Review Article | Open Access
Na Zhu, Kun Zhang, Feng Wu, Ying Bai, Chuan Wu, "Ionic Liquid-Based Electrolytes for Aluminum/Magnesium/Sodium-Ion Batteries", Energy Material Advances, vol. 2021, Article ID 9204217, 29 pages, 2021. https://doi.org/10.34133/2021/9204217
Ionic Liquid-Based Electrolytes for Aluminum/Magnesium/Sodium-Ion Batteries
Abstract
Developing post-lithium-ion battery technology featured with high raw material abundance and low cost is extremely important for the large-scale energy storage applications, especially for the metal-based battery systems such as aluminum, sodium, and magnesium ion batteries. However, their developments are still in early stages, and one of the major challenges is to explore a safe and reliable electrolyte. An ionic liquid-based electrolyte is attractive and promising for developing safe and nonflammable devices with wide temperature ranges owing to their several unique properties such as ultralow volatility, high ionic conductivity, good thermal stability, low flammability, a wide electrochemical window, and tunable polarity and basicity/acidity. In this review, the recent emerging limitations and strategies of ionic liquid-based electrolytes in the above battery systems are summarized. In particular, for aluminum-ion batteries, the interfacial reaction between ionic liquid-based electrolytes and the electrode, the mechanism of aluminum storage, and the optimization of electrolyte composition are fully discussed. Moreover, the strategies to solve the problems of electrolyte corrosion and battery system side reactions are also highlighted. Finally, a general conclusion and a perspective focusing on the current development limitations and directions of ionic liquid-based electrolytes are proposed along with an outlook. In order to develop novel high-performance ionic liquid electrolytes, we need in-depth understanding and research on their fundamentals, paving the way for designing next-generation products.
1. Introduction
The emerging attention to environmental pollution and fossil energy consumption and the rapid promotion and innovation in portable electronic devices have created a surge in demand for more efficient and clean energy storage and conversion devices. This in turn requires the suitable systems to make good use of energy storage in intermittent sources (solar or wind) and the electric or hybrid vehicle power [1]. The rechargeable batteries as efficient energy storage and conversion devices have witnessed the booming development hundreds of years with their intrinsic advantages and greatly advanced and enriched the human event process. Particularly, lithium-ion batteries (LIBs) have attracted considerable attention and dominated the commercial market and academic research in various battery systems in terms of high energy density, high operating voltage, and superior cycle stability [2, 3]. However, the rare lithium sources in the earth’s crust of LIBs limit their market share in electronic device applications. In addition, the safety issues caused by thermal runaway also limit their practical application [4].
Fortunately, several new advanced battery chemistries and concepts (not depend on lithium) have emerged in recent years, which opens up new avenues for research. These chemicals come from varieties of different elements (sodium, magnesium, and aluminum), and these materials can be classed as a metal-ion battery concept. Studies of the chemical composition of these new batteries have shown that, although some lessons can be drawn from lithium-based batteries, new insights are needed based on the chemical properties of the elements themselves to develop a commercializable technology [5–7].
Electrolytes are ubiquitous and indispensable in rechargeable batteries [8]. In all cases, an important challenge is to develop electrolytes with good electrochemical properties, contribute to reduce battery aging, enhance battery safety, and allow the system to be charged and discharged electrochemically [2]. Some works have therefore concentrated on the functional electrolyte designs, and different ways are devoted to prepare usable electrolytes with excellent characters or properties, such as high ionic transport ability, excellent chemical/electrochemical stability, nonflammability, low viscosity, and economic cost [9]. Based on the battery concepts described above, ionic liquids have been placed on the electrolyte or electrolyte components. Herein, we make an overview of the development of ionic liquid-based electrolyte in sodium, magnesium, and aluminum batteries, including basic characteristics, interfacial properties, and reaction mechanism. Then, conclusive remarks based on technical difficulties, strategic analysis, and economic aspects are proposed along with an outlook.
2. Basic Principles of Ionic Liquids
Ionic liquids, known as room-temperature molten salts, a unique class of material, consist of organic cations and inorganic/organic anions [10–13]. It is worth mentioning that the electrostatic forces of the ions can be weakened and shielded through the separation of the large molecular ions in an ionic liquid, thus lowering the melting points. Ionic liquids have application value in several areas such as the organic synthesis solvents, chemical sensing, bioscience, green chemistry, and the promising electrical energy storage system (Figure 1). Ionic liquids may be used as solvents in many processes and as electrolytes in electrochemical devices. The widespread investigation and utilization of ionic liquids in an electrical energy storage area are benefited from their desirable properties such as tunable polarity and ionic conductivity, ultralow volatility, good thermal stability, and low flammability [14, 15] (Figure 1). The synergy among the electrostatic, hydrogen bonding, and hydrophobic that exist in ionic liquids provides powerful and tunable solvency properties. Ionic liquids exhibit enhanced safety due to their negligible volatility and low flammability as well as improved feasibility benefited from their wide electrochemical window (ESW) and excellent thermal stability. The coexistences of electrostatic charge, aromatic groups, and alkyl segments in the ionic structure are beneficial to the enhanced affinity of ionic liquids for inorganic and organic materials, which facilitates the preparation of composite cell material, improves interfacial properties, and produces ordered nanostructures [2].
Generally, since an alkyl group is present around the charging center, the cation tends to be large and bulky, and this is subjected to a certain degree of charge protection. Anions generally provide a conformational isomer with close energy and undergo charge delocalization due to the possibility of several resonant structures. We classify some common cations and anions used in batteries (Figure 1). Among the anions, bis(trifluoromethanesulfonyl)imide (TFSI) and bis(oxirosulfonyl)imide (FSI) are probably the most popular along with their high ionic conductivity and stability and good solid electrolyte interface- (SEI-) forming ability. The coordination and combination of ions dominate the ionic liquid characteristics, so ionic liquid can be customized according to the specific application [5]. For example, depending on their composition, ionic liquids have different classes, essentially including aprotic (suitable for lithium batteries and supercapacitors), protic (suitable for fuel cells), and zwitterionic (suitable for ionic-liquid-based membranes) types, each suitable for a particular application (Figure 1). The investigations of various commercial ionic liquids clearly demonstrate the versatility of such material and only represents a small fraction of the theoretically possible numbers of ionic liquids.
