafion-SO2-F and Aquivion-SO2-F were explored by Lee et al.276 They examined the functional groups of the polymers and found that sulfonyl fluoride peaked at 1467 cm−1 and 796 cm−1 (SO2–F and S–F). Due to the presence of QAs, the sulfonyl fluoride peaks vanished, and a new peak (C–N) was found at 1020 cm−1. The OS
O peak was shifted following quaternization, and the sulfonamide (S–N) group was assigned at 1051 cm−1. The perfluorinated and quaternization of Nafion and Aquivion membranes were validated by the peaks that appeared, including sulfonyl fluoride, C–N, and S–N. Similarly, using the 1,4-dimethylpiperazine (DMP) cation, the Nafion with SO2-F form (Nafion-SO2-F) was quaternized. As a result of DMB functionalization, peaks like –CH3 and –CH2 were seen at 2783 and 3040 cm−1, respectively, whereas Nafion-SO2-F missed such peaks.277 The authors also noticed that the –CH3 and –CH2 peaks were stronger as the quaternization duration increased (0 to 12 h). Time is another crucial factor in producing a more robust quaternization on the membrane.
Similarly, N,N-dimethyl-butylamine (DMBA)-functionalized PIM-block copolymer PSF membrane underwent an alkaline test in 1 M NaOH before and after.307 The C–N bond of QAs at 1296 cm−1 was weaker after exposure to alkaline for 216 h; the nucleophilic substitution in the membrane may have caused this weakening peak. IR demonstrated that the functionalized ionic liquid membrane had greater alkaline stability than other cations (smaller side chains). Different cations, including DMIm, dimethylbutylamine, and dimethylhexylamine, were functionalized individually with perfluoroaklyl membrane.289 The spectra of the degraded qauternized perfluoro-based membrane were shown before and after the structural confirmations were checked using IR. Perfluoroaklyl membrane functionalized with quaternized dimethylhexylamine showed lower degradation than the membrane functionalized with other quaternized cations.
DMIm cation-functionalized PBI and PBI-Si hybrid confirmed their functional groups by IR.305 Accordingly, as a result of the cation (DIm), the imidazolium (NC–N) moiety peak appeared at 1536 cm−1, and the peak shift from 1630 to 1620 cm−1 was attributed to the benzimidazole (C
N), which may have been the cause of the functionalization of DIm and Si. Due to the dimethylimidazolium in the PBI–DIm and PBI–DIm–Si, the C–H stretching vibration was recorded at 2861 cm−1. When silica was discovered in a hybrid PBI–DIm–Si membrane, the Si–O–Si bond emerged at 1130 cm−1. The produced AEMs of structures were verified using IR absorption analysis based on the functional groups that appeared at certain wavenumbers (cm−1), such as –OH, –NH, –NH2, C–Br, C
O, C
N, benzene ring, C
C, –CF2, and C–Cl, as well as red shifting peaks and membrane structural alterations following an alkaline test.
![]() | ||
Fig. 33 TEM images of (a–f) di-quaternized PPO-based AEMs (DQ-PPO AEMs), reproduced with permission from ref. 231, Copyright 2021, Elsevier. (g) PISBr–NON-membrane, reproduced with permission from ref. 227, Copyright 2021, American Chemical Society. (h and i) QAFPK was made via a solution casting and hot-pressed electrospun modified QAFPK membrane, reproduced with permission from ref. 133, Copyright 2021, American Chemical Society. |
Interestingly, high ionic conductivity was achieved using the amphiphilic co-polymer to produce a hydrophilic and hydrophobic separation channel. Accordingly, different molar concentrations (72:
28, 62
:
38, 53
:
47) of star-shaped copolymers were synthesized, and quaternized membranes showed well-ordered microphase separation (hydrophilic and hydrophobic).127 The ionic domain sizes were increased (22–30 nm) with increasing copolymer molar concentrations, whereas linear-shaped copolymer-based AEM has a smaller ionic cluster size. Well-connected aggregated hydrophilic ion (darker region) and polymer backbone (hydrophobic) microphase separation were observed in the poly(bromoalkyl oxindolebiphenylylene)-based (PISBr) AEMs with cross-linked cation side-chain 1,2-bis(2-(2-methylpiperidine)ethoxy)-ethane (NON) (Fig. 33g).227 The side-chain (NON) may work well as a linker to improve microphase separation using a low-cost method. The hot pressing approach produced superior ionic channels for separating the hydrophilic and hydrophobic microphases during the manufacture of AEM (e.g., QAFPK) compared to the casting method.133Fig. 33h and i shows the TEM images of QAFPK membranes produced by solution casting and electrospun-modified QAFPK membranes made by hot pressing. Zhang et al. recently developed a C–NH2-linked QA-PAEK AEM.189 They found that the consistent hydrogen bond channels in the invented AEM may be supplying active sites for hydroxide transport and encouraging the formation of ion clusters favorable for increasing ionic channels and OH diffusion, thereby boosting long-range hydroxide conduction. Similarly, highly ordered microphase separation was obtained for 8% 1,6-dibromohexane crosslinked with poly(aryl piperidinium) (PAP-8%)-based AEM compared to 0% crosslinked poly(aryl piperidinium) (PAP-0%) AEM.121 Although, the ionic cluster size of PAP-8% was higher than PAP-0% due to the self-assembly of polymer backbones and accumulation of hydrophilic phases therefore, the creation of larger ionic clusters in the AEMs was promoted by the 1,6-dibromohexane-based crosslinking agent.
Crosslinked PPO with various concentrations (40, 50, and 60%) of DABDA shows that well-ordered hydrophilic (dark) and hydrophobic (bright) microphase regions were formed, and the aggregation of hydrophilic ions with increasing concentrations of DABDA.230 DABDA is one of the excellent cross-linking agents for promoting microphase separation to aid in membranes with ionic conductivity; thus, it makes sense that it is one of them (Fig. 34a–c). In another work, crosslinked naphthalene-based AEMs (e.g., NAPEK-PVP-(6, 8, and 10)-Q4) of TEM images demonstrate that increasing the QAs groups in the polymer backbone improved the size of the ionic domain, which permits better microphase separation.239 Accordingly, Fig. 34d–f depicts a TEM image illustrating different ratios (6, 8, and 10%) of QAs-functionalized NAPEK-PVP, with the darker region representing the AEM's hydrophilic area and the lighter region representing the hydrophobic polymer backbone. High microphase separation was achieved by synthesizing varied weight ratios of molecular weight 2000 of hydrophilic PEG combined with crosslinked PPO.255 PEG's increasing weight ratio in modified PPO demonstrates increased ion sizes and microphase solid separation. However, as shown by TEM pictures, only modified PPO (20PDM) exhibits no microphase separation due to no hydrophilic regions in the membrane.
![]() | ||
Fig. 34 TEM images of (a–c) PPO/DABDA-(40, 50, and 60%), reproduced with permission from ref. 230, Copyright 2021, Elsevier. (d–f) NAPEK-PVP-(6, 8, and 10)-Q4 membranes, reproduced with permission from ref. 239, Copyright 2021, Elsevier. (g–j) Zn2+ aggregation in PSF-ImCl/Zn-(0, 2.4, 2.9, and 3.8%) membranes, reproduced with permission from ref. 251, Copyright 2021, Elsevier. |
Intriguingly, imidazolium-functionalized PSF (Im-PSF) with the incorporation of different weight ratios (0 to 3.8%) of Zn2+ (Im-PSF/Zn2+) also outstandingly displays microphase separation and increasing ion cluster size with increasing the weight ratio of Zn2+.251 In contrast, no microphase separation was observed for only Im-PSF. Zn2+ ion in the Im-PSF, causing an expanding ionic cluster, which electrostatically combines with imidazolium (Fig. 34g–j). Recently, high levels of microphase separation were found in the hydrophilic N-spirocyclic (8-(4-(chloromethyl)benzyloxy)-5-azonia-spiro[4.5]decane) cation, and the hydrophobic n-octylamine functionalized with PSF AEM shows high levels of hydrophilic and hydrophobic microphase separation and the ionic cluster size was also increased with increasing hydrophobic contents.241 Similarly, a type of N-spirocyclic (6-azonia-spiro[5.5]undecane) cation was functionalized with GO (ASU-GO) in another study, and it was later further functionalized with QPPO. The functionalized PPO AEM produced ordered hydrophilic (cation-functionalized GO) and hydrophobic (PPO) microphase separation morphology.243N-Spirocyclic-based cations are key in generating microphase separation and can enhance the ionic conductivity in polymer membranes.
Well-ordered microphase separation was observed in quaternized poly(terphenyl) (m-TPN) functionalized with CH3I (Pi) compared to quaternized m-TPN-functionalized N,N-diisopropylethylamine (Py) and α,α′-dibromo-o-xylene (Be), respectively.240 The reason could be the higher degree of microphase (hydrophilic and hydrophobic) separation and increasing cations clouds in Q-m-TPN-Pi compared to others. Accordingly, the larger ionic cluster size was found to be Q-m-TPN-Pi. Interestingly, the hexyl QAs (HQAs)-functionalized SEBS AEM exhibit smaller and greater degrees of microphase separation, respectively. The pentafluorobenzoyl-functionalized HQAs-SEBS shows larger ionic clusters than monofluorobenzoyl-functionalized HQAs-SEBS.139 Without fluorobenzoyl-functionalized SEBS, however, no hydrophilic and hydrophobic separation morphology was observed. Lin et al. produced a series of poly(terphenylene) AEMs modified with piperidinium (cation).295 To this extent, multi-cation-functionalized poly(terphenylene) displays a higher microphase separated structure, and the ionic cluster size was increased compared to single cation functionalized poly(terphenylene) due to cation accumulation in multi-cation-functionalized polymer compared to single cation functionalization. As a result, cation side-chain (hydrophilic) functionalization is crucial for developing microphase separation structures during the fabrication of AEMs. According to the amount of functional or cations groups, the size of the ionic cluster increased.
![]() | ||
Fig. 35 SEM images of (a–c) various electrospun adapted polyketone-based AEMs (QAFPK-1-x-E) by solution casting and (d–f) optical photographs of the same AEMs by hot pressing, reproduced with permission from ref. 133, Copyright 2021, American Chemical Society. (g and h) Morphology and cross-sectional SEM image of cross-linked QSEBS-based AEM (C6-PQASEBS-62.3%), reproduced from ref. 203, Copyright 2020, Royal Society of Chemistry. (i and j) Morphology and cross-sectional image of block copolymer (rP(NB-MGE-b-[NB-O-BisIm+][OH−]2)-20), reproduced with permission from ref. 226, Copyright 2020, Elsevier. |
The membrane's smooth and dense structure aided in preventing fuel from leaking into energy-based devices, and this type of structure offers structural stability and mechanical qualities for energy-related applications. Accordingly, Zhang et al. observed a smooth, neat surface and dense (cross-sectional view) in bis-imidazolium-treated block-copolymer-based polynorbornene (e.g., rP(NB-MGE-b-[NB-O-BisIm+] [OH−]2)-20) AEM.226Fig. 35i and j shows morphology and cross-sectional focus of rP(NB-MGE-b-[NB-O-BisIm+][OH−]2)-20 membrane. The surface morphology of cross-linked PPO with aminopropylsilane-treated GO (PPO/AGO-8) and uncross-linked PPO AEMs (PPO/AGO-0) were similarly observed by Goel et al.,242 in contrast to cross-linked PPO (Fig. 36a and b), which produced a rough and non-uniform structured morphology. At the same time, uncross-linked PPO produced a clean surface (uniform), free of pores and densely organized (Fig. 36c and d). In contrast to non-cross-linked PPO, the authors asserted that great flexibility was discovered in cross-linked PPO AEM.
![]() | ||
Fig. 36 SEM images of (a and b) morphology and cross-section of crosslinked PPO with aminopropylsilane-treated graphene oxide (PPO/AGO-8) membrane and (c and d) uncrosslinked membrane (PPO/AGO-0), reproduced with permission from ref. 242, Copyright 2021, Elsevier. (e) PVAf, (f) crosslinked PVAf, (g) alkali![]() |
Recently, poly[2-20-(mphenylene)-5-50-bibenzimidazole] (ABPBI) was composited with PVA to fabricate electrospun nanofibers, considering that the incorporation of polymer nanofibers allows obtaining it due to ABPBI providing thermal resistance, mechanical strength, low permeability, and their durability toward the PVA membranes. Accordingly, its composition (ABPBI-PVA nanofiber) was prepared with a cross-linked (glutaraldehyde (GA)) structure, and it was analyzed through SEM.131 Composite fiber (PVA-ABPBI) morphology was unsmoothly compared to PVA fiber due to the incorporation of ABPBI, and the cross-linked composition (C-PVAf-ABPBI) morphology shows a lamellar structure (cross-sectional view) and homogeneous (Fig. 36e–h). In the case of VBC-functionalized LDPE-based AEM, a rough surface morphology and dense structure were obtained. The cross-sectional image shows the two-layer (VBC + LDPE) formation.132 Highly smooth, no separation phase, homogeneous, dense, and neat morphology was obtained for the three piperdinium cations-functionalized poly(terphenylene)-based AEM (3QPAP-x) compared to a single piperdinium-functionalized poly(terphenylene) membrane (1QPAP).295 Therefore, this study reveals that the smoothness of AEM depends on the amount of piperidinium functionalization (Fig. 36i).
It's interesting to note that surfaces made of (vinylbenzyl)trimethylammonium (VT) and N-vinylimidazole (NV)-functionalized poly(vinylidene fluoride-co-hexafluoropropylene)-based AEM were very clean, smooth, dense, uniform, and free of any other potentially dangerous substances.326 For energy-related applications, the functionalization of VT and NV are highly compatible with the copolymer to manufacture smooth AEM. Recently, dimethylimidazolium-functionalized PBI (PBI–DIm) and its hybrid with silica (Si) (PBI–DIm–Si) AEM obtained highly transparent, dense, smooth, and homogeneous surfaces.305 However, the Si on the PBI surfaces causes the hybrid structured membrane to lose its smoothness.
The organic–organic hybrid polymer composite membrane with a smooth, dense, homogeneous, and neat surface was made using PVBC crosslinked with poly(vinyl acetate).264 Compared to an organic–inorganic membrane, a composite organic–organic polymer could perform better because of its compatibility with polymer–polymer interactions. Additionally, the procedure is low-cost because poly(vinyl acetate) acts as a cross-linker agent, so no additional costly organic cross-linking agent is required. Wu et al. noticed that the polymer backbone and cross-linkers are critical in membrane morphology, which is crucial to highlight.271 As a result, when the polymer's 3 g BPPO backbone was functionalized with a medium concentration of DVB and VBC combination, it produced a tidy and uniform shape as opposed to a low (2%) concentration. However, using more than 3 g BBPO at the same medium concentration (6%) DVB shows that superior morphology was not achieved. Therefore, having the right amount of cross-linker and polymer backbone is crucial. As a result, the smooth and dense shape of the membrane relies on the weight proportion of the cross-linker and polymer backbone and the selection of composites of organic–organic and organic-inorganic matters.
![]() | ||
Fig. 37 TGA spectra of (a) PBP, PTP, and PQP-based membranes, reproduced with permission from ref. 235, Copyright 2022, Elsevier. (b) PVC and crosslinked QPVC-x membranes, reproduced with permission from ref. 129, Copyright 2021, American Chemical Society. (c) PTP-based membranes, reproduced with permission from ref. 234, Copyright 2021, Elsevier. (d) PPO-X-n membranes, reproduced with permission from ref. 232, Copyright 2021, Elsevier. |
Another research utilized the cross-linker triethylenetetramine (TETA) on PVC-based AEM. To determine the degree of quaternization, PVC was immersed in TETA solution for periods of 4, 8, and 14 h and the thermal stability of the membrane was accordingly studied.129 TGA curves indicate that weight loss occurred in stages, with the first occurring between 85 and 160 °C due to water evaporation and the second occurring at 220 °C due to the loss of TETA. The last stage occurred at 460 °C with the breakdown of the backbones (Fig. 37b). At a temperature of 430 °C, bare PVC began to degrade due to poor thermal performance. As a result, increasing the degree of quaternization via TETA over time improved the thermal stability of the PVC-based AEM. Only two stages of weight loss were discovered to be caused by aryl-ether-free PTP-based AEMs, showing that the first stage occurred between 250 and 350 °C, indicating that broken-down polymer backbones caused the breakdown of cations and weight loss at 400 °C.234 Thus, aryl-ether-free PTP-based AEM has decent thermal stability and may be employed for water electrolyzers and other applications (Fig. 37c).
Terminal alkyne (TA)-containing QA such as n-propyl (TA-nP), isopropyl (TA-iP), and cyclohexyl (TA-CH)-functionalized QPPO membrane in the iodide form demonstrate that weight loss temperature was initiated at 180 °C due to QAs degradation, and side chain decomposition was found to be in the range of 236 to 239 °C.232 The polymer backbone (PPO) was then decomposed at 400 °C. As a result, the thermal stability at high temperatures (400 °C) was improved by coupling massive QAs via an alkyl spacer (Fig. 37d). PPO was functionalized with alkyl-extender-containing dual-QAs devoted to the main chain through a spacer (PPO-SDQEC) showing low thermal stability, i.e., the side chains and cations were degraded at 150 °C, and the backbone (PPO) was decomposed in between 350 and 500 °C.113 The specific SDQEC-functionalized PPO did not exhibit significantly enhanced thermal stability.
The thermal stability of increasing amounts of cross-linking agent on quaternized polymer backbones was investigated, and it was found that when compared to quaternized polysulfone that had not been cross-linked, the rate of thermal degradation was slower for TMPTA cross-linked and TMA was used to quaternize (Q) polysulfone.202 All the prepared AEMs exhibit thermal stability up to a temperature of 370 °C; however, increasing the amount of TMPTA cross-linking on Q-polysulfone causes slower degradation than Q-polysulfone AEMs (Fig. 38a). Therefore, one of the factors affecting the membrane's thermal stability is the cross-linking agent with a higher degree that is chosen. Recently, quaternized PPO was cross-linked with aminopropylsilane-functionalized GO (AGO), demonstrating stability up to 450 °C.242 However, because AGO has a silica component, it was shown to have a slower degradation rate than Q-PPO when crosslinked with AGO. Therefore, adding silica components may improve the membrane's thermal stability.
![]() | ||
Fig. 38 TGA spectra of (a) crosslinked and uncrosslinked QCMPSF-x membranes, reproduced with permission from ref. 202, Copyright 2021, Elsevier. (b) Crosslinked VImPPO-PVA (sIPN-x) and uncrosslinked membrane (n-64), reproduced with permission from ref. 128, Copyright 2021, Elsevier. (c) Composite of MPy-functionalized crosslinked PVBC and PSF (PVBMPy-CL-x%PSF) and bare PVBC and MPy-functionalized PVBC (PVBMPy) membranes, reproduced with permission from ref. 263, Copyright 2021, Elsevier. (d) Crosslinked triblock (naphthalene) polymer (NAPEK-PVP-x-Q4), reproduced with permission from ref. 239, Copyright 2021, Elsevier. |
The weight ratio of the two polymer solutions was adjusted for membrane preparation and thermal stability. As a result, 1,8-octanedithiol was used to cross-link VImPPO, which was then mixed with a photoinitiator.128 Composite membranes were created by adjusting the cross-linked VImPPO and PVA in various ratios (8:
2, 7
:
3, and 6
:
4). Compared to other ratios, the 8
:
2 composite membrane achieved remarkable thermal stability (Fig. 38b). Also tested was a VImPPO and PVA composite without cross-linking, which exhibits poor thermal stability when compared to cross-linked VImPPO and PVA composite membranes. However, the membranes showed initial stability up to 180 °C and afterward weight loss, from 190 to 450 °C due to the decomposition of the cross-linker, polymer backbones including PPO and PVA. As a result, 1,8-octanedithiol utilized to cross-link membranes may improve their thermal stability.
Similarly, in a different study, TMHDA was used to cross-link poly(vinyl benzyl chloride) (PVBC), followed by its quaternization (Q) using MPy.263 Later, the cross-linked Q-PVBC was composited with polysulfone. The thermal stability of the composite membrane showed that the loss of water molecules caused weight loss at 130 °C; the degradation of MPy and TMHDA caused weight loss at 200 °C (Fig. 38c). Then, PVBC and polysulfone polymer backbones were finally broken down at 480 °C. However, compared to cross-linked Q-PVBC-(10 and 30%) polysulfone composites, only the PVBC membrane exhibits low thermal stability since it decomposed at 400 °C. MPy and TMHDA may thus improve the thermal stability of membranes. In another study, ionic liquid-crosslinked naphthalene and its composite with Q-PVP were found to be the weight loss temperatures at 210 and 370 °C, and this weight loss was attributed to the degradation of the crosslinkers and the polymer backbones, respectively.239 The type of composite membrane, however, did not significantly increase the thermal stability (Fig. 38d).
The cation's chemical structure greatly influences the thermal stability, which was observed by Wang et al.240 Accordingly, meta-poly(terphenylene) (m-TPN) AEMs were synthesized by quaternization with various cation structures such as CH3I, 1,4-dibromopentane, and α,α′-dibromo-o-xylene. To this end, m-TPN functionalized with CH3I shows low thermal stability than m-TPN functionalized with the spiro-cyclic structure cation of 1,4-dibromopentane and α,α′-dibromo-o-xylene membranes. Recently, the thermal stability of the chitosan membrane was investigated using TGA, showing stability up to 500 °C; however, chitosan following quaternization did not exhibit any significant stability (480 °C).250 Therefore, quaternization alone could not increase the membrane's thermal stability. Similarly, it was observed that the hexyl QAs-functionalized SEBS and the mono- and penta-fluoro benzoyl-modified hexyl QAs-SEBS had no appreciable impact on the thermal stability.139 Fluoro-functionalization does not influence hexyl QAS-SEBS membranes since all membranes have a comparable thermal effect (maximum stability 450 °C).
Another research compared the commercially available membrane FAA-3 with Orion TM1™, which has a backbone made of polyphenylene and functional groups QAs.325 The membranes exhibit a two-step weight loss, while the maximum stability was determined to be 450 °C for FAA-3 membranes and 500 °C for Orion TM1™ membranes. Jiang et al. investigated the thermal stability of bications cross-linked and uncross-linked poly(meta-terphenylene alkylene).279 However, no heat impact was seen for all of the membranes. The membranes' stable temperature was up to 360 °C, while QAs and cross-linker degradation caused weight loss at 210 °C. As a result, the thermal stability of membranes made of poly(meta-terphenylene alkylene) did not significantly change due to bication functionalization. Guanidinium salt is one component that substantially improves the membrane's thermal stability. Gao et al. developed a long multi-cation-based side chain (NQQN) cross-linked with fluorine-based PDBA AEMs with varying weight percentages (30 to 70%).283 All membranes exhibit three phases of weight loss at temperatures of 200 °C, 200 to 555 °C, and above 555 °C, and all membrane types have more robust thermal stability (555 °C) than the most current AEMs. To improve the thermal stability of membranes for applications, multi-cation (NQQN) functionalization is the optimum method. The functionalization of membranes, such as the addition of single or multiple cations, composites with other polymers, or inorganic materials, affects the thermal stability of such membranes. However, producing membranes with excellent thermal stability ought to be simple and practical for usage over an extended period.
FT-IR is one of the most effective methods for analyzing the functional groups of diverse organic and inorganic materials. The ability to recognize the functional groups would be beneficial to new researchers who are working with polymers. For example, it confirms that the hydroxyl group, the bending and deformation of ionic bonds, the aliphatic groups of polymers, the benzene ring and the sulfoxide group, bromomethyl group into QAs changes, alkyl methacrylate, CF2 group, CO group to C–NH2 due to QAs functionalization on chloromethylated polymer were found. Moreover, for the imidazolium (N
C–N) group, the changes in the structure before and after crosslinking, and the addition of inorganic particles such as silica and GO to the polymer were all confirmed. Improving the MSM by increasing the weight ratio of inorganic elements in the polymer backbone and the concentration of piperidinium in the polymer backbone is possible. These changes were attributed to larger ionic domain sizes, which was confirmed by TEM. The size of the ionic cluster may expand depending on the number of cation groups present in the membrane. Di-quaternized side chains have the potential to enhance microphase separation effectively.
SEM confirmed the smoothness, cross-sectional, and thick structure of the AEM, and it was found that the smoothness of the membrane can be lost when it is in composite form. Compared to a single polymer membrane and an inorganic–organic composite structure, a composite made from AEMs is very dense. Aromatic segment polymer structure, AEM based on piperidinium and p-tetraphenyl in the polymer is stable up to 350 °C. Though aliphatic side chains and thick QAs-linked AEMs showed that the maximum was stable up to 600 °C, this membrane type can be used for high-temperature applications. Cations such as MPi, MPy, and TMHDA made the membrane more stable at high temperatures. Unlike membranes made of a single material, composite membranes can make the membrane more stable in high temperatures. But, the inorganic–organic composite membrane is more stable at high temperatures than organic–organic polymer composites. The fluorination of the hydrocarbon polymer can also improve the temperature stability of the membrane very well. Up to 550 °C, fluorinated membranes with metal ions and metal-cation-based side chains were very stable.
![]() | (6) |
(Or)
Eqn (7) may also be used to compute the IEC of the AEMs. This equation titrates the membrane with a sodium hydroxide solution and uses phenolphthalein as an indicator after converting it to the OH− form using an HCl solution.234
![]() | (7) |
(Or)
![]() | (8) |
IEC calculation may also be assessed using eqn (8). The dry membrane (Wdry) was submerged in hydrochloric acid (HCl) at 60 °C for two days to neutralize the membrane's hydroxide ion. A KOH solution (CKOH) was used to titrate the remaining HCl molecules, and the consumed volume (Vx) was measured. V0 is the amount of KOH solution used to neutralize an HCl solution. To determine the amount of ion exchange in membranes, the reported papers provide a choice of IEC equations.128,134,187,239,240,257,287,301
Recently, PQP-100, PBP-67, and PTP-83 AEMs were produced by aromatic monomers p-quaterphenyl, p-biphenyl, and p-terphenyl and functionalizing with N-methyl-4-piperidone, respectively.235 There was no difference between these AEMs because they all had identical IEC values of 2.30 mmol g−1 at 20 °C. In another study, compared to the PEGDA (0.18 mmol g−1) membrane, the PS-DVB (1.5 mmol g−1) membrane had a higher IEC in the presence of Cl−, whereas PEGDA (0.190 mmol g−1) showed higher IEC in the HCO3− form compared to Cl−.260 An adjusted ratio of QAs (hydrophilic) and n-butyl groups (hydrophobic) was used to produce a naphthalene-based PBI membrane (NPBI-QA55-B45) to test its IEC.188 As a result, increased QA groups in the NPBI resulted in a higher IEC (0.92 to 1.80 mmol g−1) because more hydrophilic materials absorb more water (Fig. 40a). Additionally, IEC was decreased when n-butyl was functionalized with QAs-NPBI (0.83 to 1.56 mmol g−1). However, no IEC measurement was found after a certain quantity of QAs comprised NPBI-n-butyl since high WU causes dimensional change and membrane breakage. As a result, the optimization of increasing number of QAs should be considered on the AEM. Using Mohr titration, the IEC of dual QAs-functionalized PPO (DQ-PPOx-OH) could be determined with varying azidation degrees (12–26%).231 It was discovered that the IEC increased with increasing azidation degrees, from 1.35 to 2.03 mmol g−1. In order to improve the IEC of the membrane, the degree of azidation may therefore have an impact.
![]() | ||
Fig. 40 IEC profile as a function of water uptake (a) –NH-functionalized QPBI (NPBI-QAx) and n-butyl tethered NPBI-QAx (NPBI-QAx-By) membranes, reproduced with permission from ref. 188, Copyright 2021, Elsevier. (b) PPO-CH/nP/iP membranes, reproduced with permission from ref. 232, Copyright 2021, Elsevier. (c) SDQEO membranes, reproduced with permission from ref. 113, Copyright 2021, Elsevier. (d) IEC profile as a function of hydroxide conductivity for PTP-90 membrane, reproduced with permission from ref. 234, Copyright 2021, Elsevier. |
Recently, IEC testing was done on N-cyclohexyl-N-hexyl-N-methylcyclohexanaminium (CH)-functionalized PPO (PPO-CH), N-methyl-N,N-dipropylhexan-1-aminium (nP)-functionalized PPO (PPO-nP), and N,N-diisopropyl-N-methylhexan-1-aminium (iP)-functionalized PPO (PPO-iP) with varying degrees of azidation (between 31 and 44%).232 The results of this study indicated that higher azidation levels led to higher IEC (Fig. 40b). As a result, PPO-iP and PPO-nP with higher azidation (44%) groups both demonstrated higher IECs of 1.90 and 1.90 mmol g−1, respectively, when compared to PPO-iP (1.56 mmol g−1) and PPO-nP (1.56 mmol g−1) with 31% azidation. Besides, a lower IEC of 1.39 and 1.65 mmol g−1 was observed in PPO-CH functionalized with azidation levels of 31 and 44%, respectively. To this degree, the azidation degree of 44% of PPO-iP obtained higher IEC than cations-functionalized PPO with azidation degrees of 44 and 31%. Thus, the IEC depends on the chosen cation, and the degree of azidation is crucial, as revealed by this study's findings.
