Yanrong
Liu
abc,
Jiayao
Cui
ad,
Hao
Wang
abc,
Ke
Wang
abc,
Yuan
Tian
ae,
Xiaoyi
Xue
b,
Yueyang
Qiao
b,
Xiaoyan
Ji
f and
Suojiang
Zhang
*abd
aBeijing Key Laboratory of Ionic Liquids Clean Process, CAS Key Laboratory of Green Process and Engineering, State Key Laboratory of Multiphase Complex Systems, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China. E-mail: yrliu@ipe.ac.cn
bLongzihu New Energy Laboratory, Zhengzhou Institute of Emerging Industrial Technology, Henan University, Zhengzhou 450000, P. R. China
cHuizhou Institute of Green Energy and Advanced Materials, Huizhou, Guangdong 516081, China
dSchool of Chemical Engineering, University of Chinese Academy of Sciences, Beijing 100049, China
eTianjin Key Laboratory of Molecular Optoelectronic Sciences, Department of Chemistry, School of Science, Tianjin University, Tianjin 300072, China
fEnergy Engineering, Division of Energy Science, Luleå University of Technology, 97187 Luleå, Sweden
First published on 29th May 2023
As a fuel or energy carrier, hydrogen has been identified as a key way to decarbonize electricity, industry, transportation, and heating sectors. Hydrogen can be produced by a variety of methods, among which water electrolysis driven by renewable energy is sustainable and nearly carbon-free. To use hydrogen widely, storage and transportation over long distances are another key issue. Apart from storage at high pressure and low temperature, hydrogen can be stored in organic compounds via chemical bonding under relatively mild conditions. Efficient utilization of hydrogen includes hydrogen fuel cells as an alternative to internal combustion engines. From the above scenarios, catalysis and reaction media are the key factors for realizing hydrogen energy implementation. Ionic liquids (ILs) offer new opportunities due to their tunable functional groups, low vapor pressure, and stable structures as additives, solvents, and charge transfer materials. ILs are known to produce solid catalysts with controllable properties, decorate solid catalysts with modified electrons and geometric structures, and serve as electrolytes and hydrogen storage media. This review summarizes and recaps the recent progress in how ILs act as a cornerstone to support the production, storage, and utilization of hydrogen. Furthermore, critical challenges and future research directions of ILs in hydrogen energy applications are also outlined.
Hydrogen does not purely exist in nature, and its production relies heavily on energy input. It can be manufactured from various sources, including both renewable and non-renewable sources. Hydrogen production from fossil fuels has reached the highest level of technical readiness, but the process is accompanied by large emissions of CO2. Further purification and carbon capture and storage techniques require additional costs. Hydrogen production through water electrolysis has the advantages of high purity and zero carbon emissions and has been identified as the most promising green hydrogen supply technology for the future.8 The challenge associated with water electrolysis is high power consumption, which originates from the low efficiency of the electrodes, requiring active, durable, and inexpensive electrocatalysts as well as low fluid-resistance electrolyzers.9
Another obstacle that hinders hydrogen energy implementation is the cost and immaturity of hydrogen storage under mild conditions. The compressed-gas and cryogenic-liquid hydrogen storage not only causes high energy penalties and rigorous requirements for containers but may also raise safety concerns. Hydrogen storage alloys have the advantages of high safety and high hydrogen purity but face problems of low gravimetric capacity and slow kinetics. New hydrogen storage media are desirable, such as liquid organics.10 In such media, dehydrogenation is a catalytic process using highly efficient catalytic agents.11 Recently, ammonia has also been realized as a hydrogen storage carrier. Although it requires large amounts of energy for cracking and the subsequent hydrogen compression at the end-user site, ammonia is easily liquefied for transportation.12 Ammonia produced by the traditional Haber–Bosch method results in giant carbon emissions. The new method through electrocatalysis using N2 and water to produce ammonia under ambient conditions is promising but requires electrocatalysts to activate inert N2 and suppress by-product formation.