3. Advantages of Ionic Liquids as Electrolytes
Electrolyte materials are particularly crucial for the safety and electrochemical performances of advanced battery technologies. The desirable electrolytes are expected to possess several abovementioned outstanding characters. Currently, in most commercial LIBs, the standard electrolyte used is still a mixture of cyclic and linear carbonates, for example, ethylene carbonate (EC)/dimethyl carbonate (DMC) coupled with lithium hexafluorophosphate (LiPF6) [16, 17]. These organic liquid electrolytes (OLEs) generally show high ionic conductivity and good solid electrolyte interphase- (SEI-) forming ability. However, they suffer from significant downfalls particularly owing to their chemical stability and safety. The attractive properties of ionic liquids, especially the high electrochemical stability of some cationic and anionic types, make it suitable for application in high energy electrochemical devices such as lithium batteries. Moreover, the low volatility and flammability properties of ionic liquids can be regarded as their key advantages, thus providing the possibility of enhanced safety and stability.
Despite these advantages, ionic liquid-based battery systems still have some significant drawbacks, primarily high viscosity and cost. Compared to organic electrolytes, relatively high viscosity and thus low ionic conductivity limit the rate performance of ionic liquid-based battery systems. In addition, the current price of ionic liquids is too high, limiting its large-scale application in batteries. But strategies such as extending equipment life and economies of scale can help solve this problem.
In applications requiring long life and increased thermal stability, ionic liquid-based electrolytes are still hoping to overtake OLEs. The performance of the LIB prototype indicates that batteries based on ionic liquid-based electrolytes exhibit good cycle stability and capacity at low rates. In addition, ionic liquids as well as the assembled LIB can adapt to extreme conditions, such as ultrahigh vacuum spaces without heavy reinforced casings, which benefited from the wide available temperature range and low volatility of ionic liquids [5].
In recent years, post-lithium-ion battery technologies have attracted much attention, leading to many different approaches to exploring suitable electrolyte problems. The emerging development of ionic liquid-based electrolytes in aluminum, magnesium, and sodium battery chemistries is worthy to be explored and discussed.
4. Ionic Liquids as Electrolytes for Aluminum Chemistry
According to its lightweight and the three-electron transfer electrode reaction (Al3++3e-↔Al), as summarized in Figure 2(a), aluminum metal not only possesses the highest volumetric capacity (8040 mAh g-1) but also the gravimetric capacity (2980 mAh g-1) is comparable to that of lithium metal. Additionally, aluminum metal is honorably certified as the most abundant metal resource in the heart’s crust, thus leading to its fairly low price. More importantly, aluminum metal can be exposed to the air, making cell fabrication easier and more convenient and extremely improving the safety of electrochemical storage systems.
4.1. Ionic Liquid-Based Electrolytes for AIBs
For LIBs, organic solvent-based electrolytes are widely used. However, it does not apply for the AIB system. Since aluminum ions carry three positive charges and possess high surface charge density, the Coulombic interaction between cation and anion is enhanced, thereby significantly decreasing the solubility of the Al ion salt in the organic solvent, which is not conducive to the normal ionic transport of the electrolyte. The ether solvents can increase the solubility of Al ion salt with a cation-ether complex under the strong ion-dipole interaction. However, the incidental high electrodeposition polarization can be caused owing to the strong desolvation activation energy [19]. In addition, the common fluorine-containing aluminum salt easily forms a passive layer on the surface of an aluminum anode, thereby inhibiting the reversible electrostripping/deposition in an organic solvent [20, 21]. These problems make it difficult to achieve a viable electrolyte design. Therefore, conventional electrolyte selection strategies used in monovalent battery systems are not suitable for aluminum battery systems.
In fact, at the early stage, an aqueous electrolyte is firstly applied in aluminum battery systems (Al/AgO [22], Al/H2O2 [23], and Al/S [24]) due to its low cost, easy to operate, and environment friendly. However, the fatal drawbacks such as passive oxide film formation, hydrogen side reactions, and anode corrosion impede the applications of aluminum battery systems [25]. In order to ensure the reversible plating/stripping of Al, high-temperature molten salts (>100°C, e.g., NaCl-AlCl3) are utilized as electrolyte owing to the fast ion transport under the high temperature [26, 27]. In 1972, a molten salt electrolyte was applied in Al/Cl2 battery and the carbon cathode was used as the gas diffusion electrode to ensure the solid-liquid-gas interface reaction [28]. Then, in 1980, a high-temperature molten salt electrolyte was introduced to rechargeable battery systems, in which sulfides (such as FeS2 and Ni3S2) were used as cathode materials [29, 30]. The system can be known as the prototypes of AIBs. Unfortunately, high-temperature molten salt batteries require extreme conditions, accompanied by evolution of Cl2 and disintegration of cathodes, thereby limiting the development of aluminum secondary batteries [31].
Therefore, to make the rechargeable aluminum batteries practical, the design of a suitable electrolyte is crucial. Fortunately, ionic liquids have gained our attention. The attractive properties of ionic liquids, especially the high electrochemical stability of some cationic and anionic types, make it suitable for application in high energy electrochemical devices. The widespread investigation and utilization of ionic liquids in an electrical energy storage area are benefited from their desirable properties such as tunable polarity and ionic conductivity, ultralow volatility, good thermal stability, and low flammability. Moreover, the low volatility and flammability properties of ionic liquids can be regarded as their key advantages, thus providing the possibility of enhanced safety and stability. In addition, the ion-solvent complex formation with low stability promotes the rapid solvation/desolvation of aluminum ion resulting in superior transport properties. The coexistences of electrostatic charge, aromatic groups, and alkyl segments in the ionic structure are beneficial to the enhanced affinity of ionic liquids for inorganic and organic materials, which facilitates the preparation of composite cell material, improves interfacial properties, and produces ordered nanostructures. Thus, the ionic liquid-based electrolyte enables highly reversible stripping/plating of aluminum leading to the increasingly competitive AIB application in secondary batteries [32].