Intriguingly, PPO-based AEM with hydrophilic alkoxyl extender side chain and hydrophobic alkyl spacer (e.g., SDQEO) was received on the high WU, and minimal dimension change resulted in high IC.113 The dual-function PPO has the benefit of accumulating ions in the spacer between the primary and side chains to improve the conductivity. Even then, the IEC (1.13 mmol g−1) for this created AEM was not high (Fig. 40c). The N-cyclic piperidinium functionalized AEM with an alkyl ether-free bond (e.g., PTP-90) produced high IEC (2.52 mmol g−1), while the same membrane showed high WU.234 It is then proven that the membrane has a high IEC and can absorb more water when compared to PTP-75 and PTP-85-based membranes (Fig. 40d). An IEC of 1.5 mmol g−1 was received by poly(biphenyl piperidinium) AEMs based on ionic liquid and containing higher concentrations of piperidinium cation (e.g., PBPCL1.48).288 AEM's low cation concentration also exhibited a lower IEC (1.29 mmol g−1). As a result, the IECs of membranes rely on the concentration of cations within the membranes. In another study, the polymer composite polysulfone/tetraaryl-phosphonium (PS/pTAP) with the adjusted ratios 40:
60, 50
:
50, and 60
:
40 showed IEC of 1.2, 1.11, and 0.95 mmol g−1, respectively.237 The increasing ionomer pTAP amount in the PS polymer can increase the IEC due to its high hydrophilic nature to promote ion transport.
IEC was tested for the SEBS membrane grafted with shorter (CMME) and extended side chain (6-bromohexanoyl chloride) groups, after which they were quaternized with different nitrogen heterocyclic complexes.233 It was discovered that quaternized CMME-SEBS (e.g., SEBS-C1-MPy) displayed higher IEC (0.99–1.13 mmol g−1) than quaternized extended side chain-functionalized SEBS (0.44–0.48 mmol g−1). Therefore, the longer side chain could not benefit the membrane by enhancing the IEC. Goel et al. recently combined PPO with GO, which had been treated with aminopropylsilane (PPO/AGO).242 PPO was a composite with different weight percentages of AGO, which were changed by 0–8% for the efficiency tests. The IEC rose as the weight percentage of AGO in PPO increased. IEC often grows with WU. As a result, IEC (1.45–2.25 mmol g−1) and WU were raised in this study by increasing the weight percentage of AGO (2–8%) in PPO. Consequently, the hydrophilic qualities of AGO in PPO improved the IEC of the composite membrane.
In another study, when compared to bare (1.42 mmol g−1) and single antioxidant-doped (1.28 mmol g−1) membranes, the double antioxidant membrane (e.g., QP(VBC-St)-MB/EA0.3) doping (1.02 mmol g−1) did not aid in enhancing the IEC of the membrane.261 It could be because a single antioxidant-doped membrane produces better microseparation morphologies than a doped membrane with two antioxidants, which results in superior IEC. Therefore, adding two antioxidants to the membranes may not help get improved IEC. In the recent testing for IEC, several radical inhibitor structures doped with quaternized poly(4-vinylbenzyl chloride-styrene) (QP(VBC-St)) were evaluated.262 In comparison to bare QP(VBC-St) (1.72 mmol g−1), doped-QP(VBC-St) (1.44–1.38 mmol g−1) had lower IEC values. When compared to another radical inhibitor-doped AEMs (1.42–1.38 mmol g−1), the 4-tertbutylphenol-based radical inhibitor-doped QP(VBC-St) (e.g., QP(VBC-St)-4-TBP0.7) showed a higher IEC (1.44 mmol g−1). In contrast to radical inhibitor-doped AEMs, where the separation morphology influences the IEC of the membrane, QP(VBC-St) has a greater IEC because it has a well-ordered microseparation morphology.
The impact of varying the VImPPO-PVA (sIPN) semi-interpenetration polymer network ratio on IEC was recently investigated.128 The IEC values decreased from 1.46 to 0.78 mmol g−1 due to the lowering and increasing proportions of VImPPO and PVA, respectively. IEC was more significant in the higher and lower ratios (9:
1) of the VImPPO-PVA membrane (sIPN-91) than in other ratios (8
:
2, 7
:
3, and 6
:
4). As a result, VImPPO may be an excellent composite for PVA to improve the IEC. A higher PVA ratio in the composite membrane did not impact the IEC. The long-side chain structured QAs will improve the ion exchange of the AEMs. Tang et al. investigated several ratios (between 27 and 48%) of long-side chain-structured BPTMA that tethered PBI (BPTMA-PBI) evaluated for IEC.187 Finding out that the IEC (1.1–1.8 mmol g−1) rose as the proportion of BPTMA grew (27–48%) was quite interesting. Investigating the same thing using a different long-side chain structured QAs attached to AEM to evaluate the IEC would be preferable. Layered double hydroxides (LDHs), which are positively charged materials with high mechanical strength and a large concentration of hydroxyl groups, may aid in permitting better mechanical stability and hydroxide ion conduction when composited with AEMs. Due to the benefits of LDHs, Simari et al. recently explored several Mg–Al LDH compositions (1 and 5%) with quaternized polysulfone (QPSU) for IEC testing.134 Findings revealed that LDH-qPSU (e.g., qPSU-LDH5) displayed greater IEC than pristine QPSU. Interestingly, the QPSU's IEC (0.85–0.88 mmol g−1) values rise as the weight percentage of LDH (1–5%) increases, in contrast to the pristine QPSU's IEC of 0.81 mmol g−1. However, LDHs may not be the best choice to raise the IEC when combined with the membrane.
Poly(vinyl benzyl methyl pyrrolidinium) (PVBMPy), an ionomer, was recently crosslinked with varying amounts (10–30%) of TMHDA before being composited with polysulfone (e.g., PVBMPy-CL-PSF) to evaluate the IEC.263 With an increasing proportion of crosslinkers (0–30%), the IEC (3.6–1.6 mmol g−1) dropped in this study (Fig. 41a). The IEC values, however, were unaffected by the addition of polysulfone. As a result, adding polysulfone did not affect ion exchange. PVBMPy-PSF AEM demonstrated 2.1 mmol g−1 of the 15% optimized crosslinker (TMHDA). Increased IEC, on the other hand, resulted in increased water absorption and swelling ratio, which harmed the AEM. Therefore, the authors optimized a 15% crosslinked composite membrane for this research. In a different study, higher levels of crosslinking (5–15%) result in lower membrane IEC (2.99–2.74 mmol g−1).265 The ions are not exchanged due to the increasing crosslinking between membranes (e.g., 10%-PBP-c-SEBS). This investigation thus shows that improving the IEC of the membranes by increasing the crosslinking is not possible. However, since WR and IEC are strongly correlated, lowering the IEC would be beneficial for enhancing the mechanical stability of the membrane.
![]() | ||
Fig. 41 (a) IEC profile of PVBMPy-CL-x%PSF-based membranes, reproduced with permission from ref. 263, Copyright, 2021, Elsevier. (b) IEC study as a function of the conductivity/swelling ratio for N-cyclic cations-functionalized twisted poly(terphenylene) (m-TPN-x-QA)-based membranes, reproduced from ref. 240, Copyright 2021, Elsevier. (c) IEC study of uncrosslinked (AEM-0) and crosslinked quaternary poly(fluorene-piperidine)-based membranes (AEM-1 to 6), reproduced with permission from ref. 301, Copyright 2022, Elsevier. (d) P(4PA-co-2PA)-24 AEM, reproduced with permission from ref. 300, Copyright 2022, Elsevier. |
Recent attempts to enhance the IEC by introducing hydrophilic oligomers in the polymer backbones have been unsuccessful. Increasing percentages (0.5–5%) of PEGs were used to produce composites with functionalized PPO (e.g., 20PDM-(0.5–5%)PEG2000).255 Consequently, the increasing proportion of PEGs in the functionalized PPO lowered the IEC from 2.27 to 2.16 mmol g−1 compared to bare functionalized PPO. The PEG's lack of ions in the structure precluded ion exchanges; thus, the IEC was declining. Thus, the ions in the polymer serve as the foundation for the IEC in membrane improvement. The mass ratio of two polymers can be changed to increase the IEC; however, changing the mass ratio of the crosslinker and polymer results in a drop in the IEC for membranes. Accordingly, Zeng et al. studied IEC on composite membrane structures by modifying the mass ratio of two polymers (PVA:
QVBC) and adjusting the crosslinker and polymer (DVB
:
QVBC) to generate PVA0.6–0.9QVBC30–50%.257 IEC increases and reduces with the mass ratio of PVA with QPVBC (0.6–0.9%) and crosslinker (DVB) with QVBC (30–50%), respectively, rises. Therefore, ion exchange is feasible between the two polymer chains, not between the crosslinker and polymer. Thus, increasing the weight percentage of two polymers may be a better strategy to improve the IEC.
When a polymer is exposed to ionizing radiation, radiation-induced pathways cause significant changes in the polymer's intrinsic characteristics. As a result, several doses of electron beam irradiation (20, 40, and 100 kGy) on VBC-grafted ETFE-based AEM were investigated in an environment of air or N2 at either RT or below zero degrees.299 IEC (2.62 mmol g−1) was discovered to be higher for a high degree of grafting and 100 kGy radiation exposed to air atmosphere at low temperature (less than −10 °C) (e.g., 100-air-LT/RT) compared to other doses of exposure to air or N2 atmosphere at particular temperatures with a low degree of grafting. Compared to AEM made chemically, the IEC of the membrane may be improved by the low-temperature irradiation method. The method might minimize the need for hazardous chemicals and a labor-intensive manufacturing process for creating membranes. Researchers have been interested in N-cyclic-based cations because of their alkaline stability. Therefore, Wang et al. investigated various poly(terphenyl) (m-TPN)-tethered N-cyclic cations to produce membranes m-TPNPiQA, m-TPNPyQA, and m-TPNBeQA.240 It was found that m-TPNPiQA (2.54 mmol g−1) had a higher IEC than TPNPyQA (2.32 mmol g−1) and m-TPNBeQA (1.99 mmol g−1). Smaller cationic groups or microstructure configurations that result in significant ion exchanges might cause high IEC by m-TPNPiQA (Fig. 41b).
Zhou et al. investigated the impact of crosslinking on AEMs for IEC. The produced PBBP membrane was crosslinked with DBMHDA (AEM 1 to 3), BAPTES (AEM 4 to 6), and BPPO (AEM-7) to test for IEC.301 As a result, PBBP functionalized with DBMHDA (PBBP-DBMHDA) displayed a higher IEC (2.65 mmol g−1) than BAPTES (1.71 mmol g−1) and BPPO (1.55 mmol g−1). It could cause a much higher QAs occupancy rate in DBMHDA compared to BAPTES and BPPO (Fig. 41c). One of the causes may be that the silica groups in BAPTES block the hydroxyl ion exchange to the backbone. For BPPO, a smaller functional group mass per unit mass resulted in a lower IEC. Besides, the uncrosslinked PBBP AEM obtained higher IEC (1.9 mmol g−1) than BAPTES (1.71 mmol g−1) and BPPO (1.55 mmol g−1)-crosslinked PBBP. Therefore, DBMHDA could lead to better crosslinking to enhance the IEC of the PBBP compared to BAPTES and BPPO. The molecular weight ratio (0.24 to 0.63 mol%)-based copolymer (quaterphenylene alkylene and biphenylene alkylene) membranes (e.g., P(4PA-co-2PA)-24) has recently been shown to lower IEC (2.17–1.87 mmol g−1) by increasing the molecular weight ratio.300 The IEC was lower (1.85 mmol g−1) for the membrane-made biphenylene alkylene solely (Fig. 41d). Because of this, the membrane benefited from increased IEC in the copolymer form.
More cations in the AEMs mean more ion exchange between the cations and the backbone since IEC is directly proportional to the number of ion-conducting groups in the membrane. Thus, the impact of adding a fluorobenzoyl group on quaternized AEMs on their IEC was investigated. Accordingly, Munsur et al. investigated quaternized SEBS functionalized with monofluorobenzoyl (F1) and pentafluorobenzoyl (F5) to ascertain the IEC.139 Thus far, increasing fluorine substitution in the quaternized membrane did not improve the IEC. As an alternative, the IEC of quaternized SEBS (1.51 mmol g−1) was found to be greater than that of quaternized SEBS with fluorine (F1 (HQA-F1-SEBS) and F5 (HQA-F5-SEBS)) substitutes (1.49 and 1.45 mmol g−1), respectively. The restriction of ion exchange in QAs and SEBS by fluorine moieties is probably responsible for the decline in IEC. The mechanical stability caused by water uptake concerning IEC may be enhanced using a membrane of this form, reducing water absorption and swellings. Increases in IEC (1.62–2.26 mmol g−1) are shown when the ratio of poly(aryl piperidinium) with clustered cations-functionalized AEM (e.g., 3QPAP) is changed (0.3–0.5%) as compared to poly(aryl piperidinium) membrane with a single cation functionalized, e.g., 1QPAP (2.01 mmol g−1).295 Increased backbone ratio may result from improved ion exchange between clustered cations and the backbone, which raises the IEC.
Various monomers were employed in synthesizing different molarities (30–60%) of poly(ether sulfone)-based AEM, which was then tested for its impact on IEC.287 An increase in the molar ratio of monomer and the degree of bromination (e.g., 60MePh-2.07) caused the increase in IEC. The ion exchanges have undergone a comprehensive transformation because of the increased bromination in the membranes. Unlike conventional AEMs, block-copolymers AEMs do not perform well in the IEC. As a result, AEMs based on polystyrene and poly(phenylene) copolymers (e.g., HTMA-DAPP) had IECs of 1.7 and 1.5 mmol g−1, respectively.329 Incomplete or limited ion exchanges between the cation and backbones may be due to the block copolymer-based architectures of AEMs. This way, the IEC values of AEM structures based on block copolymers are lower than those found on conventional AEM structures. Recently, an AEM based on poly(arylene piperidinium) (PAPQ83) was synthesized and compared to commercially available AEM (Aemion™) in terms of IEC.294 PAPQ83's IEC value (2.29 mmol g−1) was similar to Aemion™ (2.1–2.5 mmol g−1). Thus, PAPQ83 may work on a commercial scale.
The IEC values of a single cation structured organic molecule-functionalized AEM with quaternization are lower. The functionalization of bication-structured organic compounds in AEMs has recently enhanced IEC. As a result, poly(meta-terphenylene alkylene) (m-PTPA) polymer was functionalized with bication structured compound (m-XPTPA-2N) and quaternized with TMHDA in different ratios (0–40%).279 The IEC of the bication functionalized m-PTPA was more significant (2.99 mmol g−1). However, following quaternization with TMHDA, its ratio (10–40%) is raised to reduce the IEC values (2.82–2.49 mmol g−1). As a result, the increased content of TMHDA could not exchange ions highly with the m-PTPA-2N membrane. Compared to a quaternized high number of QAs on a carbazole pendant cation-functionalized fluorinated membrane, the ion exchange was much higher in a fluorinated membrane that tethered a long alkyl chain-based carbazole pendant cation (e.g., PAEK-HQACz-0.7).110 The increasing ratio of alkyl chain-based carbazole cation increased the density of QAs in the hydrophobic fluorinated polymer to increase ion exchange and transport. In contrast, increasing the concentration without long alkyl chain-based carbazole cation with increased QAs in fluorinated polymer did not provide significantly better results in IEC. Recently, various ratios (4–10%) of (vinylbenzyl)trimethylammonium chloride were functionalized with the fluorinated polymer (PVIB), revealing the increase in the IEC (1.43–1.82 mmol g−1).326 Increasing the number of (vinylbenzyl)trimethylammonium chloride in the membrane could enhance the ion exchanges between the polymer backbone and QAs. Another research found that the IECs decreased from 2.84 to 1.55 mmol g−1 when the amount of extended alkyl chain-based bis-piperidinium cation groups in the quaternized membrane (e.g., QPBPip-Ac) was increased (from 33% to 67%).298 For the membrane to function correctly, it would need a different kind of bis-piperidinium cation group than one based on a lengthy alkyl chain.
IEC's influence on the fluorine-contained membrane with end-cross-linked structure (e.g., PHFB-VBC-DQ) was examined to determine the degree of multi-cation (30–80%).284 IEC was shown to rise (1.59–2.24 mmol g−1) when the percentage of cations (30–80%) in the crosslinked fluorinated membrane increased. The QAs tethered to the multi-cation structure may cause the fluorinated polymer backbone's enhanced ion exchange. Dimethylimidazolium-functionalized PBI composites with silicon (PBI–DIm–Si hybrid) exhibited an IEC of about 1.87 mmol g−1, as measured by Jheng et al.305 In contrast, those without an Si membrane showed 1.85 mmol g−1 hybrid membranes containing silicon increased the ion exchange. Quaternized poly(p-phenylene) and poly(arylene ether) copolymers were produced and evaluated for IEC by altering the molar ratio (2:
1 and 4
:
1) of monomer and trimer.306 It was shown that a more significant amount of monomer (e.g., QCPPAE-4/1) was used in the copolymer to get a greater IEC (1.75 mmol g−1) than that achieved with a smaller amount of monomer (e.g., QCPPAE-2/1) used in the membrane (1.08 mmol g−1). For this reason, increasing the monomer quantity is crucial for better IEC. In another study, different ratios (50 and 75%) of fluorinated polymer were recently created with cross-linked and uncross-linked THMDA for IEC.278 TMHDA cross-linked 75% fluorinated polymer (e.g., L-FPAEO-75-MIM) exhibited a greater IEC (1.8 mmol g−1) than cross-linked (1.36 mmol g−1) and uncross-linked 50% fluorinated polymer (1.32 mmol g−1). This work showed that IEC depends on cross-linking (quaternization) and polymer quantity.
Perfluorinated polymers have well-ordered structures and morphologies that assist water movement and minimal water absorption owing to their hydrophobic nature. Nafion and Aquivion, two perfluorinated polymers, were recently quaternized via TMA and evaluated for IEC.276 It was shown that the Aquivion-TMA (1.06 mmol g−1) had higher ionic conducting groups because it received a higher IEC than the TMA-treated Nafion (0.88 mmol g−1). Miyanishi et al. investigated AEM in the form of Cl− and OH− based on high molecular weight polyfluorene for IEC.100 The hydroxide form of polyfluorene-based AEM (e.g., PFOTFPh-C6-TMA (OH)) obtained high IEC, which led to the discovery that AEM-containing OH− had more ionic conducting groups than Cl−, a form of polyfluorene AEMs. Recently, it was shown that increasing the ratio (30–70%) of fluorene-based polymer precursors in multi-cation ionic liquid cross-linked fluorene-based AEMs (TQ-PDBA) boosted the IEC (1.09–2.16 mmol g−1).283 They increased the membrane's fluorine precursors to increase the number of ionic conducting groups per area of the polymer and increased ion exchanges. In their latest work, Li et al. found that the IEC (1.76 mmol g−1) of fluorine-tethered PPO followed by the tri-QAs membrane (e.g., PPO-22-3QA8F) was greater than that of tri-QAs-tethered PPO (1.71 mmol g−1) and single QAs-tethered PPO (1.72 mmol g−1).282 Higher hydrophobic and lesser hydrophilic properties of AEMs were found to have higher ion conducting groups than lesser hydrophobic and more inferior hydrophilic properties of AEMs because higher hydrophobic and lesser hydrophilic properties of fluorocarbon-tethered PPO with side chain tri-QAs received higher IEC than the lesser hydrophobic and lesser hydrophilic property one.
Recently, an attempt showed decreased IEC and hydroxide conductivity caused by the poor WU of hydrophobic crosslinked membranes.330 Higher WU is attributed to membrane swelling, mainly affecting the mechanical degradation and chemical properties. At first, the required size of the membrane was modified to the form of OH− ions according to the WU measurements. After that, the length was measured, dried, and weighed. The membrane was then sealed firmly and submerged in water for 24 h at the specified temperatures. Before degassing, WU was computed in water. After soaking, the membrane was immediately cleaned to eliminate any extra water accumulated on its surfaces, and it was then weighed and calculated. The collected measures were applied to measure the WU and SR. The membranes with OH− counter ions were soaked in water at a particular temperature, and the changes in mass and length of the membrane were observed. WU and SR were evaluated by following eqn (9) and (10), respectively.
![]() | (9) |
![]() | (10) |
Additionally, the amount of water particles absorbed per QAs group is known as the “hydration number (λ),” and it may be determined using the following eqn (11) or (12).234,239,295
![]() | (11) |
(or)
![]() | (12) |
The WU and SR were tested to be for the PQP-100, PBP-67, and PTP-83 AEMs, revealing that the p-biphenyl (PB) used to synthesize PBP-67 was claimed high WU (53.5%) and high SR (24.3%) compared to PQP-100 (20% and 20.6%) and PTP-83 (18.5% and 13.9%) at 20 °C was due to the less rigidity and high hydrophilicity of PB.235 The high rigidity and increased hydrophobicity of the p-quaterphenyl monomer in PQP-100 limit the high-water absorption and dimensional change and, thus, make it a viable AEM for water electrolyzer application (Fig. 42a). Moreover, the hydration number (λ) of PQP-100, PBP-67, and PTP-83 was 4.8, 12.9, and 4.5, respectively, and these numbers suggest that the PQP-100 received low hydration, which means less water absorption compared to others. Therefore, this AEM (PQP-100) results in reduced water absorption, and a less dimensional change will benefit the application. In another study, two different kinds of AEMs, based on poly(ethylene glycol) diacrylate (PEGDA) and polystyrene-divinylbenzene (PS-DVB), were developed by Kim et al.260 In the Cl− form, the PEGDA membrane (72%) displays higher WU than PS-DVB (18%). Later, it was found that the WU in the HCO3− form of both the membranes was greater than the Cl− form. The increased uptake of the PEGDA and PS-DVB membranes was caused by their fabrication using hydrophilic polymer (polyethylene glycol) and hydrophobic polymer (polystyrene)-based backbones, resulting in high and low WU. Furthermore, based on the findings of hydration number, it was confirmed that HCO3− membrane absorbed more water than the Cl− form of the membrane.
![]() | ||
Fig. 42 Water uptake and swelling ratio profile as a function of temperature (a) poly(aryl piperidinium)-based AEMs (PBP-67, PTP-83 and PQP-100), reproduced with permission from ref. 235, Copyright 2022, Elsevier. (b) NPBI-based AEMs, reproduced with permission from ref. 188, Copyright 2021, Elsevier. |
Increasing the QA amount in the primary NPBI very attentively revealed that the WU (7 to 35%) grew.188 However, after functionalization and adjusting the n-butyl (hydrophobic) group, the number of QAs-functionalized NPBI (e.g., NPBI-QA55-B45) showed an increment in WU (3.3 to 39.0%). As a result, this work demonstrates that WU only depends on the hydrophilic groups of QAs and is unaffected by the interference of other hydrophobic segments in the primary backbone (Fig. 42b). Additionally, the swelling ratio (SR) increased (5.4 to 11.8%) with the number of QAs in NPBI. Still, this impact later diminished (2.9 to 10.0%) after the QAs in NPBI were functionalized with hydrophobic n-butyl groups. As a result, adding hydrophobic material to QAs-functionalized membrane may lessen the swelling impact. Furthermore, the fact that the hydration number of n-butyl-functionalized QAs-NPBI (2 to 12.1 λ) was larger than that of QAs-NPBI (3.6 to 10.1λ) suggested that the functionalization of n-butyl increased the water absorption. The CuAAC click reaction results in triazole formation, which increases the number of active sites on the membrane and causes a rise in the WU and SR. Accordingly, dual QAs-tethered azidated PPO demonstrated increased WU (23.0 to 170.0%) and SR (4.6 to 70.4%) concerning the raised azidation (12–36%) of PPO and increasing ionic cluster size.231 The greater WU and high SR might not be helpful for the application. As a result, the membrane may be damaged under extremely high WU conditions that cause excessive membrane swelling. However, authors optimized the azidated manufactured membrane, i.e., DQ-PPO-17-OH, showing that the WU and SR were 60.0% and 12.7%, respectively. As a result, increasing the azidation of the polymer backbone may not be helpful due to the high WU and high SR damage to the membrane. Additionally, it was stated in this study that the WU and SR are influenced by the temperature, meaning that these values rise as the temperature increases.
At an azidation degree of 44%, the high hygroscopic nature of isopropyl-based cations-functionalized PPO (e.g., PPO-iP-44) exhibits higher WU (192.0%) than n-propyl (141.0%) and cyclohexyl (101.0%)-based cations-functionalized PPO.232 The PPO functionalized with identical cations and a 31% azidation level revealed lower WU values of 47.8, 38.8, and 22.6%. The hygroscopic isopropyl-based cation absorbs more water than other cations, which is what caused the higher WU to be measured. A higher azidation degree (44%) is another characteristic that helps the membrane absorb more water. Therefore, hygroscopic cations and azidation degrees are crucial in the WU. Moreover, these same investigations verified that higher WU and lower WU resulted in higher and lower SR, respectively. The hydration number of the PPO-iP membrane (56.1λ) was then found to be higher than that of the PPO-nP membrane (41.4λ) and the PPO-CH membrane (29.5λ) at a degree of azidation of 44%, and even PPO-iP (17.0λ) was found to have higher hydration than the other membranes (13.8 and 9.0λ) at a degree of azidation of 31%. The PPO-iP's higher hydration result demonstrated that it had a higher WU.
Due to the hydrogen bonding between the extender with the side chain and the polymer backbone when employing a hydrophilic-based alkoxyl extender, the WU and SR can be enhanced. However, very little WU and SR were seen when using a hydrophobic-based alkyl extender. Alkyl spacers are attached to the side chain and improve the WU and SR since consuming large amounts of water might encourage ion accumulation. Accordingly, Li et al. developed a hydrophilic alkoxyl extender into the side chain, and they linked it to the PPO main chain via a hydrophobic alkyl spacer (e.g., SDQEO) to the extent they did.113 At 25 °C, the alkoxy extender and alkyl spacer in the PPO were both high WU (120.1%) and showed a low SR (33.2%). Although the experiment (WU (130.0%) and SR (33.7%)) was carried out at 80 °C, there is still room for improvement. To assist the membrane and have high conductivity, the hydrophilic extender and hydrophobic alkoxyl might reduce the SR and increase the WU. N-Cyclic piperidiniums-functionalized PTP-based AEMs (e.g., PTP-90) strongly responded to an increase in WU (40–86.9%) and low SR (7.5–15.2%) with the increasing temperature (20–80 °C).234 It was proven that the low SR of PTP-based AEMs was acquired with an intense hydration number (8.7λ). For example, a low hydration number causes a low WU, and a low SR is related to high dimensional stability. Furthermore, the authors noted that the significant dimensional stability for PTP-based AEM was demonstrated at higher temperatures (80 °C).
A high SR and low OH− ion concentration in the membrane caused by an excessive WU will reduce the mechanical strength and IC of the AEMs. Therefore, a minimum quantity of water absorption would benefit the AEM's functioning. Designing an AEM with a hydrophobic moiety and long alkyl chains will reduce excessive water absorption. Accordingly, hydrophobic-based poly(biphenyl piperidinium) AEM was prepared with a higher concentration of cations β-cyclodextrin and piperidinium-based ionic liquid (PBPCL1.48), which was shown to have high WU (46.4%) and high SR (20.8%) in comparison to lower cation concentrations of similarly designed PBPCL1.29 (32.6% and 13.3%) and PBPCL0.79 (9.3% and 3.8%).288 Thus, one of the factors affecting the WU is the cation concentration in the primary backbone. Increased WU (5.4%, 5.6%, and 8.6%) was seen in the hydrophobic polysulfone with a rising hydrophilic ionomer (PS/pTAP) ratio (60:
40, 50
:
50, 40
:
60), and the external temperature (at 60 °C) also had an impact on the membranes' WU (7.8–10.3%).237 The increased WU was caused by the increase in the pTAP ionomer and higher temperatures (60 °C) because of the ionomer's hydrophilic nature and the enhanced mobility of polymer chain moieties and the production of free volume, respectively. However, this type of PS/pTAP AEM did not absorb much water, as evidenced by the finding that the WU at ambient temperature was only 8.6%.
Yu et al. investigated WU on several cations grafted onto shorter side chains (CMME) and longer side chains (6-bromohexanoyl chloride) SEBS membranes, respectively.233 At 30 °C, the WU (25–31%) of the longer side chain-grafted SEBS (e.g., SEBS-C6-Pip/MPy) was more significant than the shorter side chain-grafted SEBS (15–24%) such as SEBS-C1-Pip/MPy. Additionally, the hydration number at 80 °C showed that the longer side chain structured membrane (55–59λ) was superior to the shorter side chain-structured membrane (10–25λ). Longer side chain-tethered SEBS and shorter side chain-tethered SEBS exhibit high and low hydration numbers, 55 and 25, respectively, in the quaternization. WU was also increased as the temperature rose (30–80 °C). Therefore, the factors contributing to the membrane absorbing more water are the cation and the length of the side chains. Also, the SR of the piperidinium-functionalized shorter and longer side chain-tethered SEBS membranes was greater than that of the other cation-functionalized membranes due to the higher WU. As the temperature rises, the membranes' area and volume expand, i.e., the SR of the membrane is influenced by temperature.
As the wt% of 3-aminopropyl trimethoxysilane-functionalized GO (AGO) increased from 2–8%, the WU (24–37%) as well as SR (6.3–9.8%) were enhanced after the composite with PPO (PPO/AGO).242 But according to PPO, WU and SR were just 20.3% and 4.7%, respectively (Fig. 43a). Thus, the AGO's hydrophilic capabilities might improve the membrane's intrinsic features. In a recent study, Peng et al. found that double antioxidant doping of quaternized poly(4-vinylbenzyl chloride-styrene) (e.g., QP(VBC-St)-MB/EA0.3) significantly reduced the amount of WU (33.8%) compared to undoped (44.1%) and single antioxidant doped (59.7%) versions.261 It could cause two antioxidants to be added to the membrane to increase its hydrophobic property. Similar results were found in the SR of quaternized poly(4-vinylbenzyl chloride-styrene) doped with two antioxidants (39.6%) compared to membranes that were left bare (60.7%) and membranes that were doped with a single antioxidant (45.5%). Antioxidant doping into membranes can also be used to regulate WU.