The hydrogen fuel cell, in which hydrogen reacts with oxygen to produce electricity, is one of the most iconic technologies for the utilization of hydrogen. Hydrogen fuel cells are promising candidates to replace internal combustion engines without the limitations of the Carnot cycle and have been proven to be desirable for heavy-duty vehicles. The critical challenge is the sluggish kinetics of the oxygen reduction reaction (ORR) at the cathode. The current catalyst of choice is platinum (Pt), but its cost, mass activity, and durability are still unsatisfactory. Recent efforts have concentrated on reducing the Pt loading through alloying with Earth-abundant metals. In addition, new catalytic interfaces are also desirable to improve ion and charge transfer as well as mass transport, and to avoid corrosion.13
To ensure that hydrogen energy is more secure and affordable to release its potential, critical issues must be addressed. The abovementioned scenarios for hydrogen production, storage, and application share common problems of high cost, low efficiency, and poor stability of catalysts. Ionic liquids (ILs) could offer a solution to these problems. ILs are capable of facilitating electrocatalysis and thermocatalysis in the following ways: (1) ILs are known to act as surfactants or stabilizers to regulate the shape, particle size, and defect level of catalyst particles, thereby enhancing catalyst intrinsic activity and active site density;14 (2) ILs can be designed to contain metal and metal-free elements, which serve as sacrificial precursors to produce catalysts;15 (3) ILs can be immobilized on the surface of a solid catalyst to form a solid catalyst with an IL layer (SCILL) and regulate catalytic activity, durability, and selectivity;16 (4) ILs are favorable electrolytes or electrolyte additives due to their high ionic conductivity;17 and (5) ILs can serve as hydrogen storage media that can be hydrogenated based on imidazole or phenyl rings18 or can serve as additives to activate amineborane-based materials.19 Although ILs have the advantages described above, there are still concerns about their biological and environmental toxicity, which require careful handling during the process.20
Therefore, improving the intrinsic activity and active site density by rational design of ILs and IL–particle interactions is a new concept for enhancing the efficiency of electrocatalytic and thermocatalytic processes involved in hydrogen energy. In this respect, ILs will serve as a new cornerstone to support the implementation of hydrogen energy. ILs have been proven to be excellent electrolytes and catalysts and assist in materials synthesis.21–25 While some insightful reviews have summarized IL applications in several scenarios, such as the hydrogen evolution reaction (HER),15,26 hydrogen storage27 and fuel cells,28 the overall picture of how ILs support hydrogen energy is still missing. This review provides a summary of the recent progress in IL assistance in producing solid catalysts and IL–catalyst interactions, and their success in facilitating hydrogen production, storage, and application (Scheme 1). As we have recently carried out comprehensive reviews of the use of ILs in synthetic ammonia,29,30 this part will not be included in this review. Finally, critical challenges and perspectives on future research on IL applications in hydrogen energy are also highlighted.
ILs | Electrocatalyst | Electrolyte | Overpotential (mV vs. RHE)@10 mA cm−2 | Tafel slope (mV dec−1) | Entryref. |
---|---|---|---|---|---|