Instead of sodium and potassium cations in high-temperature molten salt, ionic liquid usually concludes long-chain organic substances as cations to decrease melting point. The organic cations can be imidazolium or pyrrolidinium, in particular imidazolium cations bearing various alkyl side chains, such as 1-ethyl-3-methylimidazolium (EMIm) or 1-butyl-3-methylimidazolium (BMIm). The anions can be halogens, like Cl- (most used in AIBs), Br-, I-, or organic anions. Then, the general ionic liquid electrolytes for AIBs can be obtained by mixing the ionic liquid and AlCl3 [33–38].
A detailed investigation on the various molar ratio AlCl3/[BMIM]Cl has been reported by our group in 2015 [18]. Notably, the molar ratio between the AlCl3 and ionic liquid is the dominant factor governing the electrochemical performance and the whole design of the cell. As the molar ratio of AlCl3 increases, the anions in the ionic liquid electrolytes also undergo a changing process of Cl-→AlCl4-→Al2Cl7-→Al3Cl10-. It is worth noting that only when the molar ratio of AlCl3 is larger than 1, that is, the ionic liquid electrolyte contains Al2Cl7- anion, the aluminum metal can be reversibly electrostripping/deposition. Because the Al2Cl7- anion possesses an asymmetry structure leading to its high electrochemical activity. As shown in Figure 2(b), the different anions in an ionic liquid electrolyte can also cause the different anode limiting potentials owing to the various corresponding anodic limiting reactions as follows:
Equations (1), (2), and (3) represent the oxidation reactions with a molar ratio equal, higher, or less than 1 : 1. As for cathode limiting reactions, a pair of redox peaks attributed to reversible Al stripping/deposition appears around 0 V when the molar ratio is higher than 1.5 : 1, which reinforces the importance of Al2Cl7- anion. The ionic conductivities of various molar ratio AlCl3/[BMIM]Cl ionic liquid electrolytes are investigated as shown in Figures 2(c) and 2(d). At room temperature, all the ionic liquid electrolytes exhibit the ionic conductivities of 10-3 S cm-1, which are adequate for the application in AIBs. The conductivities reach the highest when the AlCl3/[BMIM]Cl molar ratio is 1 : 1 and then decrease with increasing molar ratio [18].
4.2. Ionic Liquid-Based Electrolyte/Electrode Interface Interactions
4.2.1. Ionic Liquid-Based Electrolyte/Cathode Interface Interactions
In an aluminum-ion battery based on chloroaluminate ionic liquid electrolyte, the interfacial reaction between cathode and electrolyte contains three processes, namely: (1)the dissociation process of polyanions near the surface of the cathode(2)the migration of dissociated polyanions from an electrolyte to the surface of the cathode(3)the cathode accepts electrons to form an external circuit
Usually, these three processes are collectively referred to as a charge transfer process. In 2018, Yang et al. studied the effect of dissociation activation energy on the entire electrochemical process in aluminum-sulfur batteries (Figure 3(a)) [39]. In fact, in all multistep electrochemical reactions, the reaction process with the highest activation energy is the rate-determining step (Figure 3(b)). The ease or complexity of the dissociation reaction of anions, as a rate-determining step, does significantly affect the kinetic reaction rate due to the activation energy, which would intuitively determine the charge/discharge performance [37]. The researchers chose two different structures of anions and calculated the activation energy barriers of the two anions. As shown in Figures 3(c)–3(f), the results show that the dissociation activation energy per mole of brominated polyanion Al2Cl6Br- is 71 meV lower than that of Al2Cl7-. According to the Arrhenius formula, the dissociation reaction rate of Al2Cl6Br- is also 15 times faster than that of Al2Cl7-. This result is also verified by the Tafel equation. In the experiment, the electrolyte system with the anion of Al2Cl6Br- has a higher specific capacity. This work helps us understand the effect of the dissociation process of polyanions on electrochemical kinetics.
4.2.2. Ionic Liquid-Based Electrolyte/Anode Interface Interactions
The surge of AIBs has attracted research into Al metal anode, although cation-type AIBs do not face the urgent need for anode engineering because they have not been able to survive for long time cycles. Several works showed that the anode reaction significantly affected the electrochemical performance of the AIB [40–43]. Usually, there is a dense and Al-ion nonconductive oxide film (Al2O3) on the surface of the aluminum metal, which controls the reactions occurred on the interface between Al metal and electrolyte. In 2015, our group reported that a suitable Al2Cl7- concentration would cause slight pitting on the Al metal anode, which helped to properly remove the oxide film on the Al metal anode and improve charge/discharge electrochemical performance [18]. Subsequent in our another study, a noncorrosive and moisture-stable [BMIm]OTF-Al(OTF)3 electrolyte was performed in place of the corrosive [EMIm]Cl-AlCl3 electrolyte, resulting in no electrostripping/deposition activity [44]. However, when the Al metal anode is pretreated in a corrosive [BMIm]Cl-AlCl3 electrolyte, it exhibits electrochemical activity (Figures 4(b) and 4(c)). In fact, the oxide film naturally formed on the surface of the aluminum is so dense that the ionic liquid-based electrolyte is impermeable. However, corrosive Al2Cl7- corrodes and creates cracks in the oxide film, resulting in contact between the electrolyte and the active aluminum, thereby achieving electrostripping/deposition activity, as shown in Figure 4(a). This conclusion is consistent with our previous research.
However, the crystalline structure of the oxide film or the distinct property compared to bulk Al is uncertain. The surface structure of Al2O3 is highly defective (oxygen vacancies) and can be modified while immersed in a solution [45]. Thus, the thickness and subtle structure of the oxide film depend on the solution species, greatly affecting the interfacial reaction kinetics [46]. Choi et al. studied the electrochemical properties of Al metal anodes with two different thickness oxide films. The results show that the thinner electropolished aluminum foil surface is more susceptible to erosion by aluminum chloride complexes and shows higher and more stable capacity. This is due to its greater electrochemical reaction area and better contact with active aluminum ions [41]. They also presuppose surface reconstruction theory to explain the important role of oxide film at the anode/electrolyte interface [43]. Surprisingly, the reconstituted Al2O3 oxide film no longer blocks the migration of aluminum ion, which will be further explained in the future work. Here, the solid electrolyte interface (SEI) becomes the key factor in the anode reaction and more detailed investigations about its composition and structure need to be further conducted.