![]() | ||
Fig. 43 Water uptake and swelling ratio study of (a) crosslinked and uncrosslinked PPO-AGO membranes, reproduced with permission from ref. 242, Copyright 2021, Elsevier. (b) VImPPO-PVA (sIPN)-based AEMs, reproduced with permission from ref. 128, Copyright 2021, Elsevier. (c) Crosslinked PVBMPy-PSF-based AEMs, reproduced with permission from ref. 263, Copyright 2021, Elsevier. |
Modifying membranes using radical inhibitors can increase the WU. As a result, the changed QP(VBC-St) AEM with radical inhibitors (e.g., QP(VBC-St)-4-TBP0.7) showed greater WU (39–47%) than unmodified QP(VBC-St) AEM (37%).262 However, compared to other radical inhibitor-modified AEMs (39–45%), the benzophenone-based modified membrane (47%) had a higher WU. It could cause benzophenone's more hydrophilic structure compared to other radical inhibitors. The SR of the radical inhibitors-doped membranes (26–39%) showed a similar pattern, which was greater than that of bare QP(VBC-St) (27%). Coppola et al. improved the WU of an ABPBI with a PVA.131 As a result, the membrane ABPBI only exhibited 48.0% and was increased by 67.0% after compositing with PVA. Due to the PVA's hydrophilic properties, which allow it to absorb more water into the membrane, its WU was higher. The structure of the membranes is crucial for controlling the SR; for example, a membrane in the shape of a fiber (e.g., C-PVAf-ABPBI) had a lower SR (27%) than a membrane in the form of a sheet (84%) analyzed under 15% KOH solution up to 7 days. Compared to sheet-shaped membranes made by casting, the fiber (by electrospinning method) form's lower SR can result from less free space or volume to absorb less water. In a different investigation, higher WU, higher SR, and higher hydration numbers were shown to correlate with higher IEC values achieved by VImPPO-PVA (sIPN) composites.128 However, when the IEC levels fall, so do the WU, SR, and hydration values (Fig. 43b). The declining WU, SR, and hydration numbers of composite membranes may result from functional group aggregation at higher concentrations, resulting in worse performances. In comparison to other ratios of VImPPO–PVA (8:
2, 7
:
3, and 6
:
4), the composite membrane ratio 9
:
1's (sIPN-91) higher IEC values revealed greater WU (48.68%), SR (20.61%), and hydration numbers (16.41λ). While composite membranes with higher hydration numbers had higher WU, SR, and IEC, those with lower hydration numbers had lower WU, SR, and IEC. As a result, the hydration levels also affect the membrane's characteristics.
Improving the long-side chain-structured QAs-tethered membranes are expected to enhance the WU and SR. Accordingly, Tang et al. explored long-side chain structured BPTMA (27 and 48%)-tethered PBI (48%BPTMA-PBI) AEMs for testing the WU and SR.187 To this extent, the WU (37–57%) and SR (25–30%) rose along with the percentage of BPTMA (27–48%) in the PBI. Due to its hydrophobicity, however, bare PBI had lower WU (16%) and SR (7%). BPTMA may be the superior QAs to increase the membrane's hydrophilicity and improve the WU and SR. According to Simari et al., Mg–Al LDHs-modified QPSU membrane (e.g., qPSU-LDH5) demonstrated more water absorption (60%) than pristine QPSU (58%).134 Additionally, increasing the LDH wt% from 1 to 5 results in a decrease (from 60% to 58%) rather than an increase in water absorption. It could cause the blocking of the composite membrane pore structures at higher concentrations. The LDHs-modified QPSU WU findings, however, showed that there is not much water absorption by the LDHs.
According to a recent study, PVBMPy AEM's high hydrophilic content with an IEC of 3.6 mmol g−1 can lead to greater WU and SR.263 However, these values are useless for application usage because of the membrane's decreased mechanical stability brought on by greater WU and SR. As a result, PVBMPy was crosslinked with TMHDA (10–30%) before being combined with PSF and having verified WU and SR (Fig. 43c). The WU (49.4–16.3%) and SR (34.6–17.6%) decreased due to the rising crosslinker concentration (10–30%) in PVBMPy-PSF. PSF's hydrophobic nature is also one reason to lower the WU and SR of the PVBMPy-THMDA% at this stage. Thus, TMHDA and PSF were added to improve the PVBMPy performance by reducing WU, which in turn reduced the SR and can enhance the membrane health.
Similarly, increasing the crosslinking percentages (5–15%) of the PBP membrane into (SEBS-PBP-c-SEBS) revealed that decreasing the IEC resulted in lower WU (131–34.3%) and SR (13.2–91%) values.265 However, the WU and SR of the crosslinked PBP-SEBS membranes increased with rising temperature, indicating significant responsiveness to temperatures. To sum up, increased PBP crosslinking might lessen the WU, which is what causes lower SR to be attained. It could be the hydrophobic properties of PBP that are impeding the WU. Water molecules stimulate hydroxide ion conduction by the Grotthuss mechanism, which is why water absorption is directly proportional to the IC of the membranes. WU is further dependent on the membrane's hydrophilicity. Therefore, PEG, a hydrophilic oligomer, was investigated for WU and SR in the composite form with hydrophobic PPO (20PDM-PEG2000) at varying amounts (0.5–5%) of PEG.255 As the PEG percentages grew up to this point, the WU (73.4–89.7%) and SR (69.3–77.5%) both increased compared to that of pristine functionalized PPO (WU (68.2%) and SR (66.3%)). PEG can, therefore, improve the WU and SR properties of the membrane. Increased PEG, however, increased the likelihood that more water would be consumed, and swelling harmed the membrane's structural stability.
On the other hand, it was shown that the WU and SR decreased when the IEC values rose. As a result, the WU and SR were controlled by increasing the mass ratios (0.6–0.9%) of QVBC with PVA (PVA-PQVBC) and increasing the mass ratio (30–50%) of divinylbenzene (DVB) with QVBC in the composite membrane of PVA0.6–0.9PQVBC30–50% (cross-linked Q-PVBC/cross-linked PVA).257 Since the WU and SR are controlled when both QVBC with PVA and DVB with QVBC increase, the resulting crosslinked composite membrane may help improve the membrane's mechanical stability. Recently, different doses of electron beam irradiation at various atmospheres with varying temperatures have been investigated for WU and hydration numbers on VBC-grafted aminated ETFE AEMs.299 In comparison to a low degree, the smaller dosage of the beam, air, and RT, it was found that higher radiation degree-grafted ETFE had higher WU (256%) and hydration number (54λ) when exposed to a high dose (100 kGy) electron beam radiation under air environment and low temperature (e.g., 100-air-LT/RT). Possible causes include high-degree VBC grafting, high-dose electron beam treatment, and low-temperature enhancement of the high amination, which increased the WU and hydration number.
AEMs (m-TPN) were functionalized with N-cyclic-based cations to synthesize m-TPNPiQA, m-TPNPyQA, and m-TPNBeQA for evaluating the WU and SR.240 The WU and SR were found to be higher for m-TPNPiQA (52.28 and 21.14%) than m-TPNPyQA (45.12 and 16.15%) and m-TPNBeQA (30.5 and 12.5%). Due to the stiff structure, the AEMs had lower WU and SR than m-TPNPiQA and had lower IEC. Due to its high hydrophilicity, m-TPNPiQA has a greater WU, resulting in a higher SR and IEC. Additionally, it was verified that m-TPNPiQA (10.92λ) had a higher amount of hydration than m-TPNPyQA (10.36λ) and m-TPNBeQA (8.19λ) AEMs. As a result, high and low WU and SR are produced by hydrophilic and hydrophobic-based cations, respectively. Crosslinking's impact on fluorine-based PBBP membranes was observed on WU and SR by Zhou et al.301 The BPPO-crosslinked PBBP had lower WU (28.1%) and SR (5.9%) at this point than DBMHDA-crosslinked PBBP (PBBP-DBMHDA) WU (53.5%) and SR (14.3%), BAPTES-crosslinked PBBP (BAPTESDBMHDA (WU (40.3%) and SR (11.2%)), and bare PBBP (WU (90.8%) and SR (28.2%)) at 80 °C. The shorter side chains between them, which result in less volume to take in water and fewer swellings and this crosslinked membrane may have high hydrophobicity, may cause a lower value of BPPO-PBBP. In this instance, more hydrophilic cations in the DBMHDA caused a significant increase in water absorption and swellings. The OH− groups produced by the silica structure in BAPTES inhibit water absorption and result in low swelling values.
Intriguingly, WU and SR can be controlled by preparing copolymer-based AEMs. Thus, by raising the molar ratio (0.24–0.63 mol%) of their total units, the WU (18.8–9.0%) and SR (3.6–0.6%) were kept under control by the synthesized alkylene-based copolymer (quaterphenylene and biphenylene) (e.g., P(4PA-co-2PA)-24).300 In contrast, the WU (28.7%) and SR (6.1%) were enhanced in the quaterphenylene alkylene-free copolymer. Because of its hydrophobic nature, water absorption and swelling may decrease by increasing the copolymer unit's quaterphenylene alkylene molar ratio. It was also shown that the rigid/hydrophobic structure of quaterphenylene alkylene causes the hydration number of the copolymer (4.5–2.5λ) to decrease with an increase in the total molar ratio (0.24–0.63 mol%) of the copolymer unit, i.e., the water absorption of the copolymer unit was reduced with increase in the molecular weight of the total unit. On the other hand, when comparing the copolymer structure forms, the quaterphenylene alkylene-free membrane revealed a higher hydration number (7.8λ), indicating a greater capacity for water absorption.
The quantity of hydrophilic groups in the membrane determines how much water may be absorbed. Low water uptake indicates that hydrophobic membrane groups are much more abundant than hydrophilic ones. Because of this, when the number of hydrophobic fluorobenzoyl groups (from mono to penta) in the quaternized SEBS (HQA-F1-SEBS to HQA-F5-SEBS) membranes increased, the WU (150.9–136.0%) and SR (39.0–36.8%) declined at 80 °C.139 More water WU (165.9%) and SR (47.9%) were seen with the fluorobenzoyl-free hydrophilic quaternized SEBS. Therefore, the increased water and swelling in the membrane may be reduced by the hydrophobic groups of fluorobenzoyl. When compared to single cations (less concentrated cations)-functionalized backbone (21.8% and 10.9%), clustered cations-functionalized backbone (e.g., 3QPAP-0.5) exhibits an increase in WU (8.3–26.0%) and SR (2.5–9.2%).295 In contrast to cluster cation-functionalized backbone, it is uncommon to observe that the SR was better for single cation-functionalized backbone. As a result, compared to a single cation-functionalized membrane, cations in more excellent concentration cluster forms stimulate hydrophilicity to absorb more water. It was also demonstrated that clustered cation-functionalized membranes (6.4λ) had more significant hydration numbers than single cation-functionalized membranes (6.0λ), which produced higher and lowered WU, respectively. In another investigation, raising the ratio of monomers and degree of brominations (e.g., 60MePh-2.07) led to improved cationic groups, attributed to increasing the WU and SR.287 The membrane's conductivities might benefit from increased cationic group sites, thus boosting water uptake.
It is well known that a higher membrane IEC results in a higher WU. It was determined that the IEC (1.7 mmol g−1) and WU (144%) of polystyrene copolymer-based AEM (e.g., HTMA-DAPP) were superior to those of poly(phenylene) copolymer-based AEM (1.5 mmol g−1 and 98%).329 As a result, the hydrophilic nature of the polystyrene copolymer-based AEM structure may cause it to absorb a more significant proportion of water than the poly(phenylene)-based copolymer. Another investigation found that because of the PAPQ83 AEM's highly hydrophilic structure, its effective water absorption resulted in a 100% WU.294 The membrane may include hydrophobic structures, leading to a 37% decrease in WU compared to commercially available Aemion™. As a result, it is possible that WU is dependent on the hydrophilic structure of the AEM and that WU plays a significant role in ion conduction across the membrane. Water molecules move more freely at higher temperatures, allowing membranes to absorb more water. TMHDA and its ratio increase quaternization to reduce the WU and SR. Therefore, bication-functionalized AEM (e.g., m-XPTPA20-2N) without quaternization displays greater WU and SR at higher temperatures.279 Increased hydrophobic characteristics in the membrane may result from decreased WU and SR caused by quaternization using THMDA on bication-tethered AEM. It was also shown that the hydration number (15.6–10.0λ) of the bication-functionalized AEMs decreases with an increase in the cross-linking ratio (TMHDA (10–40%)). This work indicates that the water inside the membrane decreases proportionately to the cross-linking present.
Quaternized PAEK with a long alkyl chain-based carbazole pendant cation (e.g., PAEK-HQACz-0.7) demonstrated higher WU and better SW control than quaternized PAEK with a carbazole-based cation functional group but no long alkyl chain.110 When the concentration of cations increased with the increase in the number of QAs, the WU and SR rose accordingly (Fig. 44a). However, the entanglement effect was thought to be responsible for limiting the size alterations of the fluorinated membranes when a long alkyl chain was added to the fluorinated membrane. IEC results are linked to membrane SR. To that end, the SR of a (vinylbenzyl)trimethylammonium chloride-tethered fluorinate membrane (e.g., PVIB-10) rises from 13.2 to 20.5% as the IECs (1.43 to 1.82 mmol g−1) are raised.326 It may be because (vinylbenzyl)trimethylammonium chloride raises the membrane's hydrophilicity, thus increasing its SR. According to another study, when the percentage of long alkyl chain-based bis-piperidinium cation groups in the quaternized membrane (e.g., QPBPip33Ac) was raised from 33 to 67%, the WU was reduced by 53.3% to 8.7%, and the SR was decreased from 23.50% to 23.75%.298 The potential causes of this rise in hydrophobicity include a decrease in the membrane's water utilization efficiency and swelling. With increasing levels of bis-piperidinium cation functionalization, it was shown that the hydration number of the membrane dropped from 10.4 to 3.1λ, indicating that less water was absorbed.
![]() | ||
Fig. 44 Water uptake and swelling ratio profile of (a) PAEK-HQACz-x-based membranes, reproduced with permission from ref. 110, Copyright 2021, Elsevier. (b) DMBA-PIM-b-PSF-x-based membranes, reproduced with permission from ref. 307, Copyright 2021, Elsevier. |
The WU (19.6–50.7%) and SR (16.7–26.1%) at 80 °C have recently been shown to increase with increasing ratios of multi-cation (30–80%) structured compound in the quaternized fluorinated cross-linked membrane (PHFB-VBC-DQ).284 Increased water absorption and some swelling may result from the hydrophilic nature of multi-cation. Higher cation ratios resulted in more water absorption by the AEM, as seen by a rise in hydration (6.8–12.5λ). It is well known that WU is proportional to IEC. However, Jheng et al. found that WU drops as the IEC values climb.305 Dimethylimidazolium-tethered PBI AEMs demonstrated 62.7% WU and 26.7% SR. After being hybridized with silicon (Si) (PBI–DIm–Si hybrid), the WU (30.9%) and SR (10.0%) were both reduced, and the hydration number revealed that the bare membrane had a more effective water absorption rate (18.8λ) than the hybrid form (9.2λ). Accordingly, the WU and SR could be controlled if the membrane was composited with Si. Similarly, membranes composed of block co-polymer structures comprising quaternized hydrophobic fluorinated polymer and hydrophilic polymer (e.g., DMBA-PIM-b-PSF-0.91) may exert control on the WU and SR.307 The increases in WU and SR were also seen in quaternized hydrophilic polymer membranes that lacked a block-copolymer structure-based membrane (Fig. 44b). The copolymer structure that contains fluorinated hydrophobic polymer restricts the WU.
Intriguingly, the copolymers (poly(p-phenylene) and poly(arylene ether))-based AEMs (QCPPAE) were studied for their effects on WU and SR when the monomer-to-trimer ratio was altered from 2:
1 ((QCPPAE)-2/1) to 4
:
1 ((QCPPAE)-4/1).306 When the copolymers' monomer content is raised, the WU (24–40%) and SR (11–23%) are raised. Since a more significant monomer concentration results in more water absorption through the membrane, utilizing a minimum monomer concentration to control the membrane's dimensions is preferable. By constructing the membrane with a TMHDA cross-linking structure (e.g., L-FPAEO-75-MIM), it can exert control over both the WU and the SR.278 WU has recently been evaluated for quaternized perfluorinated polymers such as TMA-Nafion and Aquivion-TMA.276 The WU of the two membranes was 21% and 23%, respectively. As a result, the SR of the two membranes was only 5% comparable. It is possible that the hydrophobic character of these membranes limits water absorption, resulting in reduced swelling, and that they may be utilized for long-term applications in fuel cells, water electrolysis, and so on. Polyfluorene membranes in the OH− form (e.g., PFOTFPh-C6-TMA(OH)) were shown to have improved WU and SR over those in the Cl− form, as corroborated by Miyanishi et al.100 The results of this research make it clear that the OH− form AEMs are more hydrophilic than the Cl− form AEMs, which suggests that OH− form polyfluorene AEMs may have higher WU and SR.
Very little water WU (4.1–10.8%) and SR (1.7–3.23%) were discovered in the multi-cation ionic liquid-functionalized fluorene-based (e.g., TQ-PDBA-70%) AEMs.283 It could be because the long and flexible multi-cation ionic liquid side-chain acts as a crosslinker between the leading backbone chains, providing adequate space to store water. However, the crosslinked cations structure somewhat restricts the top chains' transport and the membranes' bulging. As a result, the multi-cation functionalized fluorene-based AEM reduced WU and swelling, which might be helpful in long-term applications. Recently, WU and SR were examined using single and tri-QAs tethered PPO and fluorine-tethered PPO with tri-QAs.282 Tri-QAs-tethered PPO (e.g., PPO-22-3QA8F) was found to have greater WU and SR than single QAs-tethered PPO and fluorine-tethered PPO-tri-QAs. Because tri-QAs-tethered PPO is more hydrophilic than other AEM, it has a greater WU and SR. Other AEMs may have weaker hydrophilic properties, reducing the WU and swellings.
![]() | (13) |
Eqn (13) shows the membrane's cross-sectional area and the distance between reference electrodes A and L.
According to research by Liu et al., p-quaterphenyl, an aromatic monomer utilized to create the AEM PQP-100 (56.7 mS cm−1) has a higher IC value than p-biphenyl (PBP) and p-terphenyl (PTP), which were used to develop the AEM PBP-67 (39.2 mS cm−1) and PTP-83 (47.2 mS cm−1), respectively.235 Compared to PBP-67 and PTP-83, whose conductivities were low, PQP-100's enhanced conductivity was caused by the well-ordered microphase hydrophilic (p-quaterphenyl)-hydrophobic (polymer backbone) separation effect. Additionally, the ionic conductivity of AEMs increased with respect to the temperature. AEMs such as the PEGDA and PS-DVB of conductivities in the form of Cl− (1.2 and 7.0 mS cm−1) and HCO3− (0.8 and 4.1 mS cm−1) were found to be the membrane in the form of HCO3− and had lower conductivity than the conductivities in the form of Cl−.260 Low water absorption-based membranes may achieve high IC in the Cl− form because the state of AEM hydration in Cl− was lower than in the HCO3− form. In another investigation, the IC value was measured at 25 and 80 °C with and without an n-butyl-functionalized QAs-NPBI (NPBI-QA55-B45) membrane.188 After n-butyl was functionalized, IC increased along with the number of QAs in NPBI (8.9 to 31.7 mS cm−1 at 25 °C). But adding more n-butyl did not significantly improve the IC. Therefore, more hydrophobic n-butyl content might lower the IC of QAs-NPBI, and the membrane at 80 °C had a better IC than at low temperature (25 °C). Increasing the number of QAs into NPBI without n-butyl showed significantly lesser IC (2.8 to 2.96 mS cm−1 at 25 °C). Therefore, to improve conductivity in the membrane, n-butyl-based groups may be helpful.
Dual QAs side chains are promising for membrane ionic conduction, resulting in improved hydroxide conductivity. Accordingly, dual QAs-tethered azidated PPO in the form of OH− membrane (e.g., DQ-PPO-17-OH) showed high IC (1.35 to 1.70 mS cm−1), although the IC depends on the temperature.231 On the other hand, the ionic conductivity of the membrane increases with increasing temperature, such as 20 °C (19.1 to 40.5 mS cm−1) and 80 °C (35.0 to 63.9 mS cm−1) (Fig. 45a). Therefore, dual QAs, azidation degrees, and temperature are all three factors that influence the IC conductivity in the membranes. In a different study, the stronger WU character of PPO-iP (e.g., PPO-iP-44) yielded higher IC (64.8 mS cm−1).232 The higher IC might cause the membrane's higher WU and good microphase separation. Consequently, the low IC values of 60.4 and 59.6 mS cm−1 were produced using the low WU features of the membranes PPO-nP and PPO-CH.
![]() | ||
Fig. 45 Ionic conductivity as a function of the temperature of (a) DQ-PPO-x-OH-based membranes, reproduced with permission from ref. 231, Copyright 2021, Elsevier. (b) PTP-x-based membranes (PTP-75, PTP-85, and PTP-90), reproduced with permission from ref. 234, Copyright 2021, Elsevier. (c) PBPCL-based AEMs, reproduced with permission from ref. 288, Copyright 2021, Elsevier. (d) SEBSCn-x-based membranes, reproduced with permission from ref. 233, Copyright 2021, Elsevier. (e) Crosslinked and uncrosslinked PPO/AGO membranes, reproduced with permission from ref. 242, Copyright 2021, Elsevier. (f) QP(VBC-St)-MB/EA0.3 membrane, reproduced with permission from ref. 261, Copyright 2021, Elsevier. |
The IC's dependence on the membranes' capacity for water absorption has been proven. Li et al. reported that the hydrophilic extender built into the side chain was coupled to the PPO via a hydrophobic alkyl spacer (e.g., SDQEO) and had a high IC (16.9–40.1 mS cm−1) at 25 °C.113 However, at 80 °C, this AEM had a high IC (87.3 mS cm−1). It might be the case that alkyl spacers encourage increasing ions to high WU, leading to increased conductivity without dimensional change. Therefore, the membrane's dual function as a hydrophilic extender and hydrophobic spacer might improve its conductivity for energy generation. Due to the membrane constructed without an alkyl ether bond by Hu et al., the IC of the N-cyclic piperidiniums-functionalized PTP-based AEMs (e.g., PTP-90) was investigated as a function of both temperature and IEC.234 The membrane's conductivity (64.4–129 mS cm−1) rose when the temperature (from 20–80 °C) and IEC (from 2.0–2.52 mmol g−1) increased (Fig. 45b). The rapid water movement in the membranes caused the conductivity to improve as a function of temperature.
Poly(biphenyl piperidinium)-based AEM (PBPCL1.48) functionalized with high cation groups demonstrated better conductivity (64.2 mS cm−1) at 30 °C than PBPCL0.79 (31.4 mS cm−1) and PBPCL1.29 (51.8 mS cm−1), which have lower cation concentrations.288 It can be the case that AEMs with high cation concentrations will improve their hydrophilicity and support ion transport through self-aggregation (Fig. 45c). In contrast, it was found that the IC of the polymer composite PS/pTAP AEM with a 40:
60 (PS/pTAP (40
:
60)) ratio rose with rising temperature. However, at 40 °C, this kind of polymer composite did not show a significant IC (4.2 mS cm−1).237 Using a specific cation and extending the side chain with membranes, the IC of the membrane could not be enhanced. As a result, MPip-functionalized longer side chain-grafted SEBS membranes (e.g., SEBS-C1-MPy) exhibit little higher IC than those functionalized with other cations (MIm, MPy, and 2-methyl-1-pyrroline) for shorter side chains.233 SEBS-C1-MPy, however, had a lower IC (23 mS cm−1). It's low WU brought on the membrane's lower IC, and when the temperatures increased, the membrane's IC did not change as much (Fig. 45d).
The combination of aminopropylsilane-treated GO and QPPO (QPPO/AGO) exhibits good ion transport and conductivity.242 Additionally, the hydrophilic pathways that facilitate ion transport were exposed by the membrane containing aminopropylsilane-treated GO forms (Fig. 45e). As a result, the conductivity of pure Q-PPO (35.2 mS cm−1) was lower than that of the Q-PPO-AGO composite (83.8 mS cm−1). The AGO's hydrophilic characteristics might cause the composite membrane's improved conductivity. Due to the dispersion of water molecules and ion transport at higher temperatures in the membrane, IC generally increased with temperature. Similarly, IC examined a single antioxidant or two antioxidants (e.g., QP(VBC-St)-MB/EA0.3) with a quaternized membrane for functionalization.261 However, functionalizing one or two antioxidants with the quaternized membranes did not affect the membrane's IC (Fig. 45f). As a result, antioxidants did not aid in increasing the IC of the membranes. Recently, Ye et al. found that the IC of modified QP(VBC-St) membranes with various radical inhibitors (RI) (QP(VBC-St)-4-RI0.7) were lower than the IC (40.8–56.1 mS cm−1) of bare QP(VBC-St) membranes (68.7 mS cm−1) at 80 °C.262 The lower IC of the radically modified membranes was caused by the membranes' poorly organized microseparation morphology or the lack of hydroxide groups in the radical inhibitors or not having functional groups for OH− conduction. In contrast to radical-modified AEMs (40.8–53.7 mS cm−1), tempamine-based radical inhibitor-modified membrane (QP(VBC-St)-TPA0.7) and 4-tert-butylphenol-based radical inhibitor-modified membrane (QP(VBC-St)-4-TBP0.7) had a higher IC 56.1 and 53.7 mS cm−1, respectively.
By combining the membrane with organic structured polymers, the IC of the membrane might be increased. Recently, PVA was crosslinked with ABPBI (e.g., C-PVAf-ABPBI) polymer using glutaraldehyde (43 mS cm−1).131 This composite material performed better in IC than bare PVA crosslinked with glutaraldehyde (28 mS cm−1) and PVA crosslinked with ABPBI fiber (36 mS cm−1) at 60 °C. The higher IC of the composite membrane may be attributable to the separation of hydrophilic and hydrophobic molecules in a well-ordered manner to increase the conductivity of the membrane. VImPPO was combined with PVA in different ratios (9:
1 to 6
:
4) to prepare AEM (sIPN).128 In comparison to PVA, VImPPO has higher hydrophilicity. Thus, the conductivity of polymer networks with an adjusted ratio of VImPPO
:
PVA (sIPN-91 to sIPN-64) from 9
:
1 to 6
:
4 was evaluated. It was discovered that the composite IC (28.8–16.4 mS cm−1) dropped for the membrane's ratios from 9
:
1 to 6
:
4. It could be because VImPPO in more significant concentrations has more hydrophilicity, which allows for more water absorption and better conductivity. Additionally, reduced water absorption and conductivity were linked to a reduction in hydrophilicity when VImPPO concentration was lower. As a result, this investigation amply demonstrated that IC was dependent on the hydrophilicity of the membrane's ratio.
Due to the membrane's strong ion-conducting functional groups, membranes with higher IEC exhibited higher IC. The longer side chain BPTMA-tethered PBI (e.g., 48%BPTMA-PBI) had higher ion conduction groups up to this point, which resulted in higher IEC.187 Due to the increasing IEC data, the IC should also increase. As discussed and proved in the IEC section, these membranes have a high IEC value. It was concluded that increasing the percentage (27–48%) of BPTMA in the PBI membranes at 80 °C will result in a higher IC (8.3–29.8 mS cm−1).187 Additionally, the longer side chain 48% BPTMA of the QAs-tethered PBI's IC had a strong temperature response, increasing the IC values from 9.9 to 29.8 mS cm−1 with a temperature rise (30 to 80 °C). The 27% BPTMA-tethered PBI also exhibits a similar pattern. As a result, longer side chain-organized QAs may help to improve the IC for membranes. Interestingly, raising the temperature from 30 to 60 °C results in a corresponding rise in the IC of the LDH-modified quaternized polysulfone (e.g., qPSU-LDH5) from 19 to 60 mS cm−1.134 The membrane's IEC determines its IC. As a result, the crosslinking degrees of the PVBMPy composite (2.4–1.6 mmol g−1) with PSF decreased as they increased (10–30%).263 To a similar extent, crosslinking degrees (10–30%) reduced the IC (40–17.5 mS cm−1) of PVBMPy-PSF AEMs (PVBMPy-CL-PSF). It could cause a narrow constant ion conduction path and hamper the ionic migration caused by the PSF polymer chain.
Wang et al. demonstrated that the IC of the membranes is independent of the amount of crosslinking.265 The conductivity was raised to 10% of the crosslinked composite membranes (e.g., 10%-PBP-c-SEBS) and then dropped at 15% crosslinking, according to the IC analysis of increasing the crosslinking percentage (5–15%) of the PBP polymer to SEBS membranes. It may have contributed to the membranes' WU quickly dropping to 15% for crosslinked composite membranes. As a result, the IC of the membranes may depend on the WU. Similarly, at 80 °C, the IC of a functionalized PPO membrane with varying PEG composite membrane (e.g., 20PDM-PEG2000) percentages (0.5–5%) was investigated.255 At this point, increasing the PEG percentages up to 2% (e.g., 20PDM-2%PEG2000) raised the IC (85.7–97.2 mS cm−1), and then IC was lowered (Fig. 46a). The conductivity may be limited due to ion buildup at greater concentrations. The obtained IC for functionalized bare PPO was 73.2 mmol g−1. However, the inclusion of PEGs can improve the PPO of IC.