[BMIM][TfO]: 1-butyl-3-methylimidazolium trifluoromethanesulfonate; [BMIM][BF4]: 1-butyl-3-methylimidazolium tetrafluoroborate; NG: nitrogen-doped graphene; [BPy][Br]: N-butyl pyridinium bromide; [C12MIM][Ac]: 1-dodecyl-3-methylimidazoliumacetate; Co0.6Ni0.4Se2-LN: loofah-like Co0.6Ni0.4Se2; [BMIM]2[MoO4]: 1-butyl-3-methylimidazole molybdate; [P4444][Cl]: tetrabutylphosphonium chloride; [P6,6,6,14]2[CoCl4]: trihexyl(tetradecyl)phosphoniumtetrachlorocobaltate; CNTs: carbon nanotubes; [OA][H2PO2]: N-octylammonium hypophosphite; [BMIM][Cl]: 1-butyl-3-methylimidazolium chloride; NHCS-W: nitrogen-doped hollow carbon spheres with nanoporous shells; Ni-MA: NiCl2·6H2O:malonic acid 1:5; GO: graphene oxide; PVEIB: poly(1-vinyl-3-ethylimidazoliumbromide); PPy: polypyrrole | |||||
[BMIM][TfO] | MoS2 | 0.5 M H2SO4 | — | 156 | 131 |
[BMIM][BF4] | MoS2/NG | 0.5 M H2SO4 | 150 | 48 | 232 |
[BPy][Br] | MoS2 | 0.5 M H2SO4 | 259 | 59 | 333 |
[C12MIM][Ac] | Co0.6Ni0.4Se2-LN | 0.5 M H2SO4 | 163 | 40 | 434 |
[BMIM]2[MoO4] | MoC | 0.5 M H2SO4 | 110 | — | 535 |
[P4444][Cl] | Ni2P | 0.5 M H2SO4 | 102 | 46 | 636 |
[P4444][Cl] | Ni12P5 | 0.5 M H2SO4 | 182 | 80 | 736 |
[P6,6,6,14]2[CoCl4] | Co2P/CNTs | 0.5 M H2SO4 | 150 | 47 | 837 |
[OA][H2PO2] | Ni2P4O12 | 1.0 M KOH | 116 | 97 | 938 |
[BMIM][Cl] | NHCS-W | 1.0 M KOH | 196 | 101 | 1039 |
Ni-MA | NiS2@GO | 1.0 M KOH | 57 | 47 | 1140 |
PVEIB | NiS2-MoS2/PVEIB/PPy/GO | 0.5 M H2SO4 | 205 | 49 | 1242 |
Lau et al. investigated the effects of aromatic and non-aromatic ILs on the synthesis of MoS2 as an HER electrocatalyst. It was found that the MoS2 regulated by [BMIM][TfO] and N-butylpyridinium trifluoromethanesulfonate ([BPy][TfO]) had higher crystallinity and possessed more unstacking/de-layering structures compared to those regulated by non-aromatic ILs,31 and such a high edge exposure was desirable for HER occurrence. The MoS2 synthesized in [BMIM][TfO] had the best HER performance, delivering 529 μA at −533 mV, which benefited from the effects of strong molecular interactions, H-bond acidity, and aromatic properties of ILs on the structure of MoS2. Liu et al. presented the mediation of [BMIM][BF4] to prepare MoS2/NG composites.32 Through electrostatic interactions, [BMIM][BF4] stabilized the edge sites of MoS2 and induced a delaminating morphology of the MoS2 crystals, thereby enhancing the HER electrocatalytic activity. In particular, with the addition of 1.0 mL [BMIM][BF4], the acquired MoS2/NG-IL10 displayed high HER performance and achieved 136.3 mA cm−2 at an overpotential of 300 mV in 0.5 M H2SO4, which is 7.4 times greater than that of MoS2 without [BMIM][BF4] mediation. Zhang et al. used [BPy][Br] as a structure directing agent to synthesize 1T/2H-MoS2.33 The result indicated that the proportion of the metal 1T phase in 1T/2H-MoS2 can be increased due to steric hindrance and the π-stacking interaction of [BPy][Br], thus promoting active site exposure and enhancing charge transfer. When the proportion of the 1T phase in 1T/2H-MoS2 reached 91.9%, the best HER performance was achieved with a Tafel slope of 59 mV dec−1. A facile synthesis strategy to directly grow a network of Co0.6Ni0.4Se2-LN on carbon cloth was investigated,34 where [C12MIM][Ac] played an important role in the construction of the 3D hierarchical architecture. The resulting Co0.6Ni0.4Se2-LN has the characteristic of a layered structure with sufficient active sites and channels for the transfer of electrolyte ions and electrons, leading to a low HER overpotential of 163 mV@10 mA cm−2 in 0.5 M H2SO4.