Dendrite formation is considered as the most adverse problem for metal anodes, which can produce a short circuit, resulting in the capacity fade and dead cell. However, there are few reports focused on aluminum dendrite formation. Recently, Chen et al. confirmed the existence of aluminum dendrite and proposed that the natural aluminum oxide film on the surface of aluminum anode in AIB can effectively inhibit the growth of aluminum dendrite and stabilize the interface between anode and electrolyte [40]. As shown in Figure 4(d), by limiting the aluminum dendrites below the aluminum oxide protective layer, it is possible to effectively prevent the aluminum anode from disintegrating during the repeated aluminum plating/stripping reaction, thereby obtaining superior cycling performance and an almost complete and flat aluminum foil after cycling (Figure 4(e)).
4.3. Reaction Mechanisms in Ionic Liquid-Based Electrolytes for AIBs
Various materials are performed to realize practical aluminum storage systems based on ionic liquid electrolyte such as carbon materials [47–62], oxide (transition metal oxide) [63–68], transition metal sulfides [69–76], selenides [77–79], conducting polymers [80], Prussian blue analogs [81–83], and sulfur [39, 84–87]. The reaction mechanisms of electron transfer with different types in AIBs based on an ionic liquid-based electrolyte are discussed below. In general, there may be two reversible energy storage mechanisms: intercalation and conversion reactions. Among them, according to the intercalated ions, the intercalation reaction can be further divided into polyanion and Al-ion intercalation.
4.3.1. Intercalation Reactions of Polyanions in Ionic Liquid-Based Electrolyte
The intercalation reaction mechanism in the ionic liquid electrolyte-based AIB means that the movable guest ions (Al-containing cations or polyanions) are reversibly intercalated into the layered host crystal lattice. Moreover, the electrolyte also determines which species can be intercalated into the positive electrode. Instead of intercalating Al3+ ions, the polyanion AlCl4- complex present in the ionic liquid was intercalated into the positive electrode. The intercalation reaction of polyanion is important and it usually exhibits superior electrochemical performance by assembling an aluminum battery. Depending on the geometry of the intercalation material, the polyanion intercalation reaction is carried out by structural flexibility and free adjustment of the spacing of the host lattice layers. Thus, during the charge/discharge process, this kind of reaction is along with a series of reversible structural changes in the host lattice [25]. Although the anion type AIBs exhibit superior electrochemical performance, their monovalent chemistry undoubtedly hampers the high capacity aluminum battery system design, and most of the research materials only show relatively low capacity. Moreover, large sizes of polyanions cause large volume expansion, which can result in irreversible damage to the battery.
The most commonly used ionic liquid-based electrolyte in AIBs is [EMIm]Cl–AlCl3, which is formed by mixing AlCl3 and [EMIm]Cl with a molar ratio of 1.3 : 1, where the ions present are [EMIm]+, AlCl4-, and Al2Cl7-. Obviously, there is no isolated Al3+, so if the insertion reaction occurs, AlCl4- is the most likely reagent because of its smallest ionic radius [88], which results in a lower steric hindrance than the other two ions.
Carbons (graphite, graphene, or amorphous carbon) have acted as intercalation host compounds with AlCl4-. Taking graphite as an example, as an attractive layered material, graphite has the ability to carry anions besides cations (an acceptor or donor of electrons) [89]. Interestingly, graphite materials can introduce another type of intercalation electrochemistry in the ionic liquid electrolyte-based AIBs. In a conventional AIB system using ionic liquid as an electrolyte, when a graphite material is used as the cathode, the polyanion AlCl4- in the electrolyte is inserted into the host structure, and electron holes are generated in the bonded π band, thereby forming an acceptor type graphite intercalation compound (GIC). In addition to carbon materials, conductive polymers are also considered to deintercalate polyanions during charge and discharge. The cathode and anode reactions in carbon materials can be expressed as follows:
4.3.2. Intercalation Reactions of Al-Ions in an Ionic Liquid-Based Electrolyte
Except for polyanions, Al-ions can also be deintercalated by some other certain types of materials. Due to the deintercalation of trivalent aluminum ions in the cathode materials, cation type AIBs theoretically exhibit higher specific capacity, making them attractive for researchers. Recently, vanadium oxide (V2O5 [65]), Chevrel phase (Mo6S8 [76]), cubic sulfide (CuTi2S4 [90]), and layered disulfide (TiS2 [90]) were shown to make reversible deintercalation of aluminum ions. These cathode materials usually feature with large interstitial sites and diffusion path for ion insertion. The proposed intercalation reactions can be summarized as follows:
where M represents the cathode materials.
The schematic illustration of a possible cathode reaction process in the ionic liquid-based electrolyte is proposed in Figure 5. During the charging process of the graphite cathode material, the generation of chlorine gas is also exhibited. As reported by Shi et al., chlorine generation occurs at about 2.4 V versus Al, and in most publications on carbon-based cathode materials, the charge cut-off potential is higher or very close to 2.4 V versus Al. Therefore, it is a reasonable problem that chlorine may be generated during charging on the carbon-based cathode [91]. In fact, in the early study of the use of graphite as a cathode material for rechargeable Al batteries [92], proposals have been made to use chlorine as an intercalation material in graphite. Yang et al. also recently reported the incorporation of chlorine into graphite from chloride-containing electrolytes [93]. Considering that the potential of chlorine generation is closely related to the charge potential, irreversible charge capacity, and severe self-discharge behavior, it is reasonable to assume that chlorine is involved in the intercalation and may not be stably hosted.
4.4. Challenges of Ionic Liquid-Based Electrolytes for AIBs
Although ionic liquid-based electrolytes show many possibilities in aluminum battery system applications, there are still some problems remaining (Figure 6(a)).
The first obstacle is the reversible/irreversible side reactions that occur on the anodic and cathodic sides, which can seriously affect the battery charge-discharge behaviors.