![]() | ||
Fig. 46 Ionic conductivity as a function of temperature of (a) 20PDM-x%PEG2000-based membranes, reproduced with permission from ref. 255, Copyright 2021, Elsevier. (b) Crosslinked QAs-PVBC/crosslinked PVA (PVA-xPQVBC30%)-based AEMs, reproduced with permission from ref. 257, Copyright 2021, Elsevier. (c) Crosslinked DBMHDA-PBBP (AEM-1, 2, 3), crosslinked BAPTES-PBBP (AEM-4, 5, 6) and BPPO-PBBP (AEM-7), and uncrosslinked QPBBP (AEM-0) membranes, reproduced with permission from ref. 301, Copyright 2022, Elsevier. (d) P(4PA-co-2PA)-x membranes, reproduced with permission from ref. 300, Copyright 2022, Elsevier. |
Water uptake in the membrane and IC are linked, meaning that IC rises with the water uptake. It is rarely seen that WU decreases while IC increases. The strategy was applied to the structure of glutaraldehyde-crosslinked PVA0.6–0.9QVBC30–50%.257 The IC values were increased and decreased in this composite structure by increasing the mass (0.6–0.9%) ratios of the two polymers (PVA and QVBC) as well as increasing the mass (30–50%) ratio of the DVB and the polymer (QVBC). Therefore, it was demonstrated that WU reduces when the mass ratio of polymers is increased as well as when crosslinkers (glutaraldehyde) are used with polymers (Fig. 46b). It has been shown that increasing the mass ratio of QVBC and PVA increases IC, but increasing the mass ratio of DVB and QVBC reduces IC. Similarly, Biancolli et al. observed higher WU, and the higher IEC of high-degree electron beam-treated VBC-grafted ETFE (e.g., 100-air-LT/RT) AEMs demonstrated low conductivity (113.5 mS cm−1), low WU, and low IEC of low-degree electron beam-treated VBC-grafted membranes (146 mS cm−1).299 Recent research has demonstrated that the high hydrophilic (WU) property of m-TPNPiQA (22.1 mS cm−1) had a higher IC than the lower WU properties of m-TPNPyQA (14.9 mS cm−1) and m-TPNBeQA (12.1 mS cm−1).240 m-TPNPiQA with a well-arranged microseparation morphology made of the hydrophilic CH3I cation and hydrophobic polymer backbone (m-TPN) to create a channel for the ions to conduct may have contributed to the higher IC.
The fluorinated segment in the polymer backbone is anticipated to increase the polymer's hydrophobicity, ultimately resulting in a higher degree of MSM. This configuration should be advantageous for developing an MSM and a highly efficient ion-conducting channel. Recently, cations such as DBMHDA, BAPTES, and the polymer BPPO crosslinked the fluorine-based fluorinated PBBP membrane.301 The DBMHDA-crosslinked PBBP (PBBP-DBMHDA) showed higher IC (74.7 mS cm−1) than BAPTES (60.0 mS cm−1), BPPO-crosslinked PBBP (36.0 mS cm−1), and bare PBBP (34.8 mS cm−1) because of its well-ordered microphase separation morphology, which provided more cations (Fig. 46c). Additionally, BAPTES-PBBP showed a morphology with well-separated hydrophilic and hydrophobic regions; nevertheless, no new cations were generated, and BAPTES silica structures blocked OH− from being transported, resulting in low WU and decreased IC. The example of BPPO-functionalized PBBP revealed that extremely low IC might be caused by lower cation generation or microphase separation morphology. As the molar ratio (0.24–0.63 mol%) of the whole unit increases, the IC (35–22 mS cm−1) of the copolymer-based AEMs (e.g., P(4PA-co-2PA)-24) decreases, which is an intriguing phenomenon to consider.300 Generally, excellent IC is accomplished by the microphase separation morphology of the membrane, i.e., the well-ordered hydrophilic and hydrophobic architecture supports the conductivity. However, as the concentration or molar ratio of the copolymer's unit rises, the conductivity falls because an increasing number of ions accumulate and block the ionic pathway, thus limiting ion conduction in the membranes (Fig. 46d).
The conductivity of the membrane might be improved in the presence of ions with high electronegativity and ionic clusters. Elevating the electronegativity and ionic clusters of fluorobenzoyl groups in the quaternized SEBS improved the membrane's IC at 8 °C.139 The quaternized SEBS with a mono- or penta-fluorobenzoyl group (e.g., HQA-F1-SEBS and HQA-F5-SEBS) added to it has a larger IC (71.2 and 87.0 mS cm−1) than when it is fluorobenzoyl-free (63.6 mS cm−1). Elevating the number of fluorobenzoyl groups from one to five resulted in a similar rise in electronegativity, and ionic clusters were responsible for the subsequent increase in IC. Due to the solid microphase separation, clustered cations functionalization with increased weight ratios (0.3–0.5%) of the backbones (e.g., 3QPAP-0.5) had higher conductivities (21.4–45.0 mS cm−1) than single cation-functionalized backbones (27.7 mS cm−1).295 The higher order of microseparation by clustered cations (hydrophilic) and polymer backbone (hydrophobic) structures may cause the rising conductivity. Increases in the monomer ratio and bromination degree (e.g., 60MePh-2.07) have recently been discovered to boost the membrane conductivity.287 The strong hydroxide conductivity may be attributed to the cationic group sites created with the increased monomer/bromide molar ratio.
In most cases, higher water intake corresponds to higher conductivity. However, the microphase separation in the membrane structures is crucial to the conductivity. Motz et al. agreed that the AEM's conductivity is not directly connected to its ability to absorb water but may result from microscopic phase separation structures.329 Polystyrene copolymer-based AEM has low IC values (50–61 mS cm−1) and high WU values (100–144%). Poly(phenylene) copolymer AEM (e.g., HTMA-DAPP), on the other hand, has a low WU (98%) and a high IC (90 mS cm−1). Compared to polystyrene copolymer-based AEM, the higher IC produced by poly(phenylene) copolymer-based AEM structure may be due to high microphase separation (hydrophilic–hydrophobic) morphologies. When the membrane is very hydrophilic, it has a high conductivity. Because of the high microphase separation and WU of this membrane structure, the PAPQ83 AEM exhibited a high conductivity (148.0 mS cm−1).294
Another research found that when the cross-linking ratio of TMHDA in a bication-functionalized membrane (e.g., m-XPTPA20-2N) increased from 10% to 40%, the conductivity reduced from 54.5 to 50.0 mS cm−1 at 80 °C.279 It may be due to lessening WU and a lack of order in the microphase separation that would occur with a higher cross-linking ratio. Developing a lengthy alkyl chain-containing hydrophilic cation-tethered fluorinated hydrophobic backbone promotes a well-ordered microphase separation structure (MSM) and characteristics that lead to high conductivity. Altering the cation ratio from low to high respectively acquired low to high dense amination group following quaternization, which allowed for very thick QAs. Increasing the concentration of QAs is essential for improving the AEM's conductivity. Accordingly, high MSM and conductivity (98.1 mS cm−1) were found in PAEK with a long alkyl carbazole side chain, followed by a tethering high density of QAs (e.g., PAEK-HQACz-0.7).110
Similarly, membrane IECs findings are linked to WU, SR, and IC. Accordingly, increases in the weight ratio of (vinylbenzyl) trimethylammonium chloride's (4–10%) increases the fluorinated (vinylbenzyl)trimethylammonium membrane (PVIB), which increases the IC (41.2–48.4 mS cm−1), leading to increases in both the IECs (1.43–1.82 mmol g−1) and the SR (13.2–20.5%).326 A higher SR and conductivity with IECs may be attributed to (vinylbenzyl)trimethylammonium chloride's hydrophilic character. Thus, the membrane's hydrophilicity is crucial to its conductivity. The membrane's microphase separation morphology (MSM) determined its conductivity. Because of this, increasing the long alkyl bis-piperidinium cation functionalization of the membrane did not result in a good MSM, meaning that not enough hydrophilic–hydrophobic segregation was produced.298 Better IC (27.7 mS cm−1) was achieved with a lower concentration of bis-cation-functionalized membrane (e.g., QPBPip33Ac) compared to the AEM with a greater concentration of bis-piperidinium cations (19.8–7.5 mS cm−1). It was discovered that the membrane has a structure of multi-cations and cross-linking, which accounts for its high conductivity.284 A higher ratio (30–80%) of multi-cations in the fluorinated cross-linked membrane (e.g., PHFB-VBC-DQ-80%) increased the conductivity in this study from 55.1 to 135.8% mS cm−1. Higher conductivity indicates that the membrane has a well-structured MSM.
In most cases, the hybrid membrane's hydroxide conductivity exceeded that of the bare membrane because of the orderly separation of hydrophilic and hydrophobic phases. Dimethylimidazolium (DIm)-functionalized fluorinated PBI bare membranes had a hydroxide conductivity of 50.2 mS cm−1. In contrast, Dim-functionalized fluorinated PBI-silicon hybrid membranes (PBI–DIm–Si hybrid) had a conductivity of 22.2 mS cm−1, as measured by Jheng et al.305 The poor conductivity was obtained because the microphase separation in the hybrid membrane did not develop correctly compared to the bare membrane. According to another research, membranes having a quaternized mixed block co-polymer structure (e.g., DMBA-PIM-b-PSF-0.91) had more conductivity than those without such a structure.307 They obtained a well-ordered MSM structure in the hybrid block-copolymer structure, consisting of fluorinated hydrophobic and hydrophilic polymers with increased conductivity. The conductivity of the copolymer-based AEM (e.g., QCPPAE-4/1) produced by Cha et al. was found to be improved by increasing the quantity of monomer (2–4%) while keeping the amount of trimer constant (1%).306 Increasing the monomer content has been shown to promote hydrophobic–hydrophilic separation for ion movement to boost the membrane's ability to conduct ions. The membrane's conductivity may be increased by developing it with cross-linking and increasing the quantity of the polymer ratio in the membrane's composition. Accordingly, an N-methyl imidazole-functionalized 50% fluorinated polymer membrane exhibited 5.4 mS cm−1, and following cross-linking (quaternization) by TMHDA, this value rose to 6.5 mS cm−1.278 When the fluorinated polymer ratio was increased to 75% (e.g., L-FPAEO-75-MIM), either with or without crosslinking, the results were 16.9 and 17.0 mS cm−1, respectively.
Perfluorinated polymers are designed with well-ordered microphase separation morphologies (MSM) to improve the conductivity. However, MSM's conductivity varies depending on the polymers from which it is made. Compared to quaternized Nafion (27 mS cm−1), quaternized Aquivion (e.g., Aquivion-TMA) was shown to have more excellent conductivity (38 mS cm−1).276 Aquivion AEM suggests that increased conductivity may be due to a shorter side chain, high IEC, and higher MSM. At 80 °C, the OH− form of polyfluorene-based AEMs (e.g., PFOTFPh-C6-TMA(OH)) was shown to have a higher conductivity (101–156 mS cm−1) than the Cl− form (52–133 mS cm−1), perhaps because the hydroxide form of the membrane had a higher hydrophilic–hydrophobic separation and higher water absorption.100
Higher IEC findings resulted in higher water uptakes related to high IC. As a result, an adjusting ratio (30–70%) of PDBA in the fabrication of multi-cation (NQQN)-functionalized fluorene-based AEMs (e.g., TQ-PDBA-30–70%) was attributed to increased IEC (1.09–2.1 mmol g−1) and WU (4.1–10.8%).283 At this stage, the membrane had a higher IC (78.8–136.2 mS cm−1). As a result, multi-cation-functionalized fluorene-based AEM may have good hydrophilic-hydrophobic separation, which would lead to good ion conduction across the membrane. In general, a higher WU membrane led to a higher IC. However, a recent study has demonstrated that IC depends not on WU but on microphase separation morphologies.282 As a result, the low WU property of single QAs-tethered PPO (35.2 mS cm−1) and fluorine-tethered PPO with tri-QAs (e.g., PPO-22-3QA8F) exhibited more excellent IC (75.5 mS cm−1) than the higher WU property of tri-QAs tethered PPO (23.5 mS cm−1). Thus, fluorine-tethered PPO-tri-QAs with high IC could achieve high microphase separation.
Wang et al. observed that QAs-NPBI with n-butyl functionalization (e.g., NPBI-QA55-B45) demonstrated 48.2% conductivity, which is still present after 1200 h in 1 M NaOH at 80 °C.188 In contrast, QAs-NPBI without functionalization showed an abrupt conductivity decrease after 50 h owing to OH− attack (Fig. 47a). The membrane's stability may benefit from n-butyl functionalization in an alkaline environment. The developed dual QAs side-chains were effective in improving the AS. As the amount of time decreased, the conductivity of the dual QAs-functionalized azidated PPO (e.g., DQ-PPO-17-OH) AEMs steadily reduced.231 The maximum initial conductivity loss (7.4%) was found for the optimized membrane (DQPPO-17-OH), indicating that the membrane had a high AS (Fig. 47b). The high ion content (dual QAs side-chains) may have contributed to the high AS. Liu et al. observed high alkaline stability for low WU characteristics of the PPO-CH membrane at 1 and 2 M NaOH at 60 °C with retained conductivity of 99.2% and 98.8%, respectively, for 216 h.232 Besides, the high WU characteristics of PPO-iP (e.g., PPO-iP-44) and PPO-nP showed high conductivity loss of 65% and 74%, respectively, at 2 M NaOH. The high AS of PPO-CH was due to high cyclohexyl substitution on QAs, which restrict the OH− ion attack. AEMs with an alkyl spacer (e.g., SDQEO) demonstrated more excellent chemical stability (77% conductivity was retained after eight days) than AEMs without one in 1 M KOH and at 60 °C due to the deteriorating effect of the electron-retreating backbone on the cations.113 Additionally, membranes are extremely susceptible to nucleophilic (OH−) attack without spacers because cations are close to the benzene ring, which raises the risk of OH− ion attack (Fig. 47c). At 80 °C under 1 M NaOH, N-cyclic piperidinium-functionalized PTP-based AEMs (e.g., PTP-90) had remarkable chemical stability, and after 934 h, 73% IC remained.234 The authors also noted good AS in a comparable PTP-based AEM with low WU and low IEC. The designed poly(biphenyl piperidinium)-based AEMs (PBPCL1.29) show high AS under 1 M NaOH at 80 °C, i.e., the membrane conductivity retained almost 87% after 480 h.288 This higher chemical stability of the AEM was due to the distinctive octopus-shaped piperidinium side chain assembly, which can ease the microphase separation, and the hydrophobic main backbone can diminish the hydroxide attack.
![]() | ||
Fig. 47 Alkaline stability as a function of time (a) NPBI-QA55-B45 and NPBI-QA55 membranes (1 M NaOH at 80 °C), reproduced with permission from ref. 188, Copyright 2021, Elsevier. (b) DQ-PPO-x-OH AEMs (2 M NaOH at 80 °C), reproduced with permission from ref. 231, Copyright 2021, Elsevier. (c) Side-chain tethered, alkoxyl extender combined with the flexible spacer on PPO-based AEMs (1 M KOH at 60 °C), reproduced with permission from ref. 113, Copyright 2021, Elsevier. (d) SEBS-x-y-based membranes (1 M KOH at 80 °C), reproduced with permission from ref. 233, Copyright 2021, Elsevier. (e) QP(VBC-St)-4-TBPx AEMs (8 M KOH at 80 °C), reproduced with permission from ref. 262, Copyright 2021, Elsevier. (f) Crosslinked VImPPO-PVA membrane (1 M NaOH at 80 °C), reproduced with permission from ref. 128, Copyright 2021, Elsevier. |
In the form of the hydrophilic ionomer and hydrophobic polymer (e.g., PS/pTAP (40:
60)), AEM obtained higher AS, i.e., 98% conductivity was retained after 300 h in 1 M NaOH at ambient temperature.237 Due to its role in maintaining the membrane's stability in alkaline circumstances, the functionalizing cation with the nonpolarizable ring-based structure in the membrane for the quaternization is crucial. Thus, at 1 M KOH at 80 °C, higher AS is demonstrated by MPy and Pip cations, which were utilized to quaternize longer and shorter side chains connected to the SEBS membrane (e.g., SEBS-C1-MPy), respectively.233 The maintained conductivity was about 87% and 84% after 600 h (Fig. 47d). Compared to other cations with functionalized side chains attached to SEBS, the heterocyclic cations (MPy and Pip) have good chemical stability and can protect the membrane from OH− attack. In another study, high chemical stability was observed in the composite PPO/aminopropyl silane-treated GO (PPO/AGO) under the environmental conditions of 5 M NaOH solution.242
Poly(4-vinylbenzyl chloride-styrene) membranes were doped with double antioxidants for the AS for the first time at 60 °C and 2 M KOH.261 As a result, diphosphite-based antioxidants-doped membranes (e.g., QP(VBC-St)-MB/EA0.3) showed significant AS (69.6% conductivity was retained after 1300 h) compared to bare AEM (42.5% conductivity was retained after 1300 h). In contrast, the doped membranes showed less stability (47.9% conductivity was kept after 1300 h) in 1 M KOH at 80 °C, while bare quaternized membranes revealed that just 21.2% conductivity was retained after 1000 h. Compared to a pure poly(4-vinylbenzyl chloride-styrene) membrane, these antioxidant-doped membranes exhibit more excellent stability, even at higher alkaline concentrations (2 M KOH). It could be the effect of a diphosphite-based antioxidant decomposing the intermediate byproduct and preventing the generation of free radicals. Radical inhibitor-modified QP(VBC-St) showed higher AS than bare QP(VBC-St) in 8 M KOH solution at 80 °C for 24 h.262 It was found to be 4-tertbutylphenol-based radical inhibitor-modified QP(VBC-St) AEM (e.g., QP(VBC-St)-4-TBP1.4) that showed higher AS (66%) than QP(VBC-St)-4-TBP0.7 (45%) and QP(VBC-St)-4-TBP0.1 (16%) AEMs (Fig. 47e). Therefore, the radical inhibitor could protect the QAs from the OH radicals. The authors also examined how more radical (OH−) ions were consumed when the concentration of 4-tertbutylphenol in the QP(VBS-St) membrane increased, thus significantly the protecting QAs. Coppola et al. investigated AS on the fiber form of a PVA-crosslinked ABPBI membrane (C-PVAf-ABPBI) in a 15% KOH solution at 30 °C and found that the membrane was more stable than bare PVA fiber or glutaraldehyde crosslinked PVA for up to 7 days.131 The ABPBI membrane may thereby improve the membrane's stability in an alkaline setting.
Yang et al. investigated AS for VImPPO:PVA semi-interpenetrated polymer networks (e.g., sIPN-91) in 1 M NaOH at 80 °C.128 The membrane's IC rose for the first two days before gradually declining as the free imidazolium ions in the NaOH solution degraded at higher temperatures (Fig. 47f). After 1000 h of soaking, the composite membrane maintained its IC of 66.5%. Similar to this, Tang et al. found that the AS dropped (100–65.3%) with an extended soaking period (0–240 h) of 48% BPTMA-tethered PBI (48%BPTMA-PBI) under 1 M KOH at 80 °C.187 It could be the cause of the nucleophilic assault that caused PBI's core to degrade. Thus, the longer side chain-tethered PBI membrane does not show much alkaline stability under high pH conditions because it declined the conductivity to 65% in a shorter time (240 h). However, by setting up a large QAs group, it may be possible to fend off OH− attack on the membrane. Interestingly, quaternized polysulfone deteriorated entirely in 360 h due to nucleophilic attack on the QAs groups in the backbone.134 It did not exhibit much AS at 1 M KOH and 80 °C. However, the positive charges of LDH electrostatically attracted the OH− ion to keep the backbone from being attacked by nucleophiles. They thus kept it stable for up to 400 h after adding the LDH composite (e.g., qPSU-LDH5). It may be best to combine positively charged elements to protect the polymer backbone from nucleophilic attack.
A considerable increase in alkaline stability may be achieved by comparing crosslinked AEMs to a pristine QAs-functionalized membrane. Most reports indicate that higher temperatures can hasten the disintegration of AEMs; the AS of the membranes at high temperatures is crucial. As a consequence, the alkaline stability of the membrane reduces as the temperature (60 and 80 °C) increases.263 Additionally, it lowered the AS of the membrane discovered in the TMHDA-crosslinked PVBMPy-PSF AEM and raised the cross-linking degree (15, 20, and 25%). It was shown that the membrane conductivity at 60 °C (27.5, 23.5, and 17.5 mS cm−1) and 80 °C (23.5, 19.8, and 15.0 mS cm−1) declines after 576 h. The hydrophilicity of the membranes may have decreased, and their QAs groups may have deteriorated with time, which would account for the AS decline. As a reason, both the temperature and degree of crosslinking affect the AS of the membrane. However, after 576 h, this crosslinked composite membrane (e.g., PVBMPy-CL-15%PS) showed stronger AS. Recently, it was found that the high AS of the PBP membrane only degraded by 9.1% at 1 M NaOH after 800 h.265 The membrane showed even higher AS and only degraded by 6.0, 4.7, and 3.2% after 800 h about 5, 10, and 15% PBP-crosslinked SEBS membranes such as 5%-PBP-c-SEBS, 10%-PBP-c-SEBS, and 15%-PBP-c-SEBS, respectively. Therefore, in highly alkaline environments, the crosslinking degrees improve the membrane AS.
Interestingly, functionalized PPO was combined with different molecular weights (400–4000) of PEG and tested for AS under 1 M KOH at 80 °C.255 With the increase in the AS, i.e., the conductivity of the composite membrane was maintained (61–68%) after 300 h by increasing the molecular weight (400–2000) of PEG (e.g., 20PDM-2%PEG2000). The 4000PEG composite with PPO was observed to have lower maintained conductivity (52%). The undeveloped microstructure (microseparation morphology) patterns may be caused by the diminishing retained conductivity of PEG at increasing molecular weights. Thus, this work demonstrates that AS depends on the structure of the membranes' microseparation. In another study, in 1 M or 6 M NaOH at 80 °C, PVA-incorporated crosslinked QVBC composite was evaluated for the AS. After 216 h, the initial conductivity decreased from 45 to 17 mS cm−1.257 Later, if the period was extended to 1584 h, the initial conductivity (17 mS cm−1) remained unchanged. After 96 h in 6 M NaOH, a second examination of the composite membrane revealed that the conductivity had dropped to 18 mS cm−1. The crosslinked QVBC combined with a PVA membrane (e.g., PVA-0.9PQVBC30%) exhibits improved AS at a mole concentration of 1 M NaOH but cannot withstand a mole concentration of 6 M NaOH.
AEM with VBC grafts recently subjected to electron beam irradiation at low or RT with a dosage of 100 kGy (e.g., 100-air-LT/RT) did not display chemical stability (high conductivity loss) when subjected to 1 M KOH at 80 °C.299 However, on irradiation with an electron beam at a dosage of 40 kGy, while the procedure was carried out at ambient temperature, demonstrated remarkable chemical stability (significantly less conductivity loss). The better chemical stability provided by a dosage of 40 kGy of AEM has considerable flexibility compared to a high-dose beam (100 kGy) of AEM, resulting in a high chain scission reaction. This flexibility may protect the backbone against attacks by hydroxide ions. AEMs treated with an electron beam have the same level of chemical stability as AEMs that have been treated chemically.
The problem of developing AEMs for energy storage and conversion devices concerns the long-term durability of the membranes in alkaline environments. The breakdown of AEMs occurs in an alkaline environment because the cationic groups and backbones are susceptible to OH− ions at high temperatures. As a result, it is crucial to develop AEMs functionalized with very stable cations. As a result, N-cyclic-based cations are very stable at high temperatures and in very alkaline environments. Wang et al. constructed m-TPN AEMs and functionalized them with various N-cyclic cations to create m-TPNPiQA, m-TPNPyQA, and m-TPNBeQA AEMs, which were then used to test an AS at 80 °C with 1, 2, and 5 M NaOH.240 In contrast to m-TPNBeQA, m-TPNPiQA and m-TPNPyQA were shown to have higher AS. As the molarity of NaOH increases, the initial IEC values of m-TPNPiQA and m-TPNPyQA are slightly reduced. When the alkaline concentration was increased from 1 to 5 M NaOH, m-TPNBeQA diminished the IEC values rapidly. The AS of m-TPNBeQA was decreasing, which may have been caused by the cation and backbone being attacked by nucleophiles, causing it to degrade.
Comparing crosslinked AEMs to bare AEMs, increased AS was seen. Compared to BAPTES and BPPO-crosslinked PBBP and bare BPPO under 2 M NaOH at 25 °C, DBMHDA-crosslinked PBBP (PBBP-DBMHDA) demonstrated more excellent AS.301 However, the conductivity of the crosslinked membranes rapidly diminished over time as the cations in the crosslinking (DBMHDA) were increasingly prone to attack by hydroxide ions. BAPTES, on the other hand, has silica (Si–O–Si) built into its structure by hydrolysis, which surrounds the functional group and shields it from OH− attack, resulting in shallow conductivity loss. However, a high IC was found in crosslinked membranes after 720 h in 2 M NaOH, compared to bare and BPPO crosslinked PBBP. The high AS of the copolymer-based membranes (e.g., P(4PA-co-2PA)-24) under 1 M NaOH at 80 °C was found by Jiang et al., with no change in the conductivity after 720 h.300 However, after being exposed to 5 M NaOH for 720 h, these copolymers-based AEMs were shown to be degraded (Fig. 48a). The n-alkyltrimethylammonium (QAs)-functionalized copolymers degrade in a 5 M NaOH environment. Still, they are stable in 1 M NaOH conditions. Additionally, 1H NMR results exhibit vinylic peaks, which signals that Hofmann elimination occurring in the pentyl spacer was used to demonstrate the stability of the membrane following 5 M NaOH treatment.
![]() | ||
Fig. 48 Alkaline stability profile of (a) poly(p-quaterphenylene alkylene)-based AEM (P(4PA-co-2PA)-47), reproduced with permission from ref. 300, Copyright 2022, Elsevier. (b) Hexyl QA-and fluorobenzoyl-tethered SEBS-based membranes, reproduced with permission from ref. 139, Copyright 2022, Elsevier. (c) Poly(meta-terphenylene alkylene)-based membrane (m-XPTPA40-2N), reproduced with permission from ref. 279, Copyright 2021, Elsevier. (d) Crosslinked end-designed oligomer (PHFB-VBC-DQ-80%) membrane, reproduced with permission from ref. 284, Copyright 2021, Elsevier. |
It is important to note that at 1 M KOH and 80 °C, the AS of fluorine-substituted (mono and penta-fluorobenzoyl)-quaternized SEBS membrane was higher than that of fluorine-free quaternized membrane.139 After 500 h, the conductivity of the mono and pentafluorine-substituted (e.g., HQA-F1-SEBS and HQA-F5-SEBS) membranes were still 93.8 and 96.6%, respectively, far higher than that of the fluorine-free quaternized membrane (88.1%). These findings suggest that the fluorobenzoyl group has high alkaline stability and protects the polymer backbone from nucleophilic attack (Fig. 48b). Therefore, membranes attached to fluorine moieties would be helpful for their increased chemical stability over the long term. Poly(aryl piperidinium) functionalized with piperidinium-based clustered cations (e.g., 3QPAP-0.5) was found to have high chemical stability under 1 M NaOH at 80 °C. Minor conductivity loss was observed from 0 to 144 h (45.0–40.1 mS cm−1); later, it was stable up to 720 h (40.1 mS cm−1).295 Due to clustered cations in the membrane protecting the backbone from OH− assault, excellent stability of the membrane was achieved. However, poly(aryl piperidinium) functionalized with a single cation similarly experienced a loss in conductivity. As a result, piperidinium-based cations, whether in a cluster or single form, improve the chemical stability in an alkaline environment.
Low WU resulting in low IEC may protect membranes from the attack of OH− ions in an extreme environment, which may explain why membranes with low IEC values have shown good alkaline stability. Cross-linking and adapting to low IEC values might help to improve the alkaline stability (AS). However, with 1 M NaOH at 80 °C, Chen et al. found that AEMs (e.g., 60MePh-2.07) with higher and lower IEC values displayed higher and lowered AS; therefore, the conductivity test in an alkaline media shows a slight decrease in these membranes.287 Compared to poly(phenylene) copolymer-based AEM, which showed modest conductivity variations (119 to 88 mS cm−1), the polystyrene copolymer-based AEM (e.g., HTMA-DAPP) exhibited very high AS with the initial conductivity (63 to 64 mS cm−1) not changing even after 500 h under 1 M NaOH.329 Unlike other AEMs, the AS values for these two copolymer-based AEMs were much higher (Table 6).