ILs can also serve as sacrificial precursors, such as metal sources, carbon sources, and heteroatom sources, to produce electrocatalysts through pyrolysis. To use ILs as metal sources, Zhang et al. prepared ordered mesoporous catalysts of MoC@C using [BMIM]2[MoO4].35 Uniform particle sizes (5 nm) of MoC with a highly crystalline structure were obtained due to the action of SBA-15 silica hard-template mediated pore channels, contributing to a low HER overpotential of 110 mV@10 mA cm−2 in 0.5 M H2SO4. Nanostructured Ni2P and Ni12P5 were developed using [P4444][Cl] as the reaction source by a microwave heating method,36 which dramatically reduced the reaction time. In 0.5 M H2SO4, Ni2P has a better HER overpotential of 102 mV than Ni12P5 of 182 mV at 10 mA cm−2. [P6,6,6,14]2[CoCl4] could serve as a source of Co and P for the preparation of Co2P/CNTs by means of a phosphidation step mixed with CNTs,37 and the resulting Co2P was uniformly distributed on CNTs to acquire only 150 mV at 10 mA cm−2 for the HER. Ying et al. used [OA][H2PO2] as a source of P to produce Ni2P4O12.38 The [OA][H2PO2] assisted synthesis not only speeds up the reaction process, but also offers a rich source of P for the manufacturing of metaphosphate, resulting in a low HER overpotential of 116 mV at 10 mA cm−2 in 1.0 M KOH. [BMIM][Cl] was applied as a source of both carbon and nitrogen to prepare NHCS-W.39 The dielectric constant of the solvent (e.g., water and ethanol) affects the dissociation/association of the IL precursor, thus affecting the morphologies and structures of the prepared carbon spheres. A solvent with a high dielectric constant is beneficial for the formation of hollow structures to enhance the water-splitting performance. The prepared NHCS-W was used as both the cathode and anode catalysts, which only required 1.61 V to achieve 10 mA cm−2. An eutectic IL of Ni-MA composed of NiCl2·6H2O and malonic acid with a mole ratio of 1:5 was pyrolyzed to produce NiS2@GO.40 The unique structure of the encapsulated NiS2 in the GO shell improves the kinetics and the exposure of the active sites. Consequently, the NiS2@GO electrocatalyst shows both a reduced HER overpotential of 57 mV and an OER overpotential of 294 mV at 10 mA cm−2 in 1.0 M KOH. The strategy of using 1-butyl-3-methylimidazolium hexafluorophosphate ([BMIM][PF6]) as a dopant to introduce oxygen vacancies in N, P, and F tri-doped MoS2/NPF-CoFe2O4 was studied by Sun et al.,41 showing that [Bmim][PF6] not only modulated the electronic structure of MoS2/NPF-CoFe2O4, but also constructed an amorphous structure with abundant oxygen vacancies, contributing to a low OER overpotential of 250 mV at 10 mA cm−2 in 0.1 M KOH (Fig. 1).
Fig. 1 (a) Schematic illustration of the synthetic strategy for the MNC hybrid, and (b) OER polarization curves without iR-compensation. (c) The elementary steps and binding energies of the intermediates during the OER. Reproduced with permission.41 Copyright 2018, Wiley-VCH. |
Poly(ionic liquid)s (PILs) are a special type of IL that have both ionic properties and long chains, and also have the potential in the preparation of water-splitting catalysts. The PIL of PVEIB was applied to prepare an HER catalyst composed of NiS2-MoS2/PVEIB/PPy/GO.42 PVEIB played a crucial role in combining NiS2-MoS2 and PPy/GO with good stability and induced the formation of the 1T phase in MoS2, leading to a small HER overpotential of 205 mV@10 mA cm−2 in 0.5 M H2SO4. Different morphologies of bismuth sulfides (Bi2S3) were synthesized by Gao et al. using poly(1-methyl-3-(4-vinylbenzyl)-imidazolium chloride) (PIL-1), poly(diallyldimethylammonium bis(trifuoromethanesulfonyl)imide) (PIL-2), and poly(3-ethyl-1-vinylimidazolium bromide) (PIL-3) as additives.43 Notably, owing to the different properties of cations and anions, the backbone architectures of the PILs could be selectively coupled to specific crystal faces, modulated and even templated for sequential growth, thus significantly enhancing their effectiveness in regulating the nucleation and growth of inorganic materials (Fig. 2a–c). As shown in Fig. 2a, nanowires, hexagonal plates, and mesostructured plates of Bi2S3 were obtained by adjusting PIL-1, PIL-2, and PIL3, respectively. The obtained Bi2S3-nanowires showed the best OER performance, close to that of RuO2 at 10 mA cm−2 in 0.1 M KOH. Ding et al. used an IL as the precursor and CoCO3 to form a novel IL containing Co species (Fig. 2d and e),44 The prepared novel IL was further reacted with CNTs by polymerization to form a PIL-Co/CNT electrocatalyst with PIL and Co species well dispersed on the CNT surfaces. CoSSPIL/CNT afforded much better OER activity (1.64 V vs. RHE at 10 mA cm−2) compared to Co3O4/CNT and CoCO3.