On the cathode side, it has been reported that the electrophilic element of the VI A group and the nucleophilic C2 carbon in the imidazole ring in the cation can react with each other, thereby causing a side reaction of the cathodic side [94]. The reactions are present in most aluminum-sulfur batteries [21]. In addition, the reduction of the chloroaluminate-based ionic liquid on the cathode surface may be another nonnegligible potential side reaction. Recently, some studies reported that in the three-electrode system, the reduction potential of the electrolyte on the glassy carbon working electrode reached 0 V (vs. Al/AlCl4-) [18]. However, during the discharge process, the reduction potential of the electrolyte can be increased due to the catalytic activity of the transition metal oxide/sulfide cathode materials to some extent [95, 96]. That is to say, the side reaction of electrolyte reduction may occur during the discharge process. To date, although no direct evidence confirms that the catalytic activity of the cathodes can contribute to the reduction of electrolyte, the “excess capacity” phenomena have been explored, which may be due to the reduction of electrolyte caused by the cut-off discharge voltage less than 0.1 V [74, 79].
On the anode side, firstly, during the charge and discharge process, side reactions based on chloroaluminate ionic liquid electrolyte are prone to occur. For example, AlCl4- will be oxidized by to form toxic Cl2, resulting in relatively low Coulombic efficiency for AIBs [18, 97–98]. Secondly, since the chloroaluminate-based ionic liquid electrolyte is extremely corrosive, the cell shells can be eroded, thereby causing side reactions [99, 100]. In 2013, the electrochemical stability of the former electrolyte was reported by Reed and Menke [99]. Interestingly, even if there is no cathode material in the battery, the charge and discharge platforms of the corresponding redox peaks can still be seen. Further analysis indicates that this reversible capacity resulted from the redox reaction of Fe and Cr with Al2Cl7-. After that, the stainless steel is to be avoided in AIBs, while the shells are made of polytetrafluoroethylene (PTFE) or pouch cells instead. It has also been suggested to use the expensive molybdenum instead of stainless steel [60, 72, 101, 102]. Thirdly, the selection of current collector is critical to the assembly of the battery, and some usual current collectors can react with the ionic liquid-based electrolyte to produce side reactions. Shi et al. [91] studied the electrochemical stabilities of different current collectors in the electrolyte during charge and discharge. The galvanostatic discharge-charge (reduction-oxidation) curves of various conductive substrates are shown in Figures 6(b)–6(g). The electrochemical oxidation of nickel (Ni) results in a distinct discharge platform at 0.8 V (vs. Al/Al3+) with high apparent capacity and repeatability in subsequent discharges. It is obvious that Ni and Ni-containing alloys are not suitable for use as current collectors for cathode materials. The same applies to titanium (Ti), platinum (Pt), and tungsten (W). However, molybdenum (Mo) and glass carbon (GC) can be stably present in the ionic liquid-based electrolyte without side reactions. Thus, it can be concluded that Mo and GC can be considered as the better current collectors of the cathode materials.
The second obstacle is the compatibility of ionic liquid electrolyte with cathode material and binder. Firstly, the compatibility of cathode material (especially the transition metal oxide/sulfide) with the ionic liquid electrolyte is almost ignored in the current literature reports. But a recent study showed that vanadium pentoxide (V2O5), one of the most common cathode materials for AIBs, had been certified to dissolve in Lewis acid ionic liquid electrolyte and form new vanadium compounds [91]. In addition, iron sulfide (FeS2) is also reported to be soluble in the electrolyte [74]. Based on the above results, we can speculate that there may be the same cases for other cathode materials reported in the literature. Thereby, in future research, the compatibility between cathode material and electrolyte should be tested in advance. Secondly, our group reported that polymeric binders such as polyvinylidene fluoride (PVDF) were incompatible with chloroaluminate-based ionic liquid electrolyte [18]. The PVDF power is slowly added to the AlCl3/BMIMCl, and the color of the electrolyte changes accordingly, which may be attributed to the reaction of PVDF with Al2Cl7- in the electrolyte.
Finally, the chloroaluminate-based ionic liquid electrolytes are highly hygroscopic, viscous, and expensive, all of which exacerbate their serious limitations as high-efficiency battery electrolytes [103–106]. These shortcomings may hinder the design of high-performance AIBs and suppress the commercial applications of ionic liquid-based electrolytes for AIBs.
4.5. Future Perspective
As mentioned earlier, the side reactions of the Al battery and the compatibility of the electrolyte with the cathode material and the binder are all caused by the unique corrosion of the ionic liquid electrolyte [107]. In order to inhibit the corrosive nature of the electrolyte, the corrosive chloride ions can be replaced by other ions, such as triflate (OTF-) and bromide (Br-). Wang et al. used 1-butyl-3-methylimidazolium trifluoromethanesulfonate ([BMIm]OTF) mixing with Al(OTF)3 to obtain a noncorrosive ionic liquid with water stability [44]. After the oxide film is removed by preimmersing the Al anode in the ionic liquids, a noncorrosive electrolyte can also be used in the RAB. Through this research, a new method is proposed. A corrosive AlCl3-based electrolyte is firstly used to create a suitable channel for Al3+, and then, a noncorrosive Al(OTF)3 based electrolyte is used to obtain the stable Al/electrolyte interface. In addition, Br- has also been used to replace Cl- to inhibit the corrosion reactions [39]. Surprisingly, the use of Br- also accelerates the kinetics of the electrochemical process of aluminum-sulfur batteries. Al-S cells have poor reversibility in chloroaluminate-based ionic liquid electrolytes, mainly due to the slow kinetics determined by the decomposition of Al2Cl7- to Al3+. Density functional theory (DFT) calculations show that the dissociation reaction of Al2Cl6Br- is easier than that of Al2Cl7- because the dissociation energy barrier of Al2Cl6Br- is lower. Therefore, a 15 times accelerated dissociation rate can be achieved by replacing all Al2Cl7- with Al2Cl6Br- in the electrolyte. The N-butyl-N-methyl-piperidine (NBMP+) cations instead of EMI+ also avoid side reactions between EMI+ and polysulfides. Therefore, the accelerated Al-S battery can be successfully achieved by using NBMPBr/AlCl3 electrolyte. As for the corrosion of the shell by the electrolyte, a new coin-cell-based AIB configuration has been prepared by applying conductive polymer PEDOT coating to prevent corrosion [54].