AEMs | IEC (mmol g−1) | WU (%) | SR (%) | IC (mS cm−1) | AS | λ | Performance | Ref. |
---|---|---|---|---|---|---|---|---|
a AEMs = anion exchange membranes; IEC = ion exchange capacity; WU = water uptake; SR = swelling ratio; IC = ionic conductivity; AS = alkaline stability; λ = hydration number. | ||||||||
PQP-100 | 2.3 | 20.0 | 20.6 | 56.7 | Medium (1 M NaOH) | 4.8 | Good | 235 |
PEGDA | 0.19 | 83.0 | — | 1.2/0.8 | Not investigated | 0.5 | Moderate | 260 |
NPBI-QA55-B45 | 1.56 | 39.0 | 10.0 | 31.7 | Low (1 M NaOH) | 12.1 | Good | 188 |
DQ-PPO-17-OH | 1.7 | 60.0 | 12.7 | 40.5 | High (2 M NaOH) | — | Good | 231 |
PPO-iP-44 | 1.9 | 192.0 | 45.0 | 64.8 | Medium (2 M NaOH) | 56.1 | Better | 232 |
SDQEO | 1.13 | 120.1 | 33.2 | 40.1 | Medium (1 M KOH) | — | Good | 113 |
PTP-90 | 2.52 | 39.7 | 15.4 | 64.4 | High (1 M NaOH) | 8.7 | Better | 234 |
PBPCL1.48 | 1.5 | 46.4 | 20.8 | 64.2 | Not investigated | — | Better | 288 |
PS/pTAP (40![]() ![]() |
1.2 | 8.6 | — | 4.2 | High (1 M NaOH) | — | Moderate | 237 |
SEBS-C1-MPy | 1.13 | 18.0 | 10.50 | 23.0 | High (1 M KOH) | 17.0 | Good | 233 |
PPO/AGO | 2.2 | 37.9 | 9.80 | 83.8 | High (5 M NaOH) | — | Better | 242 |
QP(VBC-St)-MB/EA0.3 | 1.0 | 33.8 | 39.6 | 55.0 | High (2 M KOH) | — | Better | 261 |
QP(VBC-St)-4-TBP0.7 | 1.44 | 39.0 | 41.0 | 53.7 | Medium (8 M KOH) | Better | 262 | |
C-PVAf-ABPBI | — | 55.0 | 27.0 | 43.0 | Medium (KOH) | — | Good | 131 |
sIPN-91 | 1.46 | 48.69 | 20.61 | 28.8 | High (1 M NaOH) | 16.41 | Good | 128 |
48%BPTMA-PBI | 1.80 | 57.0 | 43.0 | 29.8 | Low (1 M KOH) | — | Good | 187 |
qPSU-LDH5 | 0.88 | 58.0 | — | 60.0 | High (1 M KOH) | — | Good | 134 |
PVBMPy-CL-15%PSF | 2.1 | 40.0 | 26.2 | 33.0 | High (1 M NaOH) | — | Good | 263 |
10%-PBP-c-SEBS | 2.87 | 94.1 | 11.1 | 81.6 | High (1 M NaOH) | — | Better | 265 |
20PDM-2%PEG2000 | 2.09 | 84.7 | 74.2 | 97.2 | Medium (1 M KOH) | — | Better | 255 |
PVA0.9PQVBC30% | 1.61 | 123.7 | 21.9 | 141.7 | Medium (1 M NaOH) | Best | 257 | |
100-air-LT/RT | 2.62 | 256 | — | 113.5 | Low (1 M KOH) | 54 | Best | 299 |
m-TPNPiQA | 2.54 | 52.28 | 21.14 | 22.11 | Medium (1 M NaOH) | 10.92 | Good | 240 |
PBBP-DBMHDA | 2.51 | 53.5 | 14.3 | 74.7 | High (2 M NaOH) | — | Better | 301 |
P(4PA-co-2PA)-24 | 2.17 | 18.8 | 3.6 | 35.0 | High (1 M NaOH) | 4.5 | Good | 300 |
HQA-F5-SEBS | 1.45 | 136.7 | 36.8 | 87.0 | High (1 M KOH) | — | Best | 139 |
3QPAP-0.5 | 2.26 | 26.0 | 9.2 | 45.0 | High (1 M NaOH) | 6.4 | Better | 295 |
60MePh-2.07 | 2.07 | 130.0 | 35.0 | 45.0 | Medium (1 M NaOH) | — | Better | 287 |
HTMA-DAPP | 1.5 | 98.0 | — | 90.0 | High (1 M NaOH) | — | Best | 329 |
PAPQ83 | 2.29 | 105 | — | 148.0 | High (1 M KOH) | — | Best | 294 |
m-XPTPA20-2N | 2.68 | 140.0 | — | 52.3 | — | 14.7 | Better | 279 |
PAEK-HQACz-0.7 | 1.88 | 46.4 | 13.5 | 98.1 | High (4 M NaOH) | — | Better | 110 |
PVIB-10 | 1.82 | — | 20.5 | 48.4 | Not investigated | — | Good | 326 |
QPBPip33Ac | 2.84 | 53.3 | 23.5 | 29.7 | Medium (2 M HCl) | 10.4 | Good | 298 |
PHFB-VBC-DQ-80% | 2.24 | 50.7 | 26.1 | 135.8 | High (2 M NaOH) | 12.57 | Best | 284 |
PBI–DIm–Si hybrid | 1.87 | 30.9 | 10.0 | 22.2 | High (2 M NaOH) | 9.2 | Good | 305 |
DMBA-PIM-b-PSF-0.91 | — | 75.0 | 17.3 | 52.5 | Medium (1 M NaOH) | — | Better | 307 |
QCPPAE-4/1 | 1.75 | 40.0 | 23.0 | 65.0 | Low (1 M NaOH) | — | Better | 306 |
L-FPAEO-75-MIM | 1.80 | 95.0 | 55.0 | 17.5 | Not investigated | — | Good | 278 |
Aquivion-TMA | 1.06 | 23.0 | 5.0 | 38.0 | Low (1 M NaOH) | — | Good | 276 |
PFOTFPh-C6-TMA(OH) | 3.2 | 41.0 | 97 | 156.0 | High (8 M NaOH) | — | Best | 100 |
TQ-PDBA-70% | 2.16 | 10.84 | 3.23 | 136.27 | High (1–10 M NaOH) | 3.13 | Best | 283 |
PPO-22-3QA8F | 1.76 | 10.2 | 2.3 | 75.5 | (1 M NaOH) | 3.4 | Good | 282 |
Similarly, after 2000 h in 1 M KOH at 100 °C, the AS of poly(arylene piperidinium)-based AEM (PAPQ83) was high.294 The piperidinium-based cation may defend the skeleton from the hydroxide attack. Recently, the bication-functionalized poly(meta-terphenylene alkylene) (m-PTPA-2N) quaternized by varied ratios (10–40%) of TMHDA was found to have high AS.279 It was found that the AS of TMHDA cross-linked m-PTPA-2N with a higher ratio of 40% (e.g., m-XPTPA40-2N) was greater than that of uncrosslinked m-PTPA-2N and cross-linked with a lower percentage (10–30%) membranes such as m-PTPA10-2N, m-PTPA20-2N, and m-PTPA30-2N (Fig. 48c and its inset picture). After 300 h, only slight variations were discovered in the membrane's conductivity, and it was still able to maintain a high level of stability even after 1000 h. The quaternary ammonium groups and the polymer backbone could be protected from OH− assault by the bication structure.
Polymer backbones and QAs may be protected from nucleophilic attack by fabricating a membrane with a high cloud long alkyl chain-containing cations. Liu et al. found that the fluorinated membrane (PAEK) was prepared by the quaternization of a polymer composed of long alkyl chains-containing carbazole-based cations.110 Even after 7 days in 4 M NaOH at 80 °C, the produced AEM (e.g., PAEK-HQACz-0.7) exhibited good alkaline stability (AS). The chemical stability and membrane protection afforded by a long alkyl chain enclosed inside a cation structure may be enhanced throughout extended use. The chemical stability of the bis-cation-functionalized membrane (e.g., QPBPip33Ac) was investigated by Liu et al. for 3 days in 2 M HCl.298 In an acidic environment, its effectiveness is minimal. A decrease in the membrane mass was linked to a rise in bis-piperidinium cation concentration. Since the membrane contains the piperidinium cation, which shields the QAs from the attack of the hydroxide ions, it may be very stable in NaOH or KOH environments. Intriguingly, the multi-cation-based compound-functionalized fluorinated cross-linked membrane (e.g., PHFB-VBC-DQ-80%) was discovered to have a high AS.284 After 500 h in 2 M NaOH, the membrane retains 92.9% conductivity. The membrane's multi-cation structure may be what protects the polymer backbone and QAs from nucleophilic assault (Fig. 48d).
Under 2 M NaOH at 80 °C for 120 h, the membranes PBI–DIm and hybrid PBI–DIm–Si were discovered to have a high AS.305 Both membranes demonstrated 100% loss-free conductivity. It is possible that the decreased nucleophilic assault of OH− ions to DIm, especially for the membrane PBI–DIm, is due to sufficient water in the region of cationic side groups that dilute the hydroxide ions. Therefore, significant AS was also seen in silicon AEM hybridized with PBI–DIm (PBI–DIm–Si hybrid). However, if the authors investigated the AS for at least 500 h, the results may have been more impressive when the membranes were ready. The conductivity with time was investigated for the block copolymer-structured quaternized fluorinated PSF hybrid with hydrophilic PSF (e.g., Q-PIM-b-PSF-0.91) under 1 M NaOH at 60 °C for 216 h.307 As the time passed, the conductivity decreased; after 216 h, the conductivity stayed at 80.0%. The membrane's reduced conductivity might be due to OH− attack on polymer backbones and cations that cause the membrane to deteriorate progressively. Using an optimized ratio (2:
1 and 4
:
1) of monomer to trimer, Cha et al. found that quaternized copolymer-based AEMs made in this way exhibited moderate chemical stability.306 The conductivity (48–35 mS cm−1) decreased steadily with an increase in the days (1–7) for the 4
:
1 ratio of AEM (e.g., QCPPAE-4/1), indicating that the produced AEMs were not stable under 1 M NaOH at 60 °C. Alternatively, a 2
:
1 AEM ratio leads to 24.5–18.2 mS cm−1 conductivity for 1–7 days. The QAs groups may be degraded by nucleophilic assault, which may account for the lowering of conductivity. However, the copolymer-based AEM with a greater monomer concentration exhibited a higher AS than the one with a lower monomer level.
AS testing was conducted with 0.5 and 1 M NaOH at 25 °C on perfluorinated polymers such as quaternized Nafion (Nafion-TMA) and Aquivion (Aquivion-TMA).276 Intense hydroxide ion assault causes quaternized perfluorinated polymers to deteriorate rapidly, leading to a significant drop in IECs. These membranes might be functionalized with a solid cation to shield them against hydroxide ion attack. High molecular weight polyfluorene-based AEM (e.g., PFOTFPh-C6-TMA(OH)) was resistant to 8 M NaOH solution at 80 °C for 168 h.100 With the progression of time, the conductivity loss slowed down. However, 88% of the original conductivity remained there after 168 h. High resistance to hydroxide attack that led to high AS in hostile conditions suggest that polyfluorene-type AEM might be viable membranes. Accordingly, 1000 h of testing with 1–10 M NaOH at 80 °C revealed that the membrane, multi-cation ionic liquid-functionalized fluorene-based (e.g., TQ-PDBA-70%) AEM exhibited remarkable chemical stability.283 A steady decrease in conductivity (from 99% to 59%) was seen over 1000 h when the NaOH concentration was increased from 1 to 10 M. It is possible that the fluorine-based main chain's strong hydrophobicity, which can prevent the assault on the main backbone from OH−, contributes to the improved chemical stability of these membranes. Similarly, the membrane in the form of fluorine-tethered PPO crosslinked with tri-QAs demonstrated improved alkaline stability.282 For 7 days, the AEM showed good strength in 1 M NaOH at 80 °C. Even after 7 days, the conductivity remained at 99%. In contrast, without fluorine in the PPO-based AEMs, conductivity was preserved at 55% and 60% after OH− assault on the polymer backbone. However, fluorine-functionalized PPO-tri QAs (e.g., PPO-22-3QA8F) demonstrated that solid chemical stability and fluorine could protect the backbone against OH− assault. The characteristics of several AEMs are shown in Table 6, along with an overview of how well they performed in the IEC, WU, SR, IC, and AS and their hydration numbers (λ).
The YM results (2134, 2105, and 2127 MPa), EB (15.8, 17.3, and 14.6%), and TS (83.9, 81.4, and 82.5 MPa) demonstrated the superior performance of the aromatic monomer used to prepare 30 μm thickness of PQP-100, PBP-67, and PTP-83 AEMs, and these findings show the rigid backbones. The high molecular weight of the PQP-100 membrane obtained better mechanical properties at 30 °C.235 The membranes of the hydrophobic n-butyl group-tethered quaternized naphthalene-based PBI (e.g., NPBI-QA55-B45) were studied for TS and EB.188 The n-butyl group concentration in NPBI-Q increased from 20 to 75%, and TS (19.1–47.3 MPa) and EB (5.4–17.6%) increased. In addition, without n-butyl, Q-NPBI displayed lower TS and EB values, even though the number of QAs (25–100%) were increased due to the accumulation of highly hydrophilic (QAs), which leads to a defect in the mechanical strength of the membrane; there is no improvement in the TS (40.2–14.5 MPa) and EB (12.7–2.3%). Therefore, increasing or adding n-butyl's hydrophobic characteristics might improve the membranes' mechanical qualities. In a different investigation, completely hydrated dual QAs-treated azidated PPO-OH AEMs (e.g., DQ-PPO-17-OH) were found to have low TS (7.2 to 10.9 MPa) and high EB (19.1 to 31.8%).231 Given that the membrane was tethered with dual QAs, the obtained low TS may cause the membrane's high WU and high SR (high hydrophilic). High levels of WU and SR may cause the polymer backbone chains to get entangled, thus reducing the membrane's stiffness (tensile strength).
Liu et al. investigated the mechanical characteristics of QAs-functionalized PPO (e.g., PPO-iP-44) that included high and low cations.232 In contrast to less cation-containing QAs-functionalized PPO such as PPO-iP-44 (1599.2 MPa (YM), 34.9 MPa (TS), 2.4% (EB)) and PPO-nP-44 (1437.8 MPa (YM), 27.6 MPa (TS), and 2.0% (EB)), the highly bulky QAs-functionalized PPO (PPO-CH-44) demonstrated low YM (1212.5 MPa), TS (8.6 MPa), and EB (1%). As a result, bulky cations containing QAs caused the membrane's mechanical characteristics to decline. In contrast, fewer cations containing QAs could benefit the membrane's mechanical properties and could be the reason for less water uptake. The PPO-based AEM prepared with a hydrophilic alkoxyl extender side chain and a hydrophobic alkyl spacer (e.g., SDQEO) was found to have a low TS (10.5 MPa) and a high EB (27.3%) due to the longer side chain and the spacer's higher free volume, which reduced the inter-chain forces between the main backbone and side-chain and resulted in lower values being obtained.113 In addition, the absence of a spacer and a smaller side chain synthesized by PPO-based AEM demonstrated stronger TS. It lowered the EB because the smaller free volumes strengthen the main backbone and side-chain inter-chain forces. Due to the high viscosity of N-cyclic piperidinium, functionalized PTP-based AEM (e.g., PTP-90) showed moderate TS (29.2 MPa), low EB (6.8%), and noticeable YM (466.7).234 The TS and EB of the poly(biphenyl piperidinium)-based AEMs (e.g., PBPCL1.29) decreased and increased as the cation concentration increased in the membrane.288 It is possible to assert that the mechanical stability reduction contributed to the membranes' high IEC tendencies to increase the WU.
The mechanical stability of composite membranes with polysulfone/pTAP ratios of 60%:
40% (PS/pTAP (60
:
40)) and 40%
:
60% (e.g., PS/pTAP (40
:
60)) was examined by Arunachalam et al.237 The TS was reduced (22.5 to 12.5 MPa), and the EB was raised (1.8 to 2.5%) due to the rising ratio of ionomer pTAP (Fig. 49a). It could be because the hydrophilicity of the ionomer content increased the water absorption, causing the TS to drop and the EB to rise. As a result, an excess of the ionomer (pTAP) can degrade the mechanical characteristics of the PS/pTAP AEM. Higher WU and high SR are related to the membrane's mechanical strength deterioration. Accordingly, Yu et al. found that more extended side-chain-functionalized membranes (e.g., SEBS C1-x and C6-x) with higher WU and SR had lower mechanical strength.233 In contrast, shorter side-chain-functionalized membranes with lower WU and SR had better mechanical strength (Fig. 49b). According to Peng et al., doping one or two antioxidants in the AEMs (e.g., QP(VBC-St)-MB/EA0.3) improved the mechanical characteristics of the membranes.261 Accordingly, increasing one or more doping antioxidants led to intermediate results for the quaternized P(VBC-St) of mechanical strength, such as TS (8.2–10.3 MPa), EB (10.8–4.7%), and YM (374–773 MPa). The rising TS and YM and falling EB demonstrated the stiffness characteristics of antioxidants. When compared to the bare membrane QP(VBC-St) (10.7 MPa), the QP(VBC-St) membranes' tensile strength (6.6 to 4.6 MPa) reduced with the addition of radical inhibitors (e.g., QP(VBC-St)-4-TBP0.7).262 At the same time, the membrane's flexibility (EB) increased with the addition of the radical inhibitors (8.1 to 19.0%) compared to the bare (QP(VBC-St)) membrane (6.4%). Therefore, a radical inhibitor might increase the membrane's flexibility for applications.
![]() | ||
Fig. 49 Mechanical stability test of (a) various percentages of PS/pTAP composite membranes, reproduced from ref. 237, Copyright 2021, Springer Nature. (b) SEBS C1-x and C6-x-based membranes, reproduced with permission from ref. 233, Copyright 2021, Elsevier. (c) Crosslinked VImPPO-PVA (sIPN-x) and uncrosslinked VImPPO-PVA (n-64) membranes, reproduced with permission from ref. 128, Copyright 2021, Elsevier. (d) PVBMPy-CL-x%PSF membrane, reproduced with permission from ref. 263, Copyright 2021, Elsevier. |
It is interesting to note that structural matter can affect the mechanical stability of an AEM. As a result, PVA was crosslinked with the ABPBI polymer by glutaraldehyde to create nanofiber (e.g., C-PVAf-ABPBI) and nanosheet forms utilizing the electrospinning and immersion processes, respectively.131 In comparison to the nanosheet form (TS (2.8) and YM (64.0 MPa)), the fiber form had more excellent TS (5.0 MPa) and YM (103 MPa) values. It was discovered that the fiber form of the membrane had more stiffness than the sheet form. Additionally, it was revealed that the fiber membrane's flexibility (EB) was lower than that of the sheet kind of membrane (16.0%), at just 6.0%. Therefore, the membrane's structure firmly approached its stiffness and flexibility. Interestingly, PPO and PVA's stiff and flexible properties were combined in a composite (VImPPO:
PVA) with a range of ratios, from 9
:
1 to 6
:
4, to create an AEM (sIPN 91 to sIPN 64).128 By decreasing the PPO and increasing the PVA ratios, the membrane's TS (30.8–20.9 MPa) was reduced, and the flexibility (EB) (20.5–60.2%) was raised (Fig. 49c). Because the stiffness characteristics of PPO decreases with an increase in the flexible features of PVA in the composite form, the membrane had a low stiffness and high flexibility property. The VImPPO-PVA-mixed type demonstrated moderate mechanical stability. Therefore, PPO may increase the membrane's stiffness, while PVA was utilized to increase the membrane's flexibility.
If the mechanical characteristics are inadequate, the AEM will rupture, and the application will be unsuccessful. As a result, the AEM should need mechanical strength. The mechanical strength of various percentages (27 and 48%) of longer side chains-structured BPTMA-tethered PBI was recently investigated and compared to bare PBI membranes.187 Due to the longer side chain's-grafted membrane that interrupts the hydrogen bonding connection and subsequent increase in the macromolecular chain's distance from the backbone, the mechanical strength of the longer side chain-grafted PBI (e.g., PBI-27%BPTMA) was lower than that of bare PBI. However, the BPTMA percentage increase to 48% was attributed to lower YS, TS, and EB. The PBI membrane's stiffness and flexibility decreased as the grafting degree percentage increased. In contrast to grafted membranes, PBI, which has a complex structure, exhibits excellent stiffness and flexibility. Crosslinking and a stiff polymer backbone structure may improve the membrane's MS.
Interestingly, AEMs with a composite rigid and flexible polymer structure might improve the MS of the membranes. As a result, flexible polymer pristine membrane (SEBS) has low TS and greater EB, whereas stiff pristine polymer (PBP) exhibits high TS and lower EB.265 The benefits of these two membranes allowed for the formation of composites (e.g., 10%-PBP-c-SEBS) with a 5–15% degree of crosslinking. They were tested for MS. PBP's degree of crosslinking was increased from 5 to 15% when composited with SEBS membranes, demonstrating how the amount of PBP makes the composite membranes stiffer and causes the TS to increase while the EB decreases. Due to PBP's rigid polymer structure, membranes may be stiffer and crosslinked. Recent research has demonstrated that improving the MS may be accomplished by combining an increasing hydrophilic content with a hydrophobic polymer backbone (e.g., 20PDM-2%PEG2000).255 As a result, different percentages (0.5–50%) of PEG2000 composited with PPO were examined for the MS. It was demonstrated that as the PEG percentages were raised, the EB rose as well. However, the TS and YM values decreased as the PEG concentration increased because high WU caused the TS and YM to drop and the EB to increase. PEG can increase the flexibility of membranes by adding to the polymer backbone. Crosslinked PVBMPy-PSF composite membrane thus shows higher MS than that of bare PVBMPy.263 Additionally, increasing the crosslinking (10–30%) strengthens the membrane's MS (1.3–22.6 MPa). The growing MS of the membrane may lead to increased crosslinking, and the addition of PSF reduced the WU and SR (Fig. 49d).
Recently, the TS of the membrane reduced while the mass ratio of the two polymers, PVA and QVBC, grew (e.g., PVA-0.6PQVBC40%).257 However, an increase in the TS is shown by increasing the mass ratio of the crosslinker (DVB) with QVBC (Fig. 50a). The increase in the hydrophilic nature of the PVA ratio that led to the membrane's weakening might be caused by the declining TS. Increasing hydrophobic QVBC with the crosslinker DVB was the cause of the rising TS. As a result, the membrane's mechanical strength may be improved by the polymer's hydrophobic properties. The mechanical stability of a PBBP membrane based on fluorene (e.g., BPPO-PBBP) was investigated by Zhou et al.301 Comparatively, the TS (18.2 MPa) and EB (10.3%) of BPPO-crosslinked PBBP were higher than those of BAPTES (14.3 MPa and 7.7%) and DBMHDA (11.9 MPa and 4.9%)-crosslinked PBBP and bare PBBP (1.53 MPa and 0.9%). This crosslinked membrane's high TS and EB may be attributed to BPPO and PBBP being very hydrophobic and hydrophilic. In addition, the low TS and EB values obtained with DBMHDA can be attributed to the presence of hydrophilic cations in the structure. In contrast, the presence of silica in the BAPTES's structure could contribute to greater mechanical strength, and the presence of hydroxyl groups makes the membrane more adaptable.
![]() | ||
Fig. 50 Stress and strain analysis of (a) PVA-xPQVBC40%, reproduced with permission from ref. 257, Copyright 2021, Elsevier. (b) Hexyl QAs-and fluorobenzoyl-tethered SEBS membranes, reproduced with permission from ref. 139, Copyright 2022, Elsevier. (c) Poly(aryl piperidinium)-tethered single and clustered cations (1QPAP and 3QPAP-x) membranes, reproduced with permission from ref. 295, Copyright 2022, Elsevier. (d) Quaternized poly(styrene-ethylene-styrene) (SES-TMA-x), HTMA-DAPP and reinforced Durion membrane, reproduced from ref. 329, Copyright 2021, Royal Society of Chemistry. (e) Quaternized poly(p-phenylene) and poly(arylene ether)-based membranes (QCPPAE-x/y), reproduced with permission from ref. 306, Copyright 2017, Elsevier. (f) Crosslinked TQ-PDBA-x membranes, reproduced from ref. 283, Copyright 2020, Royal Society of Chemistry. |
The rigidity of quaterphenylene alkylene may be advantageous in the process of improving the membrane's MS. Recent research has shown that biphenylene alkylene-based AEM (e.g., P(4PA-co-2PA)-63) has a high mechanical strength of TS (35.4 MPa) and EB (17.9%).300 However, after the incorporation of quaterphenylene alkylene, the TS (37.3–42.5 MPa) of the copolymer increased with an increase in the amount of quaterphenylene, while the EB (11.7–10.4%) values decreased; this indicates that the membrane's stiffness and flexibility increased and decreased, respectively. It was conceivable to improve the stiffness of quaterphenylene despite the material's lack of flexibility because of the quaterphenylene's inherently inflexible character. Similarly, fluorobenzoyl group-grafted quaternized SEBS (e.g., HQA-F5-SEBS) had a higher MS compared to quaternized SEBS without a fluorobenzoyl group.139 Since fluorobenzoyl is hydrophobic, it does not absorb water and reduces swellings. To avoid swelling or shrinking and to make the membranes more rigid, the membranes must absorb relatively little water. In this way, the fluorobenzoyl group-incorporated SEBS (HQA-Fx-SEBS) showed low WU and SR, which led to high mechanical strength, compared to the fluorobenzoyl-free quaternized SEBS (HQA-SEBS), which showed greater WU and SR (Fig. 50b). Thus, AEMs-containing fluorine had a higher MS than those without fluorine substitution. Recently, it was discovered that the clustered cations-tethered poly(aryl piperidinium) (e.g., 3QPAP-0.3) had high mechanical strength and low flexibility (38.7 MPa and 20.4%).295 In contrast, the functionalized PAP with single piperidinium cation (1QPAP) had low mechanical strength and increased flexibility (31.8 MPa and 26.9%). However, the mechanical stability of poly(aryl piperidinium) decreased (38.7–28.0 MPa), and the flexibility increased (20.4–34.7%) with an increase in the weight ratio (0.3–0.5%) of PAP (Fig. 50c). Therefore, it was found in this study that the polymer backbone, rather than the side chain cations, is what determines a polymer's mechanical stability and flexibility.
Generally, the backbone's hydrophobic character and the cations' hydrophilic nature are responsible for the material's mechanical stability and flexibility. The mechanical strength with high and low degrees of electron beam-grafted ETFE AEM were recently investigated by subjecting it to varying amounts of electron beam irradiation in N2 atmospheres.299 When comparing the 40 kGy of electron irradiation in an N2 environment at RT (e.g., 40-N2-RT), it was found that the mechanical stability of the membrane with a high degree of grafting was not improved by electron irradiation at 100 kGy. The material's increased stiffness due to the higher degree of grafting made it very rigid. Therefore, the mechanical stability of the membranes greatly depends on the degree of grafting.
Increasing the molar ratio and degree of bromination increases the number of cation group sites, raising the IEC and WU (e.g., 40MePh-1.72).287 IEC and WU growth contributes to a less rigid membrane by reducing the mechanical stability. Therefore, increasing the WU and IEC did not aid in creating membranes with ideal flexibility and stiffness. It was found that medium IEC and WU in the membrane exhibited necessary mechanical stability (YM (480 MPa), TS (18 MPa), and EB (83%)). Compared to regular AEM made via cross-linking or grafting (SES-TMA and HTMA-DAPP), reinforced membranes often exhibited high strength (Fig. 50d). In this light, reinforced PTFE AEM (e.g., Durion) demonstrated excellent mechanical stability and lower flexibility than copolymer structure-based AEMs.329 Because of this, reinforcing increases the material's rigidity, which might be advantageous for long-term energy cell-oriented applications.
The membrane's mechanical stability (MS) rises as the number of cross-linked bication structures (e.g., m-XPTPA40-2N) rises.279 The cross-linking method supports the tangles of macromolecular chains in the membrane to improve the MS. On the other hand, the high flexibility of the membrane is due to its high molecular weight and moderate water absorption. However, this study showed that the increase in MS caused the crosslinked bication-functionalized membrane to become less flexible. The cross-linked AEM is less flexible because its extensions and chains cannot move as freely as they used to. Thus, crosslinking molecules with a bication structure also dramatically affects the materials' stiffness. Compared to fluorinated PAEK with merely the carbazole pendants connected, the stiffness and flexibility of the membrane improved when a long alkyl chain-based carbazole pendant cation was introduced into the PAEK backbone (e.g., PAEKHQACz-0.7).110 Because of the entanglement created by the long alkyl chain, the polymer's flexibility is increased (increase the elongation break). In contrast, membranes devoid of lengthy alkyl chains in their cation tethers were brittle and less flexible (decreasing the elongation break). As the amount of long alkyl chain cations in the membrane increases, its stiffness decreases, and the flexibility rises. Therefore, long alkyl chain-based cations in the polymer membrane may be appropriate for increasing the membrane flexibility in energy cells in the long term.
The membrane's mechanical stability is typically strong in its hybrid state. Compared to bare membrane PBI–DIm, PBI–DIm–Si hybrid AEM demonstrated superior mechanical strength (MS) and flexibility.305 In another investigation, the membrane's MS and flexibility were high and low in the form of a hybrid block co-polymer structure (e.g., DMBA-PIM-b-PSF-0.91).307 In contrast, the same membrane without the block co-polymer structure demonstrated low MS and high flexibility. The block copolymer structure with fluorinated hydrophobic polymer might improve the membrane's mechanical strength, whereas the hydrophilic polymer in the membrane without the block copolymer structure demonstrated excellent flexibility. Thus, fluorinated hydrophobic polymers are essential for enhancing the mechanical stability, whereas the hydrophilic polymer structures in the membrane are necessary for improving the flexibility. Cha et al. produced a copolymer-based AEM (e.g., QPPAE-4/1), and they evaluated it for MS after altering the monomer-to-trimer ratio (2:
1 and 4
:
1).306 It was discovered that a more extensive monomer content in the membrane resulted in higher mechanical strength and less flexibility than a smaller monomer quantity employed to construct a quaternized copolymer-based membrane (Fig. 50e). The membrane's mechanical strength increases with more monomer content, and the membrane remains remarkably flexible with less monomer content.