Fig. 2 (a) PILs controlled the structures of Bi2S3, (b) OER linear sweep voltammetry (LSV) curves of Bi2S3 and RuO2, (c) Tafel slope of Bi2S3 and RuO2. Reproduced with permission.43 Copyright 2016, Wiley-VCH; (d) preparation procedure of CoSSPIL/CNT, and (e) cyclic voltammogram (CV) of CoSSPIL/CNT compared to different samples of CNT, SSPIL/CNT, Co3O4/CNT, and CoCO3. Reproduced with permission.44 Copyright 2018, Wiley-VCH. |
Fig. 3 (a) OER LSV curves of the bare GC electrode, commercial 20 wt% Pt/C, CoS2 and IL-modified CoS2 and (b) the derived Tafel plots. (c) OER LSV curves of IL-modified CoS2 before and after 1000 CV tests. (d) Schematic illustration of the oxidation of water. Reproduced with permission.46 Copyright 2018, the Royal Society of Chemistry. |
PILs can also be used as surface modifiers. Pham Truong et al. demonstrated that the prepared poly(vinyl-imidazolium-methyl), i.e., poly(VImM), enhanced electrocatalytic activity by modifying the Pt catalyst.47 The hierarchical polymer brush structure accelerated the electron transfer and also the accessibility of water and oxygen. Compared to the Pt/C electrocatalyst, Pt/C/poly(VImM) exhibits a higher current density and lower OER overpotential, benefiting from the chemical composition and unique structure of poly(VImM).
Pandey et al. theoretically studied the geometric and electronic properties of a hybrid catalyst composed of [BMIM][TfO] and (TiO2)n (n = 2–12) nanoclusters, and the HER catalytic activity was predicted by the density functional theory (DFT).48 The calculation results show that the system ΔG of the hydrogen adsorption for the [BMIM][TfO]/(TiO2)5 is close to the ideal value (0 eV), suggesting a superior HER performance than that of the conventional Pt-based catalyst (Fig. 4).
Fig. 4 (a) ΔG for hydrogen evolution at equilibrium potential of the IL and IL/(TiO2)n systems, (b) summary of calculated overpotentials for the IL and IL/(TiO2)n systems, (c) HER activity volcano plot of exchange current [log(i0)] plotted against ΔG of hydrogen adsorption in an acidic medium, and (d) schematic illustration of HER on the IL/(TiO2)5 system. Reproduced with permission.48 Copyright 2021, American Chemical Society. |
Fig. 5 (a) Effects of the potential on the current density with different electrolytes for water electrolysis: (a) [BMIM][BF4], (b) [TEA-PS][BF4], and (c) KOH, and (b) effects of temperature on the current density for 0.1 M [TEA-PS][BF4] at −2.0 V and −4.8 V. Current density (i) is expressed in mA cm−2. Reproduced with permission.50 Copyright 2013, Elsevier. (c) Tafel plots for the HER on Pt/C, PtNi/C and PtMo/C in 0.1 M [TEA-PS][BF4], and the (d) schematic representation of the Volmer–Heyrovsky mechanism for the HER. Reproduced with permission.52 Copyright 2017, Elsevier. |
In addition to adding ILs to aqueous electrolytes, ILs can also serve as polymer electrolyte additives for membrane electrode assemblies (MEAs). The protic IL [Dema-H][TfO] possessed good thermal stability, and high ionic conductivity at intermediate temperatures, and was used by Thimmappa et al. to fabricate a polymer electrolyte membrane based on a polytetrafluoroethylene (PTFE).58 It was found that [Dema-H][TfO] almost completely filled the pores of the PTFE membrane. The low gas permeability was also confirmed by an H2 crossover test, in which H2 flowed to one side of the membrane (1 atm) and nitrogen flowed to the other side (50 cm3 min−1). As the temperature increased, the HER overpotential associated with the reduction of [Dema-H]+ decreased by approximately 200 mV at 100 °C. The decrease in the HER overpotential with increasing temperature can be explained by the enhancement in kinetics and decrease in pKa of [Dema-H]+ with increasing temperature. A single cell water electrolyser with the [Dema-H][TfO] impregnated membrane was successfully operated at 100 °C and 50% RH. This shows good water uptake of protic IL from the vapor phase, which is promising for intermediate-temperature water electrolysis (Fig. 6).