In addition, we know that in the AlCl3-containing imidazole-based ionic liquid, the dominant Al2Cl7- may be the only active species in aluminum deposition, and at a certain concentration, the passivation film can be appropriately removed, whereas the relative concentration of Al2Cl7− is susceptible to dynamic equilibrium (e.g., [108], K≈10-13), which promotes the conversion of Al2Cl7− to AlCl4−. In turn, the unique behaviors can cause the anodic reaction to be filled with charge and mass transfer entangled with the species. How to control the species is the key to maintaining the anode reaction cycle life. The electrolyte additive toluene is well utilized with the aim to improve the aluminum deposition kinetics [109, 110].
Since the commonly used AlCl3-containing imidazole-based ionic liquid is expensive, corrosive, sensitive to water and air, and features with a narrow electrochemical window, it possesses several difficulties in commercial application for AIBs. Therefore, novel safe, nontoxic, and stable alternative electrolytes are urgently needed and this electrolyte must be able to achieve a fast and efficient aluminum stripping/deposition reaction. In addition, a suitable electrolyte must have appropriate corrosive properties, which can dissolve the Al2O3 passivation film on the surface of the aluminum anode, making it electrochemically active, and must also avoid the destructive corrosion for the positive electrode material, current collector, and battery case, thereby suppressing the occurrence of side reactions. The enhancement and advancement of the electrolyte system will contribute a lot to the advanced aluminum battery developments.
5. Ionic Liquids as Electrolytes for Magnesium Chemistry
Magnesium is regarded as a natural choice as an anode material for rechargeable batteries due to its thermodynamic properties. Generally, magnesium has a higher volume energy density compared to other metal (such as Li); this allows the same energy to be stored in smaller batteries, which can be achieved by using magnesium metal anode [5]. Although Mg has a lower gravimetric density than Li (2205 vs. 3860 mAh g-1), it has a considerably higher volumetric density (3833 vs. 2050 mAh L-1). In addition, compared to lithium, magnesium is cheaper, environmentally friendly, and higher safety. Currently, the development of magnesium batteries is hampered by two issues [111]. First, due to the chemical activity of magnesium, the ideal electrolytes for use should provide neither protons nor accept protons, thereby inhibiting the electrochemical reaction with the passivated surface layer formation [112–114]. Second, the difficulty of Mg ion insertion in many hosts as well as the nature of the electrolyte limits the choice of cathode material [115]. So the research and process of ideal magnesium electrolytes determine the pace of development of the field.
5.1. Conventional Electrolytes for MIBs
For lithium batteries, we usually dissolve simple salts and anions in carbonate/aprotic solvents to prepare electrolyte solutions, where Li can be reversibly plated and stripped [116]. However, the nature of the mesophase formed at the electrolyte/metal anode interface is quite different, which is one of the most significant differences between the lithium and magnesium battery chemistry. In both cases, the decomposition products will form a layer on the electrode once electrolyte reduction occurs. Generally, the solid electrolyte interface (SEI) on lithium metal contributes to lithium-ion conduction. Unfortunately, this is not the case for magnesium metal. The layer formed on the magnesium metal is passivated, which prevents the transfer of magnesium ions to the electrode [117–119]. Therefore, electrolyte analogs of lithium batteries are generally not considered suitable for magnesium-based systems. Instead of taking advantage of the natural reduction product to prevent further reactions on the metal electrode (like lithium), the ideal electrolyte should possess reduction stability to avoid the unstable interphase layer formation.
The most widely known electrolytes capable of reversible plating/stripping of magnesium ions are based on Grignard reagents or Lewis acid-base pairs dissolved in ether solvents, such as (Mg2 (μ-Cl)3·6THF) (HMDSAlCl3) and (Mg2 (μ-Cl)3·6THF) (Ph2AlCl2) (Figure 7(a)). For example, as shown in Figure 7(b), magnesium ions can be successfully reversibly plated and stripped from magnesium organohaloaluminates or magnesium organoborate salts containing (Mg2 (μ-Cl)3·6THF)+ cation dissolved in etheric solvents [120, 121]. The dielectric constant of ether is very low, which limits the magnesium salt solubility. In addition, these electrolytes are confronted with safety risks due to the unstable volatile solvents and salts in the air and/or water. It is expected that the battery electrolyte contains a high concentration of an electroactive salt, thus bringing out fast ion transport [122]. And the desirable electrolyte may as well possess poor volatility and flammability to improve the safety of the battery system. Thereby, ionic liquids are considered as potential solvents for various magnesium salts.
5.2. Ionic Liquid-Based Electrolytes for MIBs
Ionic liquid can be used not only as a source of cations/anions but also as a nonflammable solvent that dissolves salt cations. In the case of magnesium electrolytes, this advantage is particularly convenient because simple salts of magnesium and conventional solvents (alkyl carbonates, esters, and acetonitrile) can form a passivate layer on the surface of magnesium metal. However, it must be taken seriously whether the ionic liquid has sufficient reducing stability to magnesium.
Research on ionic liquids for magnesium electrochemistry is still on its early stages. A lot of efforts has been made on building efficient magnesium primary batteries, but the development of rechargeable magnesium batteries is still a long way off. The ionic liquid electrolyte containing BF4- anion features with high ionic conductivity and low melting point. A mixture of magnesium triflate (Mg(CF3SO3)2) and 1-n-butyl-3-methylimidazole tetrafluoroborate (BMIMBF4) was described, which can be considered as one of the earlier reports on magnesium ionic liquid electrolytes [123]. Experimental results show that BMIMBF4 exhibits stability on a platinum working electrode at -1 to 3 V relative to Mg and the plating/stripping process of magnesium with Mg(CF3SO3)2 salt occurs at low current, as shown in Figure 7(c). Surprisingly, even with a reported Coulombic efficiency of 100%, the plating/stripping process cannot be repeated for long cycles along with the strong voltage fluctuations. According to EDS analysis, NuLi determines S, F, and Mg in the sediment when depositing at a low charge density. It concludes that even if the CF3SO3- is reduced to form a film, magnesium will still be deposited from the electrolyte system [123]. In Figure 7(d), NuLi also reports the reversible deposition of magnesium from a 1 M solution of 1 M Mg(CF3SO3)2 in N-methyl-N-propylpiperidine bis(trifluoromethane)-sulfonimide (PP13TFSI) on a silver working electrode [124]. This is due to the Ag-Mg clusters initially formed, promoting subsequent Mg deposition. Compared with BMIMBF4, the TFSI- anion has a wider electrochemical window. Subsequent reports describe the use of a mixed ionic liquid made of BMIMBF4 and PP13TFSI, and the same salt (Mg(CF3SO3)2)- [125]. Although the aim to use a mixed ionic liquid is to increase ionic conductivity and cathode stability, they also observe that the mixture provides a lower initial deposition/dissolution overpotential than each single ionic liquid system.