High molecular weight polyfluorene-based AEMs in the form of OH− and Cl− were recently investigated for MS.100 Polyfluorene-based membranes often have stiff backbones. However, the mechanical strength of membranes containing OH− and Cl− might vary. It was determined that the Cl− form of polyfluorene AEM (PFOTFPh-C6-TMA (Cl)) had more mechanical strength than the OH− form of the membrane. The explanation might be that the OH− form of polyfluorene can absorb more water, decreasing the mechanical stability. In comparison, the Cl− form of polyfluorene can resist absorbing more water, resulting in high mechanical strength. The multi-cation-functionalized fluorine-based AEM (TQ-PDBA-70%) has remarkable mechanical stability with a TS and EB of 33.4 MPa and 14.4%, respectively.283 Generally, the mechanical strength and flexibility of fluorine-based AEMs are high and stiff, respectively (Fig. 50f).
Recent research has shown that fluorine-containing block copolymers with a bend-twisted structure (NCBP-Im) exhibit higher flexibility (24.7%) and lower stiffness (28.5 MPa).285 As a result, the membrane structure of the bending and twisting type could have a high degree of flexibility. Similarly, another research found that a claw-type head with a partially fluorinated (PAEFPAE-3B-1.0-PD) membrane had a high rigidity (27.9 MPa) but low flexibility (9.6%).290 Thus, mechanical stability is determined by the structure of fluorinated membranes. In another study, according to Wang et al., the cross-linking process, also known as quaternization, increased the MS of the fluorinated membrane while simultaneously lowering its flexibility.278 Furthermore, enhancing the fluorinated polymer ratio in the membrane led to reduced MS and boosted flexibility. As a result, an overabundance of the fluorinated membrane may lower the mechanical strength of the membrane. Thus, the cross-linking (quaternization) technique does not affect the MS. Therefore, they observed that the fluorinated poly(aryl ether oxadiazole) with a TMHDA cross-linked structure (C-FPAEO-50-MIM) exhibited improved stiffness properties (28.0 MPa) and low flexibility (13.1%), whereas the same membranes lacking cross-linking structure revealed reduced stiffness (17–26 MPa) and increased flexibility (17–22%).
By increasing the quantity of bis-piperidinium-based cationic side chain in the poly(bromoalkyl oxindolebiphenylylene) (PISBr–NON-1), Zhang et al. found that the membrane stiffness (13.0–7.7 MPa) decreased and flexibility (15.7–49.2%) improved.227 As a result, cation side chains based on bis-piperidinium may enhance membrane flexibility owing to the hydrophilic feature of these chains. Similarly, increasing the proportion of piperidinium in the polymer PBMB (PDMB-Pi-0.5) reduces the membrane's stiffness and increased flexibility.144 Even though the backbone of the membranes is made of triphenylmethane, which is stiff and has high hydrophobic properties, the membranes are not very rigid. However, the high hydrophilic property of piperdinium contributes to the increased flexibility of the membrane.
The tensile strength (TS) is reduced and flexibility (elongation at break (EB)) is enhanced on increasing the crosslinking molarity in the AEM. Accordingly, with an increase in the molarity (0.5–1.0 M) of the crosslinker ether group containing DACEE in the polysulfone AEM, the TS (17.7–12.9 MPa) and EB (14.5–23.8%) were lowered and raised, respectively (e.g., PSf86-DACEE0.5).228 It may be justified by increasing the number of crosslinkers-containing ether groups, which may raise the membrane's hydrophilicity and lower its rigidity. However, compared to the ether group-containing DACEE, crosslinker DABDH had polysulfone (e.g., PSf86-DADBH0.75) and exhibited moderate TS (15.8 MPa) and low EB (17.4%). Therefore, the ether group-containing crosslinker may also contribute significantly to the mechanical stability of membranes. In another research, several crosslinkers (TMEDA, TMBDA, and TMHDA) with increased crosslinking (7.31–65.49%, 9.3–80.2%, and 8.75–90.03%) in the SEBS membrane were examined for mechanical stability (MS).203 TMHDA-crosslinked SEBS (e.g., C6-CQASEBS) had better TS (11.9–18.4 MPa) and EB (231.4–101.8%) than TMEDA (TS (8.2–13.7 MPa) and EB (191.2–88.7 MPa)) and TMBDA (TS (6.1–15.5 MPa) and (137–49%)) crosslinkers. Additionally, increasing the cross-linking degree might raise the TS while decreasing the EB of all membranes (flexibility). The TS was raised because the selected crosslinkers (TMHDA) may have a stiffer structure than other crosslinkers.
The crosslinker (DABDA) and increasing its grafting ratio (40–60%) in PPO (PPO/DABDA) caused a reduction in the tensile strength (38.1–14.2 MPa) and Young's modulus (1170–350 MPa) but caused an increase in the elongation at break (2.4–4.0%).230 The crosslinker DABDA could have hydrophilic qualities, which would boost the elastic properties of the PPO/DABDA-based AEM while decreasing its stiff properties. Therefore, as compared to higher crosslinking degrees (50–60%) of PPO, lower grafting degree (40%) of DABDA with PPO (PPO/DABDA-40) demonstrated better TS (38.1 MPa), higher Young's modulus (1170.0 MPa), and reduced elongation at break (2.4%). Researchers Qaisrani et al. found that the degree of chloromethylation in the polysulfone-Cl-DABDA (PSF-Cl-DABDA) crosslinked structure increased from 58 to 84%, and the tensile strength decreased from 25.0 to 17.5 MPa. In comparison, the elongation at break increased from 17.4 to 21.5%.220 These results were found when the chloromethylation degree was increased. Increasing the chloromethylation level lowers the tensile strength and can raise the membrane's flexibility. Elevating the chloromethylation level can raise the membrane's hydrophilic characteristics.
Consequently, selecting an appropriate crosslinker ensures the membrane's continued mechanical integrity. The lengthening of the side chains in AEMs affects the membrane's mechanical stability. Accordingly, increasing the side chain length in the quaternized polysulfone (e.g., acS10QAPSF) increased the TS, decreased the EB, increased the membrane's rigidity, and reduced the flexibility.201 Additionally, compared to sidechain-functionalized quaternized polysulfone, AEM with no side chains had extremely low TS and EB. Thus, the sidechain also affects the membrane's mechanical stability, and adding more sidechains will only improve the membrane's tensile strength. On increasing the imidazolium concentration in the PEEK (DI-PEEK) AEM, it was found that the tensile strength gradually decreased.219 It was discovered due to increased membrane plasticity caused by water absorption. Therefore, increasing the amount of imidazolium did not significantly benefit the membrane's mechanical stability.
Adjusting the monomer ratio in the synthesized PES (PES-BTP) AEM from 20 to 30% was examined for its mechanical stability, and the results revealed that increasing the monomer content in the membrane tends to lower the tensile strength while simultaneously increasing the elongation at break.184 The monomer is to blame for the membrane's increased flexibility rather than its rigidity. It might be because the monomer makes the membrane more hydrophilic, allowing it to take in more water. The ratio that is the lowest possible functionalization of 20% of the monomer PES (PES-BTP-20%) was shown to have greater flexibility (46.8%). Still, lower tensile strength (21.7 MPa), while increasing the monomer ratio, decreased the tensile strength and increased the flexibility. As a result, the monomer plays a significant part in improving the membrane's flexibility. The dried QPPO membranes and their recycled structure dried disulfide bond cross-linked RC-QPPO AEMs were evaluated for mechanical stability, and the results indicated that the recycled membrane revealed stronger tensile strength and higher flexibility than the bare dried QPPO membrane.196 For technical understanding, the mechanical properties of various AEMs have been represented in Table 7.
AEM | YM (MPa) | TS (MPa) | EB (%) | MS | Ref. |
---|---|---|---|---|---|
a AEM = anion exchange membrane; YM = Young's modulus; TS = tensile strength; EB = elongation at break; MS = mechanical stability. | |||||
PQP-100 | 2134.0 | 83.9 | 15.8 | High | 235 |
NPBI-QA55-B45 | — | 34.1 | 6.2 | Medium | 188 |
DQ-PPO-17-OH | — | 9.1 | 23.8 | Low | 231 |
PPO-iP-44 | 1599.2 | 34.9 | 2.4 | High | 232 |
SDQEO | — | 10.5 | 27.3 | Low | 113 |
PTP-90 | 466.7 | 29.2 | 6.8 | Medium | 234 |
PBPCL1.29 | — | 38.0 | 14.0 | Medium | 288 |
PS/pTAP (40![]() ![]() |
— | 12.5 | 2.5 | Low | 237 |
SEBS C1-x and C6-x | — | 53.6 | 500.0 | Medium | 233 |
QP(VBC-St)-MB/EA0.3 | 773.0 | 10.3 | 4.7 | Medium | 261 |
QP(VBC-St)-4-TBP0.7 | — | 6.2 | 14.9 | Low | 262 |
C-PVAf-ABPBI | 103.0 | 5.0 | 6.0 | Low | 131 |
VImPPO-x:PVA-x | — | 20.9–30.8 | 20.5–60.2 | Medium | 128 |
PBI-27%BPTMA | 162 | 49.0 | 32.0 | Medium | 187 |
10%-PBP-c-SEBS | — | 20.0 | 46.1 | Low | 265 |
20PDM-2%PEG2000 | 808.4 | 21.2 | 15.6 | High | 255 |
PVA-0.6PQVBC40% | — | 14.6 | — | Low | 257 |
BPPO-PBBP | — | 18.2 | 10.3 | Low | 301 |
P(4PA-co-2PA)-63 | — | 42.50 | 10.4 | Medium | 300 |
HQA-F5-SEBS | — | 24.8 | — | Low | 139 |
3QPAP-0.3 | — | 38.7 | 20.4 | Medium | 295 |
40-N2-RT | 298.0 | 23.8 | 262.0 | Medium | 299 |
40MePh-1.72 | 480.0 | 18.0 | 83.0 | Medium | 287 |
Durion | 722.0 | 28.1 | 7.2 | High | 329 |
m-XPTPA40-2N | 448.0 | 24.4 | 16.2 | Medium | 279 |
PAEKHQACz-0.7 | — | 32.5 | 43.5 | Medium | 110 |
PBI–DIm–Si hybrid | 640.0 | 32.9 | 16.9 | High | 305 |
DMBA-PIM-b-PSF-0.91 | 785.7 | 22.7 | 6.8 | High | 307 |
QPPAE-4/1 | 1669.0 | 24.4 | 2.0 | High | 306 |
PFOTFPh-C6-TMA(Cl) | — | 42.3 | — | Medium | 100 |
TQ-PDBA-70% | — | 33.42 | 14.43 | Medium | 283 |
NCBP-Im | — | 28.5 | 24.7 | Medium | 285 |
FPAE-3B-1.0-PD | — | 27.9 | 9.6 | Low | 290 |
C-FPAEO-50-MIM | 120.0 | 28.0 | 13.1 | Medium | 278 |
PISBr-NON-1 | — | 13.0 | 15.7 | Low | 227 |
PDMB-Pi-0.5 | — | 30.4 | 9.5 | Medium | 144 |
PSf86-DACEE0.5 | — | 17.7 | 14.5 | Low | 228 |
C6-CQASEBS | — | 18.4 | 101.8 | Medium | 203 |
PPO/DABDA-40 | 1170.0 | 38.1 | 2.4 | High | 230 |
PSF-Cl-DABDA-58 | — | 25.0 | 17.2 | Low | 220 |
acS10QAPSF | — | 25.1 | 18.9 | Low | 201 |
DI-PEEK | — | 47.0 | — | Medium | 219 |
PES-BTP-20% | — | 21.7 | 46.8 | Low | 184 |
RC-QPPOdry | — | 19.49 | 6.73 | Medium | 196 |
The water uptake of AEM is essential if one wishes to achieve excellent performance in water electrolysis and long-term durability in this activity.334 This is because high water uptake led to high conductivity, which enabled one to achieve high performance in water electrolysis. An advantage of AEM-based water electrolysis over traditional electrolysis is that either water or a low-concentration alkaline medium could be employed as the electrolyte.32 The cell voltage and current density are two of the most important operational parameters in electrolyzing water. In most cases, the amount of energy a cell needs to function properly is determined by its voltage. On the other hand, the creation of H2 is directly related to the current density. Accordingly, a high current density results from a fast electrochemical reaction occurring in the cell, and an AEM water electrolyzer may be operated in the current density range between 100 and 500 mA cm−2.32 However, one of the difficulties associated with the cell is that when it is working at a greater current density, bubbles are formed on the surface of the electrode. To this degree, this results in an increase in the amount of overpotential that is required. As a result, there is a need for more studies to be conducted to undertake an analysis of the optimal range of current density for the development of hydrogen. While the AEM's potential in fuel cell applications has been extensively studied, its use in water electrolysis has received comparatively less attention. Therefore, future studies should focus more on synthesizing efficient AEM for the benefit of water electrolysis for hydrogen generation. Table 8 presents the results of recent studies on using AEMs in water electrolysis for hydrogen production, and the following objectives were discussed: cell voltages employed, attained current density, temperatures, electrolytes, and durability.
AEMs | Thickness (μm) | Voltage (V) | Catalysts | Electrolyte | CD (A cm−2) | T (°C) | Durability | Ref. |
---|---|---|---|---|---|---|---|---|
a AEMs = anion exchange membranes; (A) = anode, (C) = cathode, CD = current density; T = temperature. | ||||||||
QAPPT/3% Ni–Fe LDH | — | 2.0 | IrO2 (A) and Pt/C (C) | 1 M KOH | 1.0 | 60.0 | High | 335 |
PQP-100 | 10 | 2.0 | IrO2 (A) and Pt/C (C) | 1 M KOH | 1.5 | 85.0 | High | 235 |
BD3/50EVOH | 55 ± 1.2 | 2.0 | IrO, Co (A) PtRu/C (C) | 1 M KOH | 1.57 | 70.0 | High | 336 |
TM1-MEA | — | 1.9 | IrO2 (A) and Pt/C (C) | 1 M KOH | 2.75 | 70.0 | Medium | 325 |
PBP_15 | 15 | 2.2 | Ni foam (A) and Pt/C (C) | 1 M KOH | 2.8 | 60.0 | — | 337 |
QBNTP-MP11 | 15 | 2.0 | IrO2 (A) and Pt/C (C) | 1 M KOH | 2.7 | 80.0 | High | 338 |
C-PVAf ABPBI | 30–45 | 2.0 | MEA not fabricated | KOH | 0.27 | 70.0 | Medium | 131 |
PTP-90 | 45 | 2.2 | IrO2 (A) and Pt/C (C) | 1 M NaOH | 0.91 | 55.0 | Low | 234 |
PiperION | 50 | 1.9 | IrO2 (A) and Pt (C) | Pure water | 1.0 | 55.0 | High | 122 |
HWU-AEM | 20 | 1.9 | PtRu/C (A) and Pt/C (C) | 1 M KOH | 0.5 | 60.0 | High | 339 |
HTMA-DAPP | 50 | 2.0 | Co3O4 (A) and Pt/C (C) | 1 wt% K2CO3 | 1.0 | 50.0 | High | 329 |
AF2-HWP8-75-X | 85 | 2.0 | IrOx (A) and Pt/C (C) | 1 M KOH | 0.2 | 70.0 | Very high | 340 |
PBI/mTPN-50.120 | 10 | 2.0 | Ni–Fe (A) and Mo (C) | 1 M KOH | 0.25 | 50.0 | High | 272 |
PcPBI-Nb-C2 | 20 ± 5 | 2.1 | IrO2 (A) and Pt/C (C) | 1 M KOH | 0.36 | 60.0 | Medium | 341 |
cPBI-0.4p-0.6s | 17 | 2.1 | IrO2 (A) and Pt/C (C) | 1 M KOH | 0.55 | 80.0 | Medium | 342 |
V-1.5-O-1 | 40–50 | 2.0 | NiFe2O4 (A) and NiFeCo (C) | 0.1/1 M KOH | 0.86 | 60.0 | High | 343 |
g-VBC-5-co-Sty-16-Q | 70–100 | 2.0 | Pt/C (C and A) | 1 wt% KOH | 0.8 | 55.0 | Low | 344 |
PAEK-APMBI | — | 1.9 | Nickel foam (A & C) | 10 wt% KOH | 0.5 | 60.0 | — | 147 |
SEBS-P206 | — | 2.0 | Ir/black (A) and Pt/C (C) | 0.1 M KOH | 0.68 | 60.0 | Medium | 345 |
SEBS-Py206 | 70 | 2.0 | OXYGEN-N (A) and H2GEN-M (C) | 0.1 M KOH | 0.52 | 30.0 | — | 346 |
TriPPO-50SEBS | 40–50 | 1.8 | IrO2 (A) and Pt/C (C) | 1 M KOH | 0.71 | 70.0 | High | 347 |
Fumasep FAA-3-50 (Fumatech, Germnay) | 45–50 | 2.15 | IrO2 (A) and Pt/C (C) | 1 M KOH | 1.0 | 60.0 | Low | 253 |
HTMA-DAPP | 50 | 2.0 | IrO2 (A) and PtRu/C (C) | 1% KOH | 1.0 | 60.0 | — | 348 |
Diamine-PSF/PVBC | 80 | 1.85 | NiFe (A) and NiMo (C) | 1 M KOH | 0.5 | 80.0 | High | 135 |
PSF-TMA | — | 1.8 | Pb ruthenate pyrochlore (A) and Pt (C) | Water | 0.4 | 50.0 | Medium | 349 |
qPSU/NIM-N+ | 45 ± 5 | 2.2 | Ni (A) and Pt/C (C) | 1 M KOH | 3.7 | 80.0 | High | 350 |
qPSU | 48 ± 5 | 2.2 | Ni (A) and Pt/C (C) | 1 M KOH | 4.2 | 90.0 | — | 351 |
PISPVA46 | 55 | 2.0 | IrO2 (A) and Pt/C (C) | 0.5 M KOH | 0.54 | 60.0 | Medium | 101 |
LSCPi-AEM | 50 | 1.8 | IrO2 (A) and Pt/C (C) | Water | 0.3 | 50.0 | Medium | 352 |
A-201 (Tokuyama, Japan) | — | 1.95 | Ni/(CeO2–La2O3)/C (A and C) | 1% K2CO3 | 0.5 | 60.0 | High | 353 |
Mg–Al LDH | 900 | 2.2 | CuCoO (A) and Ni/(CeO2–La2O3)/C (C) | 0.1 M NaOH | 0.2 | 70.0 | High | 354 |
PAES-TMI-0.25 | 45 | 2.0 | IrO2 (A) and Pt/C (C) | 2 M NaOH | 1.2 | 80.0 | Medium | 355 |
PAES-MQA-0.18 | 44 ± 2 | 2.0 | IrO2 (A) and Pt/C (C) | 2 M NaOH | 1.12 | RT | High | 356 |
PK_0.5 Step_3 | 70–80 | 2.0 | Nickel foam (A and C) | Water | 0.01 | 80.0 | Low | 357 |
TMA-53-MEA | — | 2.0 | IrO2 (or) NiFe (A) and Pt–Ru (or) Pt (or) NiMo/C (or) PtRu/C (C) | Water | 1.64 | 60.0 | Medium | 360 |
Sustainion® (dioxide materials, USA) | — | 2.1 | NiFe2O4 (A) and NiFeCo (C) | 0.1–1 M KOH | 0.5 | 60.0 | — | 359 |
PBPA | 50 | 2.0 | IrO2 (A) and Pt/C (C) | 1 M KOH | 4.0 | 80.0 | High | 361 |
TDMAP-50%-SEBS | 35–40 | 2.0 | IrO2 (A) and Pt/C (C) | 1 M KOH | 1.19 | 70.0 | — | 362 |
QPDTB | — | 1.9 | Cu0.7Co2.3O4 (A) and Ni (C) | Water | 0.1 | 30.0 | High | 363 |
AQ720A4-Q | — | 2.2 | NiFeOx (A) and Pt/C (C) | 1 M KOH/Water | 0.90 | 90.0 | — | 364 |
PFT-C10-TMA (MEA-1) | 56 | 1.8 | IrO2 (A) and PtRu/C (C) | 1 M KOH | 1.0 | 80.0 | High | 366 |
PFB-QA | 40–45 | 2.0 | IrO2 (A) and Pt/C (C) | 1 M KOH | 1.53 | 70.0 | High | 367 |
![]() | ||
Fig. 51 Polarization curves for the AEM water electrolysis (a) poly(aryl piperidinium)-based AEMs prepared with extended aromatic groups and durability test of the AEM, PQP-100 (1 M NaOH at 60 °C under 200 mA cm−2), reproduced with permission from ref. 235, Copyright 2022, Elsevier. (b) C-PVAf ABPBI AEMs and its durability test (15 wt% KOH at different temperatures under 200 mA cm−2), reproduced with permission from ref. 131, Copyright 2021, John Wiley & Sons Ltd. (c) Piperidinium-tethered aryl ether-free-based polymers (PTP-x) and durability of WE test with PTP-85 AEM (1 M NaOH at 55 °C under 400 mA cm−2), reproduced with permission from ref. 234, Copyright 2021, Elsevier. |
On the polydiallylammonium-incorporated cationic network sIPN AEM (e.g., BD3/50EVOH), a similar current density (1.57 A cm−2) observation may be seen.336 It was possible to achieve a higher current density because of the membrane's strong conductivity (161 mS cm−1), good mechanical stability, and elevated alkaline stability. However, increasing the quantity of poly(vinyl alcohol-co-ethylene) (EVOH) in the sIPN membranes causes a reduction in the current density since the sIPN membranes' ionic conductivity is also reduced (161–62.8 mS cm−1). The ionic conductivity of the membrane will undoubtedly have an impact on this investigation of current density. Additionally, the high alkaline stability of the membrane may have contributed to the great durability of the sIPN AEM, which has been shown to exhibit no significant voltage fluctuations up to 300 min at 1.6 V at 70 °C.
Interestingly, water electrolysis was performed by commercial AEMs TM1-MEA and FAA-MEA under various mixtures of solutions such as KOH/KOH, KOH/water, water/KOH, and water/water.325 To this degree, TM1-MEA showed high current density under all conditions of solutions compared to FAA-MEA due to high conductivity. Additionally, TM1-MEA showed maximum current density (2.75 A cm−2) in KOH/KOH than other conditions at 1.9 V. Besides, TM1-MEA-conducted electrolysis study in water/water conditions exhibited very low performance, and the obtained current density was 0.175 A cm−2. Therefore, higher current density was obtained by TM1-MEA under KOH/KOH conditions compared to FAA-MEA. It could be the polymer structure of TM1 AEM and suggested possible conditions for water electrolysis with at least a mixture of KOH rather than only water. This membrane's durability was evaluated for 50 h, and it was discovered that after 200 cycles, the current density of the TM1-MEA was only reduced by 32%, compared to 85% for the FAA-MEA, indicating that the TM1-MEA had greater stability. As the cell's temperature rose from 50 to 80 °C, so did the breakdown rate of the TM1-MEA membrane (from 40 to 321 mV h−1). Therefore, raising the temperature of the cell affects its membrane.
In another investigation, the performance of poly(biphenyl piperidinium)-based AEMs (e.g., PBP) at varying thicknesses (15–60 μm) was evaluated.337 It was found that the thickness of the AEMs had a significant impact on the current density. Accordingly, the current density (2.8–1.54 A cm−2) decreased as the thickness (15–60 μm) of AEMs rose, and this might be because lower and higher thickness membranes had higher and lower series resistance, respectively. High-resistance membranes may have poor conductivity and low current density, while low-resistance membranes have high conductivity and high current density. As a result, PBP (PBP-15) with a 15 m thickness showed excellent performance at 2.8 A cm−2 at 2.2 V in 1 M KOH electrolyte at 60 °C. The higher current density by the membrane was likely caused by its higher conductivity (185 mS cm−1) and thinner thickness (15 μm). When dual-methylpiperidinium AEMs (e.g., QBNTP-MP11) were used to crosslink poly(binaphthyl-co-terphenyl piperidinium) at electrolyzer settings of 2.0 V, 1 M KOH, and 80 °C, significant current density (2.75 A cm−2) was found on the polymer.338 Due to the cross-linking of dual-methylpiperidinium, the high current density of the membrane may cause the membrane's high conductivity (181.2 mS cm−1). Conversely, cross-linker-free AEM (QBNTP) had poorer current density (1.9 A cm−2) at the same voltage, which may have been caused by lower ionic conductivity (140 mS cm−1). Compared to QBNTP at 1 M KOH, 80 °C, and an applied current density of 0.2 A cm−2, the QBNTP-MP11's durability was superior for up to 100 h. The QBNTP-MP11's exceptional durability was made possible by its strong ionic transport, high mechanical strength, and great chemical stability. On the other hand, QBNTP showed poor performance after 40 h, which might result from membrane breakdown because of their poor mechanical and alkaline stability.
The current density of the membrane (e.g., C-PVAf-ABPBI) was measured at cell voltage 2.0 V and different temperatures, such as 50 and 70 °C, in the presence of 15 wt% KOH.131 Because of the strong ion mobility at high temperatures, the membrane's current density rose from 0.23 to 0.27 A cm−2 as the temperature increased from 50 to 70 °C (Fig. 51b). As a result, the temperature affects the current density in the membranes for water electrolysis. Despite this, it has been shown that the current density of the C-PVAf-ABPBI is rather low. The AEM durability test was completed in 60 min, revealing that the starting cell voltage (1.98 V) did not vary significantly. Conducting a durability test on the membrane for at least 50 h would be preferable. As a result of ionic conductivity, the AEMs' current density may vary. By analyzing PTP-based AEMs, Hu et al. verified that PTP-75 achieved 0.767 A cm−2 at 2.2 V. In contrast, a higher current density of 0.910 A cm−2 was detected in this report for alkaline water electrolysis.234 When the temperature is raised from 55 °C to 75 °C, the current density of the AEMs rises from 0.910 A cm−2 to 1.0 A cm−2; this rise in current density could be caused by the AEMs' enhanced conductivity (Fig. 51c). The authors carried out a durability study for the PTP-85 AEM. It was assessed at a current density of 0.4 A cm−2 at 55 °C. Cell voltages are seen to rise (to a maximum of 2.8 V) with time (to a maximum of 120 h), which may result from a slight deterioration of the membrane electrode assembly.
The membrane with a lower hydroxide conductivity requires a higher voltage to work well in water electrolysis, whereas the membrane with a greater conductivity performs well at low voltages. For water electrolysis, commercially available AEMs such as Sustainion, Aemion, and PiperION were tested, and the findings showed that they produced 2.25, 2.24, and 2.05 V, respectively, at current densities of 1.0 A cm−2 in water. Because of their low hydroxide conductivities, Sustainion and Aemion could operate at higher voltages; however, piperION, which had a higher conductivity, could function at a lower voltage. PiperION showed a voltage deterioration rate (VDR) of 0.67 mV h−1 at a current density of 0.5 A cm−2 at 55 °C for 175 h. In contrast, other commercial membranes failed to withstand and shut down above 2.9 V during the same test.122 Since the hydroxide conductivity improves with temperature, the operating voltages of the PiperION show a drop as the temperature rises, resulting in improved performance. Remarkably, the PiperION has a durability of 175 h when compared to membranes such as Sustainion and Aemion. Curiously, at 1.9 V, high water uptake-AEM (HWU-AEM) has good performance under pure water input to obtain a current density of 0.254 A cm−2 at 60 °C in contrast to utilizing 1 M KOH, which exhibits a current density of 0.911 A cm−2 at 30 °C.339 On the other hand, it was discovered that the HWU-AEM membrane has very high durability, demonstrating a modest increase in cell voltage from 1.97 to 2.0 V after being tested for 150 h. In another study, under the circumstances of 1 wt% K2CO3, AEM–water electrolysis with a poly(phenylene)-based membrane and a binder poly(fluorene) ionomeric gave the best result, which was 1 A cm−2 at 2 V.329 For the durability test, at a current density of 750 mA cm−2, it was discovered that the VDR was 0.27–0.55 mV h−1 after being measured for more than 400 h.
Aemion+® is a poly(bis)imidazolium ionenes-based commercial membrane with strong ionic conductivity, high mechanical strength, and great chemical stability. As a result, polyolefin-reinforced Aemion+® (e.g., AF2-HWP8-75-X) AEM is utilized for water electrolyzer testing.340 Although this reinforced membrane has a high ionic conductivity (112 mS cm−1), it exhibits low current density (0.2 A cm−2) at the applied voltage of 2.0 V when tested with 1 M KOH at 70 °C. Therefore, the membrane's current density could not rely on ionic conductivity but rather on the structure and backbone of the polymer. However, this reinforced membrane may demonstrate high durability up to 5000 h and twelve months with virtually unaltered conductivity operation. Although this membrane could have a low current density, its continued usage might be useful in water electrolyzers for hydrogen production. In an electrolysis system, the efficiency of the cells may depend critically on the concentration of KOH.