Fig. 6 (a) SEM surface morphology of (a) the unfilled PTFE membrane, (b) [Dema][TfO] impregnated PTFE membrane and below images of both membranes, (b) ionic conductivity of [Dema-H]+[TfO]− filled in the PTFE membrane filter, and (c) OER and HER in IrO2 and Pt/C MEA with humidified (100% RH) N2 at 5 mV s−1 (inset figure for the lower potential region) at 20 °C. Reproduced with permission.58 Copyright 2020, Elsevier. |
In contrast to the well-studied borohydride ILs, also known as tetrahydroborates (containing BH4), octahydrotriborate ILs (containing B3H8−) have also attracted a great deal of interest as intermediates in the thermal decomposition of borohydride ILs. For example, Chen et al.62 prepared guanidine octahydrotriborate ((N3H6)C(B3H8)) from NaB3H8 and guanidine chloride by ball milling or reaction in THF. Theoretically, ((N3H6)C(B3H8)) possesses a hydrogen storage capacity of 13.8 wt% due to the feature of 6H in the cation and 8H in the anion. The corresponding dehydrogenation characteristics were investigated. The dehydrogenation occurred in multiple steps when the solution was treated at 100 °C. After 95 hours, approximately 6.5 equivalents or 12.9 wt% of hydrogen was released, with approximately 4 equivalents of hydrogen being released in the first hour.
In addition to the hydrogen storage properties of ammonia borane and borohydride ILs, five-membered B/N anionic chain complexes and liquid organic hydrogen carrier ILs also showed excellent hydrogen storage properties. For example, the synthesis and dehydrogenation of four kinds of five-membered B/N anionic chain ILs, [Bu4N][BH3(NH2BH2)2H], [Et4N][BH3(NH2BH2)2H], [C(N3H6)][BH3(NH2BH2)2H], and [C(N3H5CH3)][BH3(NH2BH2)2H], were systematically investigated by Chen et al.63Fig. 7 shows the release of hydrogen from these four prepared IL aqueous solutions under the action of catalysts. Among the catalysts investigated, RuCl3 showed excellent activity, catalyzing the release of hydrogen from all four ILs within 2 min (about 33–34 mL, 8 equivalents of hydrogen). The catalytic activities of NiCl2 and Pt/C were similar, but lower than that of RuCl3, while no catalytic activity was observed for FeCl3. This work provides insight into the ILs with the five-membered B/N anionic chain and further development of the chemistry of boron and nitrogen.
Fig. 7 Metal-catalyzed hydrolytic H2 release from the aqueous solution of (a) [Bu4N][BH3(NH2BH2)2H], (b) [Et4N][BH3(NH2BH2)2H], (c) [C(N3H6)][BH3(NH2BH2)2H], and (d) [C(N3H5CH3)][BH3(NH2BH2)2H] at room temperature.63 Copyright 2021, Wiley-VCH. |
Hydrogenation of ILs was performed by Stracke et al.18 and no hydrogenation of the imidazole ring was found, which means that in principle not any imidazole-based IL could be used as the liquid-organic hydrogen carrier (LOHC) system. Subsequently, Deyko et al.64 developed ILs containing carbazole molecules as hydrogen-carrying sites using bis(trifluoromethylsulfonyl)imide imidazolium-based ILs (Fig. 8b), and compared the imidazole cation with carbazole. The results suggest that it was better to use the –(CH2)3– linkers to connect N and Si atoms. N-(3-Trimethylsilylpropyl)carbazole was successfully hydrogenated with high selectivity and no by-product with a low boiling point was formed. When the Pt/C was introduced as a hydrogenation and dehydrogenation catalyst, the theoretical total gravimetric hydrogen capacities of the synthesized ILs were 2.05 and 1.58 wt%, respectively.
Fig. 8 (a) Reaction paths involved in the preparation of imidazolium ionic liquids containing cyclohexane moieties. Reproduced with permission.18 Copyright 2007, American Chemical Society. (b) New ionic liquids for hydrogen storage. Reproduced with permission.64 Copyright 2020, Elsevier. |
Fig. 9 Dehydrogenation yields of EDB supported by imidazolium-based ILs. Reproduced with permission.67 Copyright 2013, Elsevier. |
ILs are also excellent catalysts for LOHC. Søgaard et al. used a cationic iridium (Ir) complex immobilized in [PPh4][NTf2] as a catalyst to study the dehydrogenation process (Fig. 10).69 They combined a molten salt catalyst immobilisation strategy with a homogeneous catalyst to achieve efficient low-temperature hydrogen storage/release of LOHC. Furthermore, the method allowed easy separation of catalyst and LOHC, enabling the release of large amounts of hydrogen and very small amounts of dissolved noble metal complexes.