However, the most common ionic liquids include anions (such as TFSI), which are usually reductively unstable with magnesium metal. In fact, the TFSI anions in ionic liquids easily form a protective layer on Mg metal-containing Mg and TFSI decomposition products [126]. Early studies mentioned above claimed that Mg can be reversibly plating/stripping from ionic liquid electrolytes containing ionic magnesium salts [123]. But the results can not be copied [127, 128]. Thus, the satisfactory ionic liquid electrolytes for magnesium-ion batteries require not only wide electrochemical window, high salts solubility, and high ionic conductivity but also interface compatibility and stability. The structure, composition, and ratio of ionic liquid electrolyte were designed to optimize the magnesium/electrolyte interface, so as to find an ideal ionic liquid electrolyte that can effectively deposit and dissolve magnesium reversibly at room temperature. The anode material matched with the electrolytic liquid phase was developed, and the energy was stored and released by the breaking and bonding of the chemical bonds, so as to realize the effective charging and discharging cycle of the battery. It is expected that this basic work will lay a foundation for the in-depth research and extensive application of high-performance rechargeable magnesium batteries.
5.3. Ionic Liquid-Based Polymer Electrolytes
The development of ionic liquid-based polymer electrolytes for magnesium systems has also attracted researchers’ attention. The electrolyte can reversibly transport Mg2+ ions, in which various Mg salts are coupled with the polymer chain, plasticizer, and/or filler for enhanced ionic conductivity [129].
Morita et al. [130–132] synthesize a solution consisting of poly-(ethylene oxide) PEO-PMA (polymethacrylate) matrix and 1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide (EMITFSI) ionic liquid dissolving Mg(TFSI)2. Then, the ionic conduction as a solid-state electrolyte for magnesium polymer batteries has been studied. They observe that ionic conductivity depends on the content of ionic liquids which contain Mg(TFSI)2 and the highest conductivity of the polymer gel is more than 10-4 S cm-1. Yoshimoto et al. [133] further report a flexible, self-supporting (medium mechanical strength) transparent polymer electrolyte based on ionic liquids, which contains three components: PEO-PMA polymer matrix, Mg(TFSI)2, and ionic liquid EMITFSI or DEMETFSI. The authors successfully prepare a gel polymer electrolyte using poly(ethylene glycol) monomethacrylate (PEM) and poly(ethylene glycol) diethylacrylate (PED) as prepolymers. An appropriate amount of Mg(TFSI)2 is dissolved in EMITFSI or DEMETFSI at room temperature. In a Mg(TFSI)2/EMITSFI or Mg(TFSI)2/DEMETFSI ionic liquid solution, add a 3 : 1 molar mixture of PEM and PED, which contains a small amount of a photo-induced free radical initiator, and add 2-phenylbenzene acetone. The liquid is cured to an aluminum plate by UV radiation in an Ar atmosphere, and finally, a crosslinked polymer matrix PEO-PMA with salts in an ionic liquid is produced. The ionic conductivity of these polymer electrolytes varies with the content of ionic liquids containing Mg(TFSI)2.
5.4. General Strategies to Develop Ionic Liquid-Based Magnesium Electrolytes
Although ionic liquids can dissolve magnesium salts at high concentrations, they are troubled by the reduced ionic conductivity and electroactive species migration numbers. This is due to its high viscosity, and the additional ions as well as Mg can also migrate when applying an electric field. The addition of tetrahydrofuran (THF) as a cosolvent helps to increase the ionic conductivity with the decreased viscosity. Particularly, Cheek et al. point out that Mg can be reversibly plated/stripped in the 1 M PhMgBr in THF/1-butyl-1-methylpyrrolidinium trifluoromethanesulfonate (BMPCF3SO3) [134]. Yoshimoto’s group report that 1.0 M DEMETFSI has an ionic conductivity of 7.44 mS cm-1 [135]. The addition of ether into ionic liquid-based electrolytes makes a great significant increase of ionic conductivity and a decrease in viscosity (in both cases an order of magnitude reduction) [136]. The results of spectra [137, 138] and single crystal [139] indicate that the charged Mg substance in the TFSI-based ionic liquid-based electrolyte may be a large anionic complex. On the contrary, when oligomeric ether is added, it shows that Mg is mainly coordinated by ether oxygen. The combined electrochemical and spectroscopy results indicate that the nature of Mg’s charge-carrying substance is a crucial ingredient affecting the deposition/dissolution process.
Optimizing the cations in the imidazolium-based ionic liquid using TFSI-anions can further increase the current density [140, 141]. Kakibe et al. prove the magnesium cycle by replacing the alkyl group on the cation (such as [C2mim]+) to avoid undesired reaction between the magnesium surface and the C2 proton on the imidazole ring [141]. Interestingly, it also shows that the varied ratio of the TFSI-/FSI- anion in a binary ionic liquid mixture can also make a raise in current density. Gao et al. point out that the speciation behavior of Mg2+ is crucial for assessing the feasibility of ionic liquid electrolytes in Mg-based batteries [116, 142]. The authors investigate two types of alkoxy-functionalized ionic liquids (one with alkoxy-pyrrolidinium cations and another with alkoxy-ammonium cations) and report better reversibility than the N-butyl-N-methylpyrrolidinium bis(trifluoromethanesulfonyl)imide (Pyr14TFSI) usually studied. Huie et al. investigate ionic liquids with different cations (pyridine, imidazolium, piperidinium, and pyrrolidinium) in a solution containing Mg (TFSI)2 salt and DPGDME or an acetonitrile cosolvent [143]. A correlation between the high conductivity of ionic liquid with unsaturated rings and short carbon chain lengths can be observed. However, these ionic liquids also exhibit lower oxidative stability. A composition of Mg(TFSI)2 dissolved in 40% 1-ethyl-1-methyl-pyrrolidinium bis(trifluoromethanesulfonyl)imide (EMPTFSI) and 60% acetonitrile exhibits a conductivity of 40 mS cm-1, and the best performance is obtained, with voltage stability of 5.23 V relative to Mg0/Mg2+.