Recently, MEA (catalysts (Ni/Fe and Mo) and binder (SEBS)) with PBI/mTPN-50.120 AEM demonstrates that the current density (from 0.01–0.172 A cm−2) increases when the KOH concentration rises from 0.01 to 1 M.272 This membrane was put through its durability test at a current density of 0.25 A cm−2. Almost all cell voltages (1.98 V) remained stable for over 200 h. The high chemical stability membrane of crosslinked piperidinium-functionalized norbornene-introduced alicyclic PBI membrane (e.g., PcPBI-Nb-C2) may not perform well since the current density of the membrane was determined to be 0.368 A cm−2 at 2.1 V.341 The membrane could not achieve a significant current density despite having a high OH− conductivity (105.5 mS cm−1). The designed membrane's durability was also tested for a short time (8.3 h) at 60 °C with a current density of 0.15 A cm−2; however, it passed with excellent performance since the voltage was maintained between 1.49 and 1.51 V. N-Spirocyclic-quaternized copolymer PBI AEMs with flexible alicyclic and stiff benzene incorporation (e.g., cPBI-0.4p-0.6s) show strong ionic conductivity, mechanical stability, and alkaline stability.342 But this type of structured membrane (long-side chain-based piperidinium and N-spirocyclic QAs) could only deliver modest performance in a water electrolyzer. For example, the current density by the cPBI-0.4p-0.6s-based AEMWE was found to be 0.553 A cm−2 at 2.1 V, 1 M KOH, and 80 °C. Due to the high conductivity (110.3 mS cm−1), the prepared membrane achieved a modest performance. Also, the membrane's durability was average, and there were only small changes in the voltage (1.78–1.85 V) after 10 h of testing. It might be better if the study lasted at least 100 h.
The crosslinkers and electrolyte content affect the membrane's current density as well. Accordingly, 1,6-hexanedithiol cross-linked polynorbornene-based AEMs (V-1.5-H-1) and 3,6-dioxa-1,8-octanedithiol cross-linked polynorbornene-based AEMs (V-1.5-O-1) were put to the test in a water electrolyzer with 1 and 0.1 M KOH and an applied voltage of 2.0 V at 60 °C.343 Better current density (0.86 A cm−2) was achieved with 1 M KOH and 3,6-dioxa-1,8-octanedithiol crosslinked AEMs system than 1 M KOH with 1,6-hexanedithiol-crosslinked AEMs system (0.513 A cm−2). However, the current density was reduced when using 0.1 M KOH for the membranes V-1.5-O-1 (0.60 A cm−2) and V-1.5-H-1 (0.21 A cm−2). Therefore, the concentration of electrolytes has a significant impact on the membrane current density. Although, at 40 °C, the V-1.5-H-1's ionic conductivity (112 mS cm−1) was slightly higher than that of V-1.5-O-1 (108 mS cm−1). However, a lower current density is seen in V-1.5-H-1 than V-1.5-O-1. The fact that V-1.5-O-1 had a superior current density might be attributed to the effects of the crosslinker (3,6-dioxa-1,8-octanedithiol) and the molarity of the electrolytes (1 M KOH). Additionally, the V-1.5-O-1's durability test was superior to the V-1.5-H-1's at 0.1 M KOH, 60 °C, and at a current density of 0.2 A cm−2, perhaps owing to the membrane's high alkaline stability. There were no appreciable changes in voltages over time up to 70 h. Vinylbenzyl chloride/styrene-tethered copolymer-based membranes were shown to have low durability and moderate current density (0.8 A cm−2) at 2.0 V in the 1 wt% KOH solution (e.g., g-VBC-5-co-Sty-16-Q).344 This structured copolymer membrane had poor ionic conductivity (5.7 mS cm−2) and maybe low alkaline stability, resulting in a moderate current density and short durability (stable for only 3 h). Therefore, an additional study on membrane modification is required to increase the conductivity and alkaline stability of the membranes in light of AEMWE performances.
Imidazolium-functionalized membranes are difficult to test in alkaline electrolysis because OH− attacks the C2 position and breaks the membrane. It is possible that adding C2-methyl to the membrane could make it more stable under alkaline conditions. The chemical stability of the AEM could also be improved by adding N3-butyl. Thus, C2-methyl and N3-butyl-incorporated imidazolium-functionalized PAEK membrane (e.g., PAEK-APMBI) was studied for water electrolysis performance at 2 A cm−2 in 10 wt% KOH solution at 60 °C, and they were compared with commercial AEM (FAA-3).147 The high conductivity of imidazolium-functionalized PAEK obtained better results (2.53 V at 2.0 A cm−2). It had better electrochemical performance than the FAA-3 membrane (2.63 V at 2.0 A cm−2) as it was made. The benefit of the C2-methyl substitute in the membrane is that it improves the alkaline stability and keeps the imidazolium from being attacked by hydroxide.
The longer side chain-crosslinked piperidinium-functionalized membrane (e.g., SEBS-P206) was put through tests for water electrolysis using clean water and 0.1 M KOH, and the results showed that its current density was 0.275 A cm−2 and 0.680 A cm−2, respectively.345 Thus, the KOH solution that uses the AEMWE system can potentially attain higher results. Similarly, longer side chain-crosslinked pyrrolidinium-functionalized membranes, such as SEBS-Py206, demonstrate greater current densities (0.5 A cm−2), which were achieved in the presence of 0.1 M KOH, compared to lower current densities (0.17 A cm−2) when using pure water and a cell voltage of 2 V.346 Nevertheless, the current density of the SEBS-Py206 was lower than SEBS-P206 AEMs in this study's testing of water electrolysis using a PGM-free catalyst or conductivity issues. Triazole may be utilized to crosslink the two polymers to enhance the AEMWE performances and mechanical qualities. Accordingly, PPO-SEBS with triazole crosslinking (such as TriPPO-50SEBS) was found to have a high current density (0.71 A cm−2) compared to PPO-SEBS without crosslinking (such as PPO-50SEBS), which had a low current density of 0.42 A cm−2.347 Ionic conductivities of high (77.3 mS cm−2) and low (66.0 mS cm−2) values may be caused by larger and lower current densities by the membranes. Triazole may, thus, increase the membrane conductivity and hence increase the current density. The crosslinked PPO-SEBS's durability was extremely high and lasted for 3.8 h as opposed to the triazole-free PPO-SEBS, which degraded after 0.06 h at a current density of 0.2 A cm−2.
High levels of current density (>0.5 cm−2) are often required for hydrogen production on an industrial scale when electrochemical techniques are used. Therefore, electrolysis was investigated by Wan et al. using a commercially available AEM (FAA-3-50) in conjunction with a manufactured MEA (Pt/C and IrO2 catalyst and Nafion ionomer).253 At 2.15 V, the membrane attained a current density of 1 A cm−2. Despite this, the commercial AEM's performance was inferior to that of the developed PTFE/LDH composite membrane (1 A cm−2 and 1.88 V). It was discovered that additional catalyst usage (LDH) with mass transfer and the improved OER (oxygen evolution reaction) performance of LDH were the reasons for the composite membrane's greater performance (Fig. 52a). The cause may also be that the interfacial resistance between the catalyst and FAA-3-50 is greater than that between the catalyst and PTFE/LDH. At a current density of 1 A cm−2, the researchers found that an increase in temperature from 20 to 60 °C resulted in a drop in the voltage from 2.1 to 1.88 V. Because an increase in temperature facilitates the passage of hydroxide ions, the electrolyzer's ohmic resistance will noticeably decrease as the temperature rises. This might be the cause of the falling voltage. In addition, the transfer of charge that occurs at the interface between the catalyst layer and the membrane in the MEA is caused by the mobility of OH− ions. Therefore, in the MEA with a composite membrane, the number of OH− ions that flow between the electrode and membrane increases, ultimately resulting in a higher degree of reaction but a lower cell voltage. When evaluated for durability, the two membranes, PTFE/LDH and FAA-3-50, showed strong and poor stability for more than 180 h at a current density of 0.5 A cm−2, respectively (Fig. 52b). Although the catalysts were washed out of the electrodes, the voltage was raised by 8 h for FAA-3-50, demonstrating that the catalyst and flat AEM interface did not retain the catalyst particles during continuous operation, which resulted in instability. At the same time, the three-dimensional structure of the PTFE/LDH composite was used to hold the catalyst particles that increase the stability during continuous operation.
![]() | ||
Fig. 52 (a) Schematic depiction of the alkaline AEMWE, the polarization arcs of PTFE, FAA-3-50, and PTFE/LDH-3 AEMs, and with PGM (IrO2//Pt/C) and PGM-free (LDH//CoP) catalysts and (b) durability test of the different AEMs (1 M KOH solution at 60 °C under 500 mA cm−2), reproduced with permission from ref. 253, Copyright 2021, Elsevier. |
According to recent research, utilizing the alkali metal media available in the electrolysis system may aid in designing and optimizing AEMWE.348 The polarization of AEMWE showed that an increase in the electrolyte (KOH) concentration resulted in an improvement in the operating performance of the electrolyzer while simultaneously leading to a reduction in the cell's high-frequency resistances. Research demonstrates that additional hydroxide ions play a crucial function, including acting in the AEM, the catalyst layer's ohmic resistance, and working in cell performance reaction rates. In addition, it was discovered that the cell, including the electrolyte 1 M KOH, performed better than the cell containing only water. Therefore, it was determined that an electrolyte solution with a high pH in AEMWEs is essential to accomplish a high level of cell electrolyze performance. It was discovered that ionomers with strong ion exchange capacity and a greater interface with electrocatalysts might increase the AEMWE performance. In addition, the significance of conductivity in the electrode and electrolyte supplied in the electrode for the performance of the AEMWE was shown by the high level of performance achieved when the 1 M KOH solution was circulated.
Mechanically reinforced and high ion conduction diamine cross-linked PSF and PVBC-MPy-based AEMs have recently been hooked up to improve water electrolysis performance.135 Cell performance was proven to be dependent on membrane conductivity. It was found that the AEM (e.g., M-6#) lasts in the cell for up to 48 h, after which the cell voltage stays the same (Fig. 53a). Better results for OER were achieved using polysulfone-based AEMs (e.g., PSF-TMA) in alkaline water electrolyzers with the inexpensive catalyst Pb-ruthenate-pyrochlore (Pb-R-P).349 This research aimed to quantify electrolyzers' stability and long-term performance and identify the root causes of performance degradation. Results from the study demonstrated that water electrolysis at 50 °C produced a current density measurement of 0.4 A cm−2 at 1.8 V. These composite-structured membranes may perform poorly for water electrolysis in hydrogen production due to their low current density.
![]() | ||
Fig. 53 (a) Durability profile of an AEMWE cell with diamine crosslinked AEM (e.g., M-6#) at 500 mA cm−2, used 1 M KOH solution at 80 °C, reproduced with permission from ref. 135, Copyright 2021, Elsevier. (b) Polarization curves for AEMWE-one cells with various PISPVA-x blended AEMs, reproduced with permission from ref. 101, Copyright 2020, Elsevier. (c) Polarization curves and durability test of AEMWE with membranes (SCPi and LSCPi) at 200 mA cm−2 in 1 M NaOH solution at 50 °C, reproduced from ref. 352, Copyright 2019, Royal Society of Chemistry. (d) Polarization curves for various electrolytes experiments used with A-201 membrane and the inset picture show durability test of the commercial AEMs, reproduced with permission from ref. 353, Copyright 2017, Elsevier. (e) AEMWE polarization curves for various commercial membranes performed at different temperatures (40 to 60 °C) in 1 M KOH, reproduced with permission from ref. 359, Copyright 2020, Elsevier. |
The significant current density (3.7 A cm−2) for composite-structured quaternized polysulfone-TMA-functionalized silica AEMs (e.g., qPSU/NIM-N+) at 2.2 V under 1 M KOH at 80 °C.350 This composite membrane may be particularly useful for producing hydrogen in real-world applications because of its high current density and high membrane conductivity (109 mS cm−1). Long-term behavior is crucial for AEMs; hence, consideration should be given to their resilience over time when developing AEM electrolysis. As a result, the test was run at 2 V for 15 h while submerging in a 0.5 M KOH solution, and it was found that the AEM remained stable for up to 14 h. Interestingly, the straightforward organic method to create quaternized polysulfone (qPSU) AEM demonstrated a much greater current density (4.2 A cm−2) at 2.2 V under 1 M KOH at 90 °C.351 It was shown that the current density might not depend on ionic conductivity, although the qPSU's ionic conductivity (77 mS cm−1) feature allowed for achieving a very high current density. However, the membrane's increased current density at higher temperatures is the sole issue with the reported study.
Recently, higher amounts of imidazolium-functionalized blended AEMs (e.g., PISPVA46) with high conductivity and high hydration characteristics have shown superior electrochemical performance than smaller amounts of imidazolium-functionalized blended AEMs such as PISPVA37 and PISPVA 28.101 Since PISPVA46's charge transfer, ohmic, and mass transport resistance was lower than those of PISPVA37 and PISPVA28, thus, PISPVA46 had good electrochemical performance. The PISPVA46 membrane's stability was discovered to be somewhat less stable, dropping from 0.35 A cm−2 to 0.22 A cm−2 from 30 min to 80 h, respectively (Fig. 53b). Long side chain (LSC) and short side chain (SC) QAs, connected to AEM-containing piperidinium (Pi) groups, respectively, demonstrated electrolysis of pure water at 50 °C.352 The research found that LSCPi and SCPi AEMs had a current density of 0.3 and 0.2 A cm−2 at 1.8 V, respectively. Comparing LSCPi AEM to SCPi AEM, the improved performance of the potential cell with LSCPi AEM was attributed to their lower ohmic resistance or greater membrane conductivity (Fig. 53c). On the LSCPi AEMs, the possibility of delamination between the membrane and catalyst was attributed to the degradation of MEAs, resulting in moderate durability. It was found that there was a rapid increase in cell voltage after 35 h.
Another research examined the effects of changing the catalyst loading (10–40 mg cm−2), ionomer quantity (9–33%), electrolytes (K2CO3, KHCO3, KOH, and water), and temperature (40–80 °C) on AEMWE performance.353 Catalyst loading between 30 and 40 mg cm−2 was optimal, with 10 mg cm−2 being the bare minimum due to the increased overpotential. It refers to the extra energy required to start the water-splitting process above and above what is anticipated on the principle of thermodynamics. It was shown that just 9% of the ionomer was enough to prevent cracking in the catalyst layer. Nonetheless, a nonlinear connection existed between temperature and voltages, with higher temperatures resulting in lower onset voltages. Better performance was seen with a 1% K2CO3 electrolyte compared to 1 M KOH, 1% KHCO3, and water owing to the electrolyte's stronger stability and lower overpotential (Fig. 53d). Finally, the 1% K2CO3 electrolyte at 60 °C showed the greatest performance, with 0.5 A cm−2 for 1.95 V. In addition, a durability test over 200 h demonstrated the same promising results for hydrogen production, with the performance remaining almost constant at 0.5 A cm−2 at 2.08 V. LDHs generally have high ionic conductivity, can transport more OH− ions, and are quite stable over the long term. To investigate water electrolysis in an alkaline environment containing 0.1 M NaOH, Zeng et al. developed an inorganic-based AEM (cold pressed Mg–Al LDH membrane).354 At a temperature of 70 °C and a cell voltage of 2.2 V, the integrated inorganic membrane-containing MEA demonstrates an electrolysis performance that produces a current density of 0.208 A cm−2. The findings of the cell durability test showed that the voltage of the cell grew slightly during the whole of the study, exhibiting strong stability when using either 0.1 M NaOH (voltage increased from 1.88 to 1.95 V) or 0.1 M Na2CO3 (voltage increased from 2.03 V to 2.09 V).
Under various conditions, including the electrolyte concentration and temperature, the electrolysis performance of dense side chain-functionalized partially-fluorinated PAES membranes was evaluated.355 It was observed that the electrolyte concentration (0.5–2.0 M NaOH) and temperature (20.0–80.0 °C) affected the amount of current that was flowing. This indicates that the resistance dropped as the electrolyte content and temperature rose. The current density that could be attained in this investigation was 1.2 A cm−2 under 2 M NaOH and at a temperature of 80 °C. The membrane's durability (e.g., PAES-TMI-0.25) was determined by exposing it to 2 M NaOH at 0.5 A cm−2 at 30 °C for 480 h. It was discovered that there were slight variations in the voltage over time and that the voltage decreased as the concentration of NaOH increased. This was considered an indicator of the durability of the battery's performance. A study evaluated a partially-fluorinated PAES membrane (e.g., PAES-MQA-0.18) with extremely flexible side chains and numerous QAs for water electrolysis at two different NaOH molarities (1 and 2 M).356 The AEMs of such types exhibit higher performance because of their strong ionic conductivity (125.5 mS cm−2). As a result, the membrane produced a high current density (1.12 A cm−2) at a voltage of 2.0 V while operating in a 2 M NaOH electrolyte at 30 °C. When the electrolyte concentration was 1 M NaOH, the same membrane could not demonstrate a superior current density (0.50 A cm−2). Therefore, the electrolyte solution's concentration could improve the AEMWE's current density. At 0.5 A cm−2, 2 M NaOH, and 30 °C conditions, the durability test was high up to 480 h, and a very low degradation rate was discovered. It could be due to partial fluorination, numerous QAs, and flexible side chains that helped to increase the membrane's alkaline stability.
Recently, it was shown that morpholinium-tethered polyketone (e.g., PK_0.5 step 3) with low conductivity (0.029 mS cm−1) had extremely poor water electrolyzer performance, only showing a current density of 0.15 A cm−2 at 2.0 V in pure water at 80 °C.357 The AEM's poor ionic conductivity may have caused the low current density. Therefore, mopholinium may not be the best modification to enhance the polyketone membrane's characteristics. To guarantee a secure and productive operation of electrolyze, one of the most significant tasks is the measurement of hydrogen crossover. Accordingly, Bensmann and colleagues researched an in situ approach for determining the hydrogen gas crossover that occurs during the water electrolysis of polymer electrolyte membranes.358 The electrolysis measurement is based on the electrochemical output of the flux, which can transform the mass flux into an electric current. The method that has been suggested has a simple setup and a high degree of measurement precision; these qualities make it suitable for use in electrolysis applications at an industrial level.
In another research, increasing the concentration of an ionomer (e.g., TMA) with MEA resulted in higher current density when compared to bare MEA under pure water-electrolyte conditions, and the influence of phenyl groups-loaded ionomer was shown to be identical.360 Nevertheless, the current density varied depending on the electrolyte solution, with pure water obtaining a greater current density than the NaOH solution owing to an acidic phenol from the phenyl group being neutralized by the NaOH solution. Moreover, the same research discovered that the durability of higher amounts of ionomer MEA was poorer than smaller amounts of ionomer MEA because more elevated amounts of MEA do not retain the catalyst particles throughout the operation. Similarly, in another study, MEA was equipped with various commercial AEMs such as Sustainion, AEMION, and A-201, which showed that Sustainion obtained better electrochemical performance (Fig. 53e), i.e., the resistance was less than that of others under low concentration KOH solution at a current density of 0.5 A cm−2 and cell voltage of 2.1 V.359 Additionally, the cell voltage of Sustainion-MEA was decreased with increased temperature due to increased membrane resistance. Therefore, the temperature does not play a greater role in the commercial AEM for cell efficiencies.
AEM has high conductivity, high mol. weight, and high mechanical strength, which might be advantageous for water electrolysis. Accordingly, Miyanishi et al. promised that the polyfluorene-based AEM (such as PFOTFPH-C6-TMA) could be highly useful in water electrolysis because they had high conductivity (100 mS cm−1), high mol. weight (170–240 kDa), low swelling, high mechanical strength (25–42 MPa), and high chemical stability (under 8 M NaOH).100 Moreover, they possessed minimal swelling characteristics, which may have been an advantage of the AEM. Recently, high conductivity (123 mS cm−1), durability, and efficiency of 7-bromo-1,1,1-trifluoroheptan-2-one-incorporated poly(biphenyl alkylene) (e.g., PBPA) AEM were discovered to have excellent performance in water electrolyzer, which obtained a very high current density 4.0 A cm−2 at an applied voltage of 2.0 V, using 1 M KOH at 80 °C.361 Furthermore, the PBPA membrane-based AEMWE responded effectively to increased KOH concentration from 0.1 to 1 M, corresponding to an increase in the current density from 1.3 to 2.8 A cm−2. The excellent durability of the PBPA membrane was discovered to be very high, with no significant changes in the voltages up to 3500 h at the conditions of 1 A cm−2, 1 M KOH, and 60 °C, which could be attributed to the incorporation of fluorine moiety in the membrane, which is attributed to the membrane's high alkaline stability. Because of its great durability (membrane degradation rate less than 6.5 μV h−1) and high current density (4.0 A cm−2), PBPA might be a viable choice for real-time hydrogen synthesis utilizing a water electrolyzer. Due to the hydroxyl groups present in its structure, the crosslinker tris(dimethyl aminomethyl) phenol (TDMAP) may form a hydrogen bond with H2O, increasing the conductivity as it does so. Accordingly, TDMAP-crosslinked bromohexyl pentafluorobenzyl SEBS membrane (e.g., TDMAP-50%-SEBS) exhibits excellent conductivity (109.9 mS cm−1), yet this crosslinked SEBS membrane may demonstrate current density of 1.19 A cm−2 at 2.0 V under the 1 M KOH at 70 °C.362 Crosslinked SEBS performed better in water electrolysis (1.19 A cm−2) compared to the FAA-3-50 membrane (0.96 A cm−2), which may be related to FAA-3-50's lower conductivity (80.4 mS cm−2). However, no study on durability was found for this investigation.
The stability of the functionalized methacrylate monomers with the fluorinated alkane side chain (2,2,2,-trifluoroethyl methacrylate (TFEMA)) to the quaternized poly(DMAEMA-co-TFEMA-co-BMA) membrane (e.g., QPDTB) was found to be the current density, which remained steady for 300 min.363 However, the water-electrolyte used in the AEMWE with QPDTB achieved a low current density (0.1 A cm−2) at a cell voltage of 1.9 V when the temperature was kept constant. Perfluorinated commercial membrane Aquivion® (e.g., AQ720A4-Q) integrated QAs achieved excellent current density (0.9 A cm−2) at a voltage of 2.2 V and functioned at high temperature (90 °C). No durability testing was carried out in this study.364 According to Chen et al., the characteristics of QAs-tethered poly(fluorenyl ether) (PFEQA) membranes make them highly applicable in water electrolysis.365 These characteristics include high hydroxide conductivity, mechanical strength, and relatively moderate chemical and thermal stability. The developed AEMs showed features capable of use in water electrolyzers, which can function in pure water without electrolytes. TMA-functionalized poly-(fluorene-alt-tetrafluorophenylene) (e.g., PFT-C10-TMA (MEA-1)) has high durability and is very effective despite its 56 m thickness. Under 1 M KOH at 80 °C, AEM was measured at 1.8 V, showing 1 A cm−2 and at 0.1 V, showing 0.0 A cm−2.366 As a result, applying voltages is crucial for achieving a current density. Up to 1000 start–stop cycles, the voltage of the MEA-1 was seen to remain unchanged, demonstrating its outstanding durability. However, the PFT-C10-TMA membrane (MEA-2) was tested after being adjusted to a thickness of 35 μm. Current density (1.0 A cm−2) and durability (1000 start–stop cycles) were determined to remain unchanged at the same voltage. Due to their very durable structure, the generated fluorocarbon AEMs (MEA-1 and MEA-2) may be a suitable candidate for real-time use in hydrogen generation.
Suzuki cross-coupling was employed to produce aryl ether-free and propyl spacers in a polyfluorene-polymer membrane (e.g., PFPB-QA) with pendent ammonium in the side chains, which was tested for water electrolysis.367 The polymer's propyl spacer may have the advantage of increasing flexibility and having the capacity to contain a large ionic cluster. In this work, water electrolysis was tested using AEMs and ionomers. Accordingly, PFPB-QA membranes employing ionomers such as PFPB-QA and QA-PPO compared these results to commercial membrane FAA-3-50 using ionomers such as FAA-3, PFBQA, and QA-PPO. It showed that PFPB-QA produced current densities of 1.53 and 1.19 A cm−2, respectively, using the ionomers PFPB-QA and QA-PPO. Contrarily, the ionomers (FAA-3, PFBQA, and QA-PPO) utilized in FAA-3-50 exhibit lower current densities (0.942, 0.946, and 1.05 A cm−2). As a result, it was investigated using a synthetic polymer membrane known as an ionomer that worked well in the electrolysis system. It showed a current density of 1.53 A cm−2 at a voltage of 2.0 V in 1 M KOH electrolyte at a temperature of 70 °C. This may be because the PFPB-QA membrane had a greater conductivity (122 mS cm−1), but the FAA-3-50 membrane had a lower conductivity (76.9 mS cm−2). Additionally, a durability test over a very short time (0.4 h) was performed, and the results showed that PFPB-QA was almost stable up to 0.4 h, whereas FAA-3.50 was unable to do so because PFPB-QA has an aryl-ether-free structure and FAA-3.50 has an unstable PEEK structure. Table 8 shows various AEMs used in water electrolysis, and their corresponding parameters, including current density and durability, have been summarized.
To make large quantities of hydrogen at an industrial scale, researchers need to find low-cost, highly stable electrocatalysts with identical catalytic performance to catalysts for water splitting at greater current densities. Any acidic, basic, or neutral environment suits HER and OER. HER works well in acidic electrolytes because there are enough protons (H+) to adsorb on the electrode surface and react. Still, utilizing alkaline and neutral electrolytes is not easy because water dissociation weakens the kinetic process.369 In contrast, basic electrolytes are highly favored as OER catalysts for water splitting because of their superior activity and great stability. At the same time, acidic solutions cause the catalyst to disintegrate and quickly inactive the catalyst. HER or OER catalysts do not perform significantly in a neutral medium. Suppose the neutral-based water-splitting process could be satisfied by the catalyst. In that case, the real-time seawater splitting system may be conceivable without the desalination matter for pH settings. This would produce a greater quantity of hydrogen in a more environment-friendly manner. Metal oxides, hydroxides, perovskites, and carbon-based electrocatalysts have been used for water-splitting. Carbon-based catalysts are particularly effective in producing OER catalysts because of their large surface area, high conductivity, and excellent corrosion resistance. Despite this, they have a high cost of synthesis and are a notoriously difficult technique. On the other hand, scientific communities are trying very hard to synthesize an effective OER catalyst with a low price and a straightforward procedure for employing it in AEMWE. Therefore, a discussion will be made of low-cost-based effective electrocatalysts for the catalytic process, leading to hydrogen generation in an AEMWE system in the following sentences.
For hydrogen to be produced via water electrolysis, purified water is required. On the other hand, if only clean water were utilized to generate hydrogen everywhere, then significant problems with water supply may emerge. Therefore, the electrolysis of sea salt water, which is one of the most plentiful characteristics in the world and is regarded as a component bestowed by nature in the generation of hydrogen and seawater desalination, without concerns about the use of clean water, is viable options. Accordingly, the AEM-water electrolyzer (AEMWE) that Park et al. developed with a composed Ni–FeOOH (AEMWE-NF) catalyst was evaluated for its electrolysis capability using 1 M potassium hydroxide and seawater.370 The constructed configuration revealed that a current density of 0.729 A cm−2 could be attained, demonstrating a cell efficiency of 76.3%. Additionally, they examined the electrolysis performance of AEMWE-IrO2, which revealed a lower performance (73.2%). These results demonstrated that performing AEMWE utilizing seawater while maintaining high efficiency is feasible. The endurance of these catalyst-composed AEMWE was examined, and the results showed that AEMWE-NF was superior to IrO2-composed AEMWE (high voltage loss) over 15 h at 0.5 A cm−2. Due to the superhydrophilic surfaces of the Ni–FeOOH catalyst, the AEMWE-NF had a greater efficiency than the AEMWE-IrO2, which was the case. Therefore, the Ni–FeOOH surface possesses the superior property that generated oxygen gas is swiftly removed from the surface of the electrode, which led to cell voltage being proficiently used due to the mass transference resistance in the electrolyzer being efficiently minimized. This resulted in the cell voltage being used to its full potential.
The OER is an important process in various energy storage applications, such as water electrolysis; however, its sluggish reaction and large overpotentials make it difficult to employ in these applications. For optimal oxygen evolution reactions (OER), an electrocatalyst has to have a lot of surface area and strong electrical conductivity. Many investigators have exerted great effort for seeking ideal catalysts to enhance the reaction speed and stability. In light of this, Busacca et al. developed Ni–Mn and Ni–Co-based catalysts on carbon fiber for use in OER in AEMWE.371 The electrochemical test was carried out in water and 6 M KOH solution. The results showed that the Ni–Mn-based catalyst performed better due to its high electrical conductivity. Additionally, the OER activity of both the catalysts improved with increasing temperature when tested in the alkaline solution, in contrast to when tested in water. Recently, the active bifunctional electrocatalyst graphitic nitride-carbon fiber (e.g., g-CN-CNF-800) was designed by Park et al. for AEMWE testing, and the electrochemical performance test was studied by modifying the loading of the catalyst.372 The investigation discovered that loading a thicker catalyst layer might reduce the electrolysis performance. The researchers then refined the catalyst loading, and they found that 6 mg cm−2 led to achieving the optimum current density (0.734 A cm−2) when the voltage was set to 1.9 V (Fig. 54a). In addition, g-CN-CNF-800 shows an excellent electrolysis performance compared to IrO2 catalyst, i.e., the performance of the carbonaceous (g-CN-CNF-800) catalyst was 53% greater than IrO2 at a current density of 1.9 V. Due to their outstanding electrochemical activity and stability for energy conversion in cell operations, the functionalization of NiFeOOH on nitrogen-doped carbon-based materials (e.g., a-NiFeOOH/N-CFP) is now a desirable electrode.373 Additionally, doping a high electronegative nitrogen element into a carbon substrate may provide the carbon atoms with an unequal supply of electrical charge, which enhances the oxygen electrochemistry on nitrogen-doped carbon materials. Additionally, plasma doping nitrogen enhances the electrode's surface wettability, which enhances the integration of electrodes with the carbon substrate. The amorphous phase and nitrogen-doped carbon support were the only compensations to achieve a current density.