Fig. 10 Catalytic dehydrogenation of 2-methylindoline to 2-methylindole in the liquid–liquid biphasic catalyst system using Crabtree's catalyst. Reproduced with permission.69 Copyright 2019, The Royal Society of Chemistry. |
Many efforts in the literature are devoted to disclosing the mechanisms underlying the positive effects of ILs. In the case of enhanced proton conductivity, recent work by Wang et al.83 proved proton shuffling promotional effects of ILs on the ORR needed fine tuning. The authors screened a series of ILs as additives to Pt/C for the ORR and found a volcano correlation between ORR activities and pKa values. [MTBD][NTf2] was proven to be an optimal promoter with a pKa of approximately 15, which was close to the pKa of the product of the ORR rate-limiting step: H2O (pKa = 15.7). The “perfect match” of the pKa value and facilitated ORR performance were attributed to the hydrogen-bonding structure and the tunneling transfer of the proton. The larger product of Boltzmann probability and the proton vibration wavefunction overlap the integral of [MTBD]N-H+⋯OHPt, PμSμv2. With matched pKa, the largest PμSμv2 indicates a much higher hydrogen tunneling rate/probability for the ORR rate-determining step. The correlation between IL acidities and promotion effects on ORR activity was also investigated by Wippermann et al.84 In addition, Zhang et al.85 found that the Pt/C ORR performance had a strong dependence on the cationic side-chain length of imidazolyl ILs. If the alkyl chains were too short (e.g., 1-ethyl-3-methylimidazolium bis(trifluoromethanesulfonyl)imide ([EMIM][NTf2])), the IL could not sufficiently enhance ORR activity because of their limited hydrophobicity as well as the low capability of suppressing the formation of nonreactive oxygenated species on Pt.
By contrast, the IL with too long alkyl chains (e.g., 1-hexyl-3-methylimidazolium bis(trifluoromethanesulfonyl)imide ([C6MIM][NTf2]) and 1-decyl-3-methylimidazolium bis(trifluoromethanesulfonyl)imide ([C10MIM][NTf2])) would lead to a significant loss of the electrochemical active surface area (ECSA). The optimal chain length was needed to achieve a balance between effective inhibition of the formation of oxygenated species on the Pt surface and less passivation of active sites. Avid et al.86 further emphasized the roles of IL in the porous structure of the Pt/C in improving the ORR performance. The alkyl imidazolium ILs with different alkyl chain lengths were loaded into the Pt/C and the impacts of the IL chemical structure on the ORR performance were investigated. As shown in Fig. 11a and b, the ILs that could penetrate into the pores of carbon black would significantly enhance the proton and O2 transfer inside the pores. Additionally, the impacts of ILs on the ORR activity were conformational structure-dependent. Fig. 11c and d show that the combination of butyl and methyl functional groups promoted the ORR activity of Pt/C the most. By contrast, the butyl/methyl and butyl/bimethyl functioned imidazolium-based ILs were less beneficial or even detrimental. The authors highlighted the trade-off between proton conductivity and O2 transfer resistance, and pointed out the important roles of electrochemical ECSA change, pore-filling degree, hydrophobicity, and attraction/repulsion with Nafion in the rational design of SCILL towards the ORR.