In addition to TFSI anion, Mg(BH4)2 has also been proven to be a salt capable of Mg deposition, and there have been reports of using Mg(BH4)2 as the source of Mg in ionic liquid electrolytes [144, 145]. And Kar et al. verify the results in an alkoxyammonium-based ionic liquid named [N2 (20201) (20201) (20201)] [NTf2] [144]. Su et al. develop a piperidinium-based ionic liquid-based electrolyte, namely, 1-butyl-1-methylpiperidine bis(trifluoromethanesulfonyl)imide (PP14TFSI). The dissolved magnesium salt is Mg(BH4)2. The ionic conductivity of the electrolyte ranges from 1 to 3 mS cm-1 [145].
The ionic liquids can be regarded as additives to add into magnesium liquid electrolytes, which have made valuable progress [145–147]. Cho et al. find that the addition of 1-allyl-1-methylpyrrolidinium to the Grignard reagent solution can improve its oxidation stability by 1.0 V [146]. Pan et al. develop a new electrolyte by adding DEMETFSI ionic liquid and observe that its ionic conductivity increases linearly with increasing DEMETFSI concentration [147]. Figure 8 provides all kinds of prominent ionic liquid electrolytes used in MIBs.
6. Ionic Liquids as Electrolytes for Sodium Chemistry
Lithium technology powers a lot of our modern devices and is therefore well known. The sodium chemistry shares commonalities. Because of the following reasons, SIBs are expected to meet the requirements for sustainability and performance in various applications [148]. (1) Compared with lithium, sodium has a higher abundance and a wider distribution. The abundance of Na in the earth is about 24000 ppm (20 ppm for Li), which are almost unlimited resources [149]. The price of commonly used Na-containing raw materials (Na2CO3) is therefore more than 100 times lower than Li2CO3. (2) Although sodium is heavier, subsequent capacity loss (estimated) can be compensated by replacing heavy copper current collectors (which need to be used in LIBs) with lighter aluminum and leading to greatly reduced costs compared to LIBs. (3) Na metal exhibits a low redox potential (2.71 V relative to SHE), resulting in a rather high battery voltage. (4) The ionic radius of Na is larger than that of Li, which causes reduced solvation in polar solvents. It is known that desolvation can control the migration of alkali metals at the interface, so this weaker interaction promotes the insertion of Na+ [150].
6.1. Motivations for Ionic Liquid-Based Electrolytes for SIBs
The electrolyte is an essential component of SIBs, and it transfers and balances the charges between the electrodes in the form of ions [17, 151]. The electrolyte is located between the cathode and anode and interacts closely with the electrodes. Therefore, it has been increasingly recognized that the electrolyte will profoundly affect the battery system, not just the transport medium for shuttle ions. Over the past few decades, based on such electrolyte characteristics, the researchers have made great efforts to find the best combination of suitable electrolytes for SIBs. Despite the research on SIBs in the 1980s, the renaissance of SIBs has just begun, so there is currently no perfect “standard” electrolyte. In academic and industrial research and development, carbonate-based electrolyte solutions with sodium salts are regarded as one of the most suitable electrolyte media for SIBs. However, organic electrolytes are flammable and volatile. In the case of battery abuse, the electrolyte may cause the electrolyte to burn. And the internal pressure of the battery may be enlarged by volatilization, which may cause a fire or explosion of the battery [152]. In addition, their high sensitivity to thermal history and temperature fluctuations limit their application in emerging devices [153, 154]. Therefore, it is very urgent to develop a next-generation electrolyte with high safety for SIBs, such as ionic liquid-based electrolytes.
Although ionic liquids are not economical choices due to the high cost, since the amount of electrolyte required in a battery system is very limited, the overall cost of the battery system is not significant [155]. Moreover, with the increasing research on the synthesis and recovery of ionic liquids, their costs are expected to decrease accordingly [156, 157]. It is worth noting that using an ionic liquid as an electrolyte, other functions of the battery, such as enhanced durability and performance/cost ratio, may compensate or eliminate battery cost issues. Figure 9 shows the Ragone diagram of various battery systems [148]. It can be seen from the figure that the sodium-ion battery system using an ionic liquid as the electrolyte has a good specific energy and power at room temperature. Words cannot be reached.
6.2. Cations and Anions for Ionic Liquid-Based Electrolytes
In this section, different cations and anions of ionic liquids for SIB electrolytes are summarized. It is worth noting that the usual research focuses on cations such as imidazolium, pyrrolidinium, ammonium, and anions of TFSI-, FSI-, and BF4- [158]. The corresponding chemical structures of these anions and cations are listed in Figure 10 and Table 1. In addition, the combination of cations and anions can be freely selected and meet the ideal performance requirements of ionic liquid [159].
|
Fluoro complex anions (BF4-) are widely used to obtain low melting salts previously [160–161]. But their low stability to hydrolysis and high viscosity prevent them from being used as ideal electrolytes in battery systems. Although some works prepared some more stable hydrolysis of BF4- in SIBs, based on their hydrophilicity, the synthesis of them is not economical and convenient, so it is not attractive in practical. Sulfonamide-based (also known as sulfonimide-based) ionic liquids are popular due to their ease of synthesis and high ionic conductivity. By expanding its organic framework, the cations in the ionic liquids change more than the anions. Although melting points usually decrease when large or long substituents are introduced, viscosity and ionic conductivity generally increase and decrease, respectively. Therefore, asymmetric cations having a short alkyl chain are generally preferred. Nonaromatic pyrrolidinium cations possess higher reduction stability than imidazolium cations and are therefore widely used in battery systems [162]. In the following review, the research progress and application prospects of three different cationic (imidazolium, pyrrolidinium, and ammonium) ionic liquid-based electrolytes are discussed.
6.3. Imidazolium-Based Ionic Liquid Electrolytes
In practical applications, imidazolium-based ionic liquids, such as 1-ethyl-3-methylimidazolium (EMIm-) and 1-butyl-3-methylimidazolium (BMIm-), have attracted widespread attention because they feature with relatively low viscosity