![]() | ||
Fig. 54 (a) Schematic depiction of AEMWE with catalyst, g-CN-CNF-800 (anode) and Pt/C (cathode), polarization curves for AEMWE with different amount loadings of g-CN-CNF-800 and cell performance test of AEMWE with catalysts such as g-CN-CNF-800 and IrO2, reproduced with permission from ref. 372, Copyright 2018, Elsevier. (b) Schematic depiction of the AEMWE (anode and cathode catalysts: NiFeS@Ti3C2 MXene), polarization curve for the catalysts, and durability test of different catalysts, reproduced with permission from ref. 376, Copyright 2023, Elsevier. |
Proton exchange membrane water electrolyzers (PEMWEs) and alkaline water electrolysis (AWE) are well-established industrial technologies for producing H2. Despite this, AWE has some demerits, such as forming toxic carbonates after OH radicals react with CO2 in the atmosphere to reduce the IEC and the system's overall effectiveness.125,374 As for PEMWEs, they use expensive platinum (Pt) metal electrocatalysts, which makes them a high-priced technology, with the possibility of degradation when long-term stability testing is conducted. It is also not durable when combining a PEMWE with renewable technologies.59,374,375 In addition to these two approaches, AEMWEs, which combine the benefits of PEM and alkaline electrolyzers, are now the subject of intense study. AEMWEs have a compact design, superior carbonate formation prevention, and employ non-precious metal catalysts such as Ni, Fe, and Co. In a while, an environment with a high alkaline concentration and high temperatures may cause the membrane to degrade. Therefore, Niaz et al. recently looked into the effects of two distinct operating voltage conditions on the deterioration behavior of AEMWE over 1000 h and carried out an experiment to find a major degradation factor concerning the operation modes, such as short and long periodic operating modes.374 They observed that the long periodic operating mode resulted in superior stability, with the ohmic and non-ohmic electrical resistance staying practically constant throughout the test. To develop novel catalyst materials with high catalytic activity, high conductivity, and electrocatalytically excellent catalyst supports for electrolysis, Chanda et al. constructed NiFeS–Ti3C2 MXene electrocatalysts on Ni foam framework for AEMWE.376 Compared to electrocatalysts such as Pt/C–RuO2 cells in AEMWE, it achieved cell efficiency of 67.6% and attained a current density of 0.4 A cm−2 at a cell voltage of 1.85 V. The study's durability was found to be quite good, allowing continuous operation of the cell for up to 750 h with little fluctuations in current density (Fig. 54b). Because of the low running expenses of this sort of system, the proposed electrolyzer is among the best-performing AEMWE systems in recent years.
Another research discovered that the current density was dependent on the electrocatalysts; for example, at a voltage of 1.85 V, NiCoO–NCo/C (0.504 A cm−2) exhibited a larger current density than NiCo/C (0.071 A cm−2), which showed a lower current density in AEMWE in the single-cell setup.377 For this reason, the form of NiCoO–NCo/C electrocatalysts that is used is a crucial consideration for use in water electrolysis. In addition, it was observed that the durability of the electrocatalyst was very steady for 10 h, but the NiCo/C electrocatalyst showed a rapid increase after just 3 h of use. Moreover, operating a 5-cell setup with electrodes such as NiCoO–NiCo/C (cathode) and CuCoO (anode) obtained a high current density of 0.74 A cm−2 at 9.25 Vstack (Fig. 55a–e). The 5-cell stack AEMWE demonstrated superior results than the single-cell AEMWE, and the current density grew quickly while utilizing this method. The 5-cell configuration's endurance was discovered to be degrading at a rate of 2.0 mV h−1. Furthermore, it was found that the cell efficiency was initially 69% (0 h), but it later revealed an efficiency of 59%. High-purity H2 (99.95%) may be produced using the newly invented 5-cell system, equivalent to commercial grade H2 (99.999%).
![]() | ||
Fig. 55 (a) Photograph of 5-cell stack AEMWE, (b) polarization curve for the catalyst NiCoO–NiCo/C in AEMWE, (c) durability and cell efficiency test of 5-cell stack AEMWE (1 M NaOH at 50 °C), (d) photograph of the H2 generation setup using 5-cell stack system, and (e) comparison of the purity of generated H2 with that of H2 produced standard method, reproduced with permission from ref. 377, Copyright 2021, Elsevier. |
Recently, the electrocatalyst with chemically etched CuCo-based AEMWE has shown excellent performance in oxygen evolution reaction (OER).378 This system also has a higher current density (1.39 A cm−2) owing to the electrocatalyst's high electrical conductivity. The durability of the CuCo-based AEMWE was tested for 64 h and found to be good. The cell voltage slightly increased from 1.66 to 1.74 V (Fig. 56a). In contrast, the IrO2 electrocatalyst-based AEMWE system showed lower performance in OER, and the obtained current density was only 0.9 A cm−2 due to the low conductivity of the electrocatalyst. Similarly, for AEMWE, an effective electrocatalyst (such as H–Co0.9Fe0.1-CNF) was used.379 It was discovered that the cell voltage was almost stable and only increased slightly with the current density up to 1.7 V with a high current density of 0.794 A cm−2, which may have been caused by the electrocatalyst's high electrical conductivity (Fig. 56b). The H–Co0.9Fe0.1-CNF-based electrocatalyst was found to have good performance in terms of stability, with no noticeable variations in cell voltage up to 290 h. Recently, high AEMWE electrochemical performance was achieved to attain current densities of 1.5 and 1.15 A. cm−2 at 1.9 and 1.8 V, respectively, by adjusting the MEA manufacturing process, MEA characteristics, architectures, ionomer-membrane selection, and operating conditions.380 The parameters of low ohmic, high charge transfer, and high mass transport resistance maximized the cell performance.
![]() | ||
Fig. 56 (a) Schematic depiction of AEMWE setup, the polarization curves of AEMWE by the performance of electrocatalyst under 1 M KOH solution at 45 °C, inset image is generated hydrogen bubbles through the cathode, durability tests up to 64 h of the designed AEMWE (at a current density of 500 mA cm−2 under 1 M KOH solution at 45 °C), reproduced with permission from ref. 378, Copyright 2020, Elsevier. (b) AEMWE polarization curve of H–Co0.9Fe0.1-CNF in 1 M KOH at 80 °C and durability test of electrocatalyst for 290 h, reproduced with permission from ref. 379, Copyright 2020, Elsevier. |
Researchers have been drawn to transition metal sulfides because of their low cost, rapid charge-transfer kinetics, high conductivity, high active sites, high mechanical strength, stability, and high storage. Nevertheless, nickel sulfides in crystalline form have strong OER energy resistance. As a result, by building a heterojunction interface, the performance of OER might be improved by tuning the electrical structure. A catalytic arrangement based on heterojunctions might reduce the energy consumption, enabling a low-cost hydrogen economy. As a result, Wang et al. developed a high-conductivity heterojunction-structured Ni2P/Ni7S6 catalyst for OER performance.381 Compared to a single catalyst system (Ni2P, Ni7S6 and Ni), they discovered that the new heterojunction-structured catalyst needs overpotentials of 330 mV to attain 1.0 A cm−2 and is durable at 0.5 A cm−2 for 700 h. Compared to lower temperatures, the heterostructure catalyst on Pt/C at 75 °C revealed the best cell performance, with a current density of 1.0 A cm−2 at 1.88 V. The developed catalyst-supported AEMWE exhibits 1.0 A cm−2 of current density for 140 h at 1.88 V and 75 °C in a 1 M KOH environment. Substantial water splitting for industrial use might be accomplished using the advanced heterostructure electrocatalyst since it is very active and stable. Pt is a very active catalyst for HER; however, its sluggish kinetics results in a high cost due to the large loading of Pt quantities needed. Intermetallic structured catalysts showed better HER activity in AEMWE than Pt/C catalysts, suggesting a potential solution. As a result, very stable intermetallic MoO2/Ni cathode and Ni foams were prepared for AEMWE HER testing.382 A current density of 0.55 A cm−2 was obtained at 2 V at 60 °C and was found to be stable for more than 300 h at 1 M KOH at a current density setting of 0.5 A cm−2. The highly porous structure of MoO2/Ni in this work enables high active areas for water electrolysis, and the study was built up to make MoO2/Ni with an extent of 78.5 cm2, which was evaluated in a three-cell AEMWE stack and might be beneficial for industrial use.
To achieve optimal cell efficiencies, it is essential first to develop OER catalysts that are both effective and economical. As a result, one of the low-cost-based binaries or ternary electrocatalysts might play a significant role in the OER process for improved water electrolysis applications. Accordingly, the researchers observed that ternary electrocatalysts exhibited greater electrochemical performance than binary electrocatalysts in the AEMWE system.383 Moreover, when placed in an alkaline solution, the ternary electrocatalysts offered a very low OER overpotential. It was discovered that the ternary electrocatalyst had a durability of not changing at a current density of 0.5 A cm−2 for up to 3.5 h in a highly alkaline medium at 60 °C. The electrocatalyst Co–Fe ratio was recently modified and tested for electrochemical performance in AEMWE.384 The results showed that the onset potential decreased as the catalyst Co content increased, while the OER activity increased as the catalyst Co content increased. Examining the stability of AEMWE with Co–Fe and Pt/C catalysts at 1.8 V reveals that the current density did not change after 16 h. Recently, a porous transport layer on the anode material considerably impacted the OER's AEMWE.385 Higher porosity benefited numerous transport capabilities, such as water and oxygen, while decreasing the adherence of the catalyst layer to the porous transport layer. Furthermore, the Ni-based porous transport layer had a lower operation voltage than the stainless steel-based porous layer. Nevertheless, by sintering the materials, the operation voltage may be increased. The thickness of the porous layer (1.2 mm) has a negligible influence on the electrolyzer process. In this study, AEMWE with enhanced porous transport layer material achieved an electrolyzer voltage of 1.64 V at 1.0 A cm−2 and at a cell voltage of 2.0 V, it acquired a current density of 3.5 A cm−2.
Perovskite oxides are extensively used as OER electrocatalysts due to their low cost, high durability, cation deficiency, flexible electronic structure, high conductivity, high catalytic material, and eco-friendliness. Perovskite oxides also have high conductivity and high catalytic material. Hence, Chen et al. developed a hybrid structured perovskite catalyst to boost the oxygen diffusion rate, ultimately increasing the OER.386 The hybrid perovskite has superior OER activity, as demonstrated by the measurement of 0.1 A cm−2 achieved at 0.36 V in an environment containing 1.0 M KOH. Moreover, the hybrid-structured perovskite catalyst was applied as the anode of WEs, and this application was ascribed to a current density of greater than 1.0 A cm−2 at a voltage of 1.8 V. The durability of the AEMWE catalyzed by a hybrid catalyst was evaluated for 40 h at a current density of 0.5 A cm−2; despite this, the AEMWE that was formed exhibited a degradation rate of 3 mV h−1, indicating low durability.
Similarly, the commercial membranes Fumasep and Sustainion were employed as AEMs to investigate the electrolysis of water utilizing Pt/C as an electrocatalyst.387 A schematic diagram of the AEMWE setup is provided in Fig. 57a. At a voltage of 2 V and a temperature of 60 °C, the authors discovered that the electrolysis performance of Sustainion, which attained a current density of 2.74 A cm−2, was superior to that of Fumasep (1.40 A cm−2). The catalyst (Pt/C)-coated carbon paper and carbon cloth testing for electrolysis performance revealed that catalyst-coated carbon cloth had high electrochemical activity. Based on its superior performance, only Sustainion was tested for stability using Pt/C-coated carbon paper and Pt/C-coated carbon fabric for 205 h at 60 °C. Compared to employing Pt/C-coated carbon paper, the Pt/C-coated carbon cloth demonstrated great stability without any changes throughout the test.
![]() | ||
Fig. 57 (a) Schematic depiction of the AEMWE, polarization curve for the AEMWE using a Sustainion® membrane performed through Pt/C coated on cathode materials such as carbon cloth and Sigracet carbon paper, and durability of the AEMWE using Sustainion® membrane at 1 A cm−2 for 25 h on Sigracet carbon and 205 h on carbon cloth, reproduced from ref. 387, Copyright 2023, Elsevier. (b) Schematic depiction of the AEMWE cell setup and inlet gas arrangements for hydrogen pump tests, reproduced with permission from ref. 388, Copyright 2019, Elsevier. |
Hydrogen may be stored under pressure in a cell or stacked on a particular design of an AEM water electrolyzer. The manufacture of compressed hydrogen would facilitate the storage and consumption of hydrogen by allowing for direct storage and enhancing the energy output for impending energy conservation practices. Therefore, Ito et al. established a hydrogen pressure of 8.5 bar in the cell of the AEM water electrolyzer and examined the effect of hydrogen pressure on the electrolysis performance of water, including current density, cell voltage, anode gas, and the quantity of water in the produced hydrogen.388 The investigation findings demonstrated that a pressurized AEM-based water electrolysis method enabled the synthesis of hydrogen with low relative humidity. Schematic depiction of home-made AEMWE cell outlines and inlet gases arrangement for hydrogen pump tests for anode catalysts such as Pt/C (upside) and CuCoOx (downside) are shown in Fig. 57b. The configured cell was utilized for AEM water electrolysis as well as used for the hydrogen-pump test. To measure current in the hydrogen pump tests, the voltage applied to the AEMWE cell by the DC supply is 0.50 V (downside) and 1.80 V (upside). The gas transfer for the anode measurement causes the current to fluctuate over time in these hydrogen pump experiments.
The impact of electrolytes and their various concentrations on the performance of AEMWE was examined.389 In the electrochemical performance, electrolytes such as K2CO3, KOH, and water were utilized, and it was determined that the electrolysis system functioned more efficiently in the favorable environment of KOH or K2CO3 than in water, owing to the membrane's high resistance in the water. In addition, the K2CO3 (around pH of 12) electrolyte performed better than KOH for electrolysis due to its lower resistance, and increasing the concentration (0.1–10 wt%) of K2CO3 resulted in an even greater improvement in electrolysis performance. Researchers have found catalysts that can replace expensive noble metal catalysts, such as metal oxides, spinel ferrites, organic matter, and perovskites, after searching for non-noble-based OER catalysts for several years. Among them, spinel ferrites are drawing interest because of their cheap cost, great durability, and excellent catalytic characteristics, which make them useful in catalysis. But spinel ferrite cannot be used as a catalyst in the electrolysis process for industrial applications because it has low conductivity. However, Pandiarajan et al. recently designed a spinel ferrite (Ce0.2MnFe1.8O4) catalyst as the anode material.124 In brief, Ni was coated on the cathode, and the Ce0.2MnFe1.8O4 catalyst was used as the anode material, and both were set up to provide MEA and AEM (FAA-3-PK-130). In the electrolysis test, the applied cell potential was found to rise slowly, which caused the current density to rise to 0.3 A cm−2 at 1.8 V. In the stability test, the current density (0.39 A cm−2) was almost the same after 100 h at 2 V. Thus, a spinel ferrite-based catalyst could improve the electrolysis performance and keep the OER stable over time. Also, the Pt–Ti meshes were used to make the electrolyzers last longer by reducing anode corrosion.
On the other hand, due to its naturally high availability and large supply, salt water is an excellent source for water electrolysis. However, seawater electrolysis is harder than water electrolysis because salt (chloride anions) produces a chlorine evolution reaction (CER), which increases the resistance of the AEM and could cause all the metals in the cell to corrode. At higher overpotentials, the CER can race with the OER, which causes all metal-based catalysts to corrode. Also, ions in the seawater, such as magnesium and calcium, can settle on the surface of the electrode during the electrolysis process and cover up the active catalytic spaces. Research into the fundamentals of electrolysis in seawater has been extended to prevent these issues. Accordingly, Xing et al. developed active, low-cost, and durable (1000 h) seawater-splitting electrocatalysts (NiFe-LDH/Pt/PTL (anode) and RANEY®-Nickel/Ni (cathode)) for the AEMWEs.390 The performance of the constructed electrolyzer cells, when operated with seawater, shows improvement, and after approximately 1000 h of processing time, the same current density (0.3 A cm−2) was attained.
Transition metal oxides (TMO) or a combination of various transition metals show an excellent performance toward OER catalysis due to the possibility of tunable stoichiometry of the lattice point.391 This could be because transition metal oxides can be made with different transition metals. TMOs are not good conductors of electricity, and the isotherm process in the OER is caused by the oxygen vacancies created by electrocatalytic reduction.392 But making TMO through orderly hierarchical growth in two-dimensional or three-dimensional structures gives the OER high electrochemical stability. Thus, recently, a hydrothermal method was used to make NiFe2O4 in a hierarchical structure to improve the catalytic activity of OER in AEMWE.393 When the hierarchical catalysts were made, it was found that they had a high catalytic activity toward OER. The study (NiFe2O4 catalyst) had a current density of 2.7 A cm−2 at 2.2 V, more than that recorded with an IrO2 catalyst at the anode for AEMWE based on FAA3-50 AEM. However, the 72 h durability test showed that it needed to be fixed to work longer. AEMWE is a new technology that uses platinum-free catalysts and pure water as the electrolyte to make it cheap to make hydrogen, and this kind of electrolyzer also has the capacity to create compact big cell stacks. Thus, considering the problems with PEMWE, an AEMWE has a great chance of supplanting the technologies. For the industrial uses of AEMWEs, it is important to find stable catalysts for OER that don't use platinum metal. To solve this problem, Osmieri et al. made a La–Sr–Co catalyst for OER and studied the effect of ionomer-to-catalyst and binder-to-ionomer on AEMWEs on water and how long they last.394 Both 1% K2CO3 and 0.1 M KOH, which are used in AEMWE operations, reveal that the electrolyzer's performance is far lower than that in water. Table 9 shows various electrocatalysts used in AEMWE, and their corresponding parameters have been summarized.
Catalysts | AEMs | Thickness (μm) | Voltage (V) | Electrolyte | CD (A.cm−2) | T (°C) | Durability | Ref. |
---|---|---|---|---|---|---|---|---|
a AEMs = anion exchange membranes; (A) = anode, (C) = cathode, CD = current density; T = temperature. | ||||||||
IrO2 (A) and Pt/C (C) | X37-50 Grade T (dioxide materials, USA) | — | 1.7 | 1 M KOH + seawater | 0.72 | 50.0 | Very high | 370 |
NiMn2O4/CNF (or) NiCo2O4/CNF (A), and Pt/C (C) | FAA3-50 (Fumatech, Germany) | 50 | 1.8 | 6 M KOH | 0.08 | 50.0 | — | 371 |
IrO2 (A) and Pt/C (C) | Fumapem FAA-3-50 (FuMA-Tech Inc, Germany) | — | 1.9 | 1.0 M KOH | 0.73 | 60.0 | — | 372 |
α-NiFeOOH/N-CFP (A) and MoNi4/MoO2-nickel foam (C) | FAA-3-50 (Fumatech, Germany) | — | 2.08 | 1 M KOH | 0.01 | 50.0 | High | 373 |
NiFe2O4 (A) and NiFeCo (C) | Sustainion 37–50 (dioxide materials, USA) | — | 2.09 | 1 M KOH | 0.93 | 50.0 | Medium | 374 |
NiFeS@Ti3C2 MXene/Ni (A & C) | Sustainion™ X37–50 RT (dioxide materials, USA) | — | 1.85 | 1 M KOH | 0.4 | 50.0 | High | 376 |
CuCoO (A) and NiCoO–NiCo/C (C) | X37-50 Grade T (dioxide materials, USA) | — | 1.85 | 1 M KOH | 0.44 | 50.0 | Medium | 377 |
IrO2 and CE-CCO (A) and Pt/C (C) | X37-50 Grade T (dioxide materials, USA) | — | 1.8 | 1 M KOH | 1.39 | 45.0 | Medium | 378 |
H–Co0.9Fe0.1-CNFs (A) and Pt/C (C) | FAA-3-50 (Fumatech, Germany) | — | 1.7 | 1 M KOH | 0.79 | 80.0 | High | 379 |
IrO2 (A) and Pt/C (C) | FAA-3-50 (Fumatech, Germany) | 50 | 1.9 | 1 M KOH | 1.50 | 50.0 | — | 380 |
Ni2P/Ni7S6 (or) Ni2P (or) Ni7S6 (or) Ni foam (A) and Pt/C (C) | Fumasep FAA-3-50 (Fumatech, Germany) | — | 1.88 | 1 M KOH | 0.1 | 75.0 | High | 381 |
Ni foam (A) and MoO2/Ni (C) | FAA-3-PK-130 (Fumatech, Germany) | — | 2.0 | 1 M KOH | 0.55 | 60.0 | High | 382 |
Ni–Fe–Co (A) | A201 (Tokuyama Corp. Japan) | — | 2.2 | 4 M NaOH | 1.0 | 60.0 | High | 383 |
Co2Fe1 (A) and Pt/C (C) | Fumapem® FAA-3 (FuMA-Tech GmbH, Germany) | — | 1.8 | 1 M KOH | 0.13 | — | High | 384 |
IrO2 (A) and PtNi (C) | XION™-72–10CL (Xergy Inc, USA) | 40 | 2.0 | 0.3 M KOH | 3.5 | 60.0 | Medium | 385 |
Ni foam (A) and Pt/C (C) | X37-50 (dioxide materials, USA) | — | 1.8 | 1 M KOH | 1.0 | 60.0 | Low | 386 |
Stainless steel (A) and Pt/C (C) | Sustainion X37-50 Grade RT | — | 2.0 | 1 M KOH | 2.74 | 60.0 | High | 387 |
CuCoO (A) and Pt/C (C) | A201 (Tokuyama, Japan) | 28 | 2.2 | 0.1–10 wt% K2CO3 | 1.0 | 50.0 | — | 389 |
Ce0.2MnFe1.8O4 (A) and Ni (C) | Fumasep® FAA-3-PK-130 (Fumatech, Germany) | 130 | 1.8 | Water | 0.3 | 25.0 | High | 124 |
NiFe-LDH/Pt/PTL (A) and RANEY®-Nickel/Ni (C) | Sustainion® X37-50 Grade T (dioxide materials, USA) | — | — | Seawater | 0.3 | 60.0 | High | 390 |
NiFe2O4 (A) and Pt/C (C) | Fumasep® FAA3-50 (Fumatech, Germany) | — | 2.2 | 1 M KOH | 2.7 | 60.0 | Medium | 393 |
LaxSr1−xCoO3−δ (A) and PtRu/C (C) | PiperION-A40-HCO3 | 40 | 1.9 | Water | 0.3 | 70.0 | Medium | 394 |
Studies should concentrate on the stability of electrocatalysts and the maintenance of their performance in neutral environments. Additionally, in water electrolysis systems based on a single-cell stack, there have been relatively few reports of the performance of the water electrolysis equipment deteriorating. Low alkaline durability and low conductivity of membranes are the major problems that decrease the ohmic drop, which is attributed to the energy loss in the AEMWE. A complete understanding of catalysts in electrocatalytic mechanisms should be further investigated. Many electrocatalysts have been analyzed considering performances for water electrolysis, and their catalytic mechanism in the cell is not yet fully explored. The practical use of the developed non-PGM-based water electrolysis catalysts could be a potential direction in industrial applications. Using alkaline stable and low-cost AEMs with non-PGM catalysts will be appropriate for reducing the total cost of H2 production and the operation of AEMWE. The start-up expense of the whole system may be reduced with the use of noble metal-free catalysts in AEM water electrolysis. Ni–Fe and other metal compositions make up the vast majority of OER catalysts; this might be the reason for their relatively cheap cost and average efficiency. In the future, the research and development of PGM-free, low-cost catalysts that can achieve high catalytic activity reactions will be an important part of the water electrolysis process. The performance of hydrogen-generating systems operating at laboratory and industrial scales must be compared, and laboratory AEMs must be scaled to industrial cell stack systems.
Microphase separation morphology (MSM) is a crucial key for the AEMs for ionic conductivity, and SAXS, AFM, TEM, and SEM can confirm these morphologies of AEMs. It was found that MSM is founded on a variety of hydrophobic and hydrophilic configurations in the membrane. Degree of quaternization, a higher level of hydration, alkoxy groups and extender with side chains, and increasing the number of quaternary amines groups are all necessary to achieve good MSM. Interestingly, the ability of the membranes to segregate microphases was significantly improved by adding a fluorinated component to the polymer backbone. New structural insight into the possibility of a high conductivity AEM is obtained by maintaining consistency between hydrophilic and hydrophobic phase modification during microphase separation. 1H NMR may confirm the organic structures of AEMs, although well-soluble polymers in deuterated solvents provide the best results. FT-IR is a low-cost analytical method for studying organic and inorganic compounds and their composite structures. It might benefit researchers globally due to its accessibility and affordability. XPS detects surface elements of modified and unmodified organic/inorganic composites. Due to its high cost, rural researchers may find it difficult to employ this technique for polymers and their organic/inorganic composites. Aliphatic side chains and densely structured QAs-tethered AEMs showed high thermal stability. Unlike single polymer-constructed membranes, composite membranes have better thermal stability. The fluorination of hydrocarbon polymers results in increased membrane thermal stability.
Conductivity and water uptake determine membrane IECs. High IEC may cause high conductivity and water absorption. Therefore, membranes with high conductivity increase the IEC. Due to their strong conductivity, fluorinated membranes have higher IEC than hydrocarbon membranes. Membrane hydrophilicity affects water intake and swelling ratio. Higher hydrophilic may increase water absorption and swelling ratios, thus weakening membrane dimensional stability. Organic/organic composite hydrocarbon membranes increased the WU compared to single and inorganic/organic polymer membranes. Hydrophobic copolymer-structured AEMs with hydrophobic extenders to the hydrophilic backbone have lower water absorption and swelling ratio. Reinforced membranes decrease water swelling and improve the membrane stability. Fluorinated membranes are hydrophobic and have limited water absorption compared to hydrocarbon membranes. Water absorption and membrane MSM influence IC. Due to water absorption, hydrophilic membranes may have higher IC than hydrophobic ones. The organic/organic composite's IC relies on the membrane's hydrophilic groups. However, the IC of these composite membranes could be lower that of the inorganic/organic composite. Fluorinated membranes have higher IC than hydrocarbon membranes because fluorine's electronegativity facilitates ion separation and movement. The very low IC in AEM suggests that it may be chemically unstable due to OH− assault on the immobile ion. To a large extent, the conductivity of an AEM may be hampered by the Hoffmann elimination process and the degradation of QAs groups by nucleophilic substitution. The IC is crucial to the AEM's functionality since greater OH− conductivity enables significantly larger current densities.
The membrane's alkaline stability (AS) relies on its functionalized cations such as cyclohexyl substituted QAs, heterocyclic cations, long side-chain quaternary ammonium groups, N-spirocyclic, and N-cyclic piperidinium as well as the high density of cations. Due to their aromatic nature and reduced nucleophilic attack, copolymer-structured AEMs have high AS. High AS membranes with low WU and IEC protect backbones from nucleophilic attack. Pore-filled membrane method can improve the mechanical properties using porous substrates. Inorganic/organic composites (e.g., PBI–DIm–Si) preserved 100% conductivity in alkaline conditions. Fluorine-substituted quaternized AEMs were more stable in alkaline solution than fluorine-free AEMs, and the membrane preserved the conductivity better than hydrocarbon-based membranes. Due to low water absorption, cross-linking improves the membrane mechanical stabilities (MS). TMHDA and DABDA, which are hydrophilic and decrease membrane stiffness, may not improve MS. Due to limited water absorption, inorganic particles strengthen inorganic–organic composite membranes, endowing them with higher MS. Organic or inorganic rigidity may boost membrane MS. Hydrocarbon-based membranes are mechanically weak. Fluorinated quaternized membranes have higher MS than bare ones.
In terms of electrochemical performance, blended AEMs functionalized with imidazolium have excelled. Interestingly, longer side-chain-functionalized AEMs may be more advantageous than shorter side-chain-functionalized AEMs in improving the water electrolysis performance. The electrolyte conditions can affect the performance of AEM in water electrolysis. Therefore, KOH conditions demonstrated higher performance than water. As the temperature rises, the strong ions become more mobile, allowing for a higher current density in the membrane. An increase in current density may not be beneficial if a higher cell voltage is the cause of a higher temperature. Water electrolysis for hydrogen generation may be carried out with lower voltages using membranes with higher conductivities. The membrane's current density depends on the electrolytes because the AEM achieved a lower current density in water than in KOH. In an electrolysis system, the efficacy of the cells may be highly dependent on the electrolyte concentration, i.e., the current density increases as the KOH concentration rises. Compared to commercially available AEMs, AEMs synthesized in the lab exhibited a higher current density. The best results using AEMs to electrolyze water and using electrocatalysts for AEMWE are as follows: the best water electrolysis is shown in 1 M KOH at 80–90 °C with the following components: AEM: qPSU/NIM-N+, catalyst: Ni (anode)-Pt/C (cathode), voltage: 2.2 V. AEM: qPSU, catalyst: Ni (anode)-Pt/C (cathode), voltage: 2.2 V. AEM: PBPA, catalyst: IrO2 (anode)-Pt/C (cathode), voltage: 2.0. These AEMs work effectively to achieve greater current densities of 3.7, 4.2, and 4.0 A cm−2, respectively. AEMWE's best performance so far is 2.74 A cm−2 at 2.0 V for the following configuration: Sustainion X37-50 Grade RT (AEM), stainless steel (anode), and Pt/C (cathode) in 1 M KOH at 60 °C.
This journal is © The Royal Society of Chemistry 2023 |