Fig. 11 (a) The scheme of the ILs filled in the pores of carbon black supports. (b) The TEM images of Pt/C-IL catalysts. The gravimetric performance (c) and geometric performance (d) comparisons of catalysts modified with different ILs. Reproduced with permission from ref. 86. Copyright 2022, Springer Nature. |
Although the IL promotional effects are conventionally ascribed to the enhanced O2 transfer, quantitative investigations on the transfer resistances are relatively rare in the literature. Recently, Huang et al.87 have used the 1-methyl-2,3,4,6,7,8-hexahydro-1H-pyrimido[1,2-a]pyrimidinium 1,1,2,2,3,3,4,4,4-nonafluorobutane-1-sulfonate ([MTBD][C4F9SO3]) IL and boosted the ORR performance of PtCo/C by 48%. By conducting the limiting current measurements at different O2 partial pressures, the O2 transfer resistances were quantitatively determined. The results suggested that the IL resulted in an approximately 15% drop in oxygen diffusion resistances compared to its counterpart in the same electrode structure. In addition, the PtCo/C + IL showed a more durable O2 diffusion resistance during the 30 k cycles of stability tests, which was significantly better than the pristine PtCo/C. Another nontrivial mechanism of the IL-enhanced ORR is that the ILs alleviated the specific adsorption of the sulfonate functional group of Nafion on the Pt surface, which was detrimental to the ORR activity88,89 according to the ATR-SEIRAS method used to quantitatively investigated the specific adsorption of sulfonate anions. In the absence of ILs, the –SO3− specific adsorption was visible at 1125 cm−1. By contrast, after introducing [MTBD][beti] ILs, the vibrational absorption of –SO3− in the IR spectra diminished within the investigated polarization potentials.
Fig. 12 The TEM bright-field images of PtFeNi nanowires + ILs (a) before and (b) after 30 k cycles of durability tests. The ORR activities of PtFeNi nanowires + ILs and control group samples (c) before and after the stability tests. Reproduced with permission from ref. 94. Copyright 2019, Elsevier. (d) The scheme of sequential procedure loading [MTBD][beti] onto Pt/C electrocatalysts and the corresponding TEM image of Pt/C + IL. The (e) ORR activities of fresh Pt/C + [MTBD][beti] and Pt/C. The (f) ORR activities loss, (g) ECSA retaining ratios and (h) Pt concentrations in the electrolyte of Pt/C + [MTBD][beti] and Pt/C after the accelerated durability tests. Reproduced with permission from ref. 95. Copyright 2019, Copyright 2019, American Chemical Society. |
In summary, ILs have been proven effective in promoting the performances as well as enhancing the durability of electrocatalytic ORR via regulating the hydrogen bonding, hydrophobicity, proton/O2 transfer resistance, oxygenated species suppression, etc. Comparing the prosperous new catalyst developments, ILs are still limited to relatively few candidates. Although the general mechanisms underpinning the IL promotional effects have been proposed, quantitative investigations to prove a solid correlation between IL chemical–physical properties and specifically enhanced ORR kinetics are still lacking. The accurate description of the conformational structures of the IL-modified catalysts layer is still ambiguous. As a prosperous research topic of PEMFCs, new ILs with a clear understanding of the correlation between their structure and promotional effects are expected to further advance both the catalytic activities and durability of ORR electrocatalysis.
In the case of ILs acting as catalyst supports and promoters, the understanding of the spatial and electronic relationship between the catalysts and the ILs is still limited. For catalysts with porous structures, the configurations of the ILs filled inside the pores and the effects of the confined structure on the functions of the ILs are less investigated. On the other hand, the PILs can also provide meso/microporous structures for catalysts. The confinement effects of PILs, including electronic, chemical, and steric effects, on the catalytic performances are still ambiguous. Another key issue is the stability of the IL catalyst composite structure. Considering that the IL-promoted catalytic reactions usually take place in a liquid environment with vigorous stirring, the leaching of ILs from the IL catalyst composites needs careful consideration. Finally, with the involvement of ILs in a catalytic reaction makes the mechanisms quite complex. In situ spectrokinetic characterization with high spatial/time resolutions and theoretical calculations on multi-scales are proposed to unravel the accurate and comprehensive reaction networks.
ILs acting as catalyst preparation regulators can enable the production of catalysts with different morphologies, sizes, and structures compared to conventional surfactants. However, the origin of these unique structures and the interaction of ILs with catalysts during the preparation process is less well understood. Time-resolved investigations on the morphology/crystalline structure evolution of catalysts during the IL-assisted preparation are highly recommended. Spectroscopic and theoretical investigations on the interaction between ILs and crystallites are required to disclose the mechanisms underlying the regulation effects.
Compared with conventional hydrogen storage organics, ILs show some unique advantages, including extremely low vapor pressure, high H2 storage density, self-catalytic effects, etc. For their practical application, the weight, cost penalty, and potential biological and environmental toxicity of certain ILs added to the hydrogen storage system would need to be addressed. The reversibility of ILs during H2 storage/release requires further evaluation.
This journal is © The Royal Society of Chemistry 2023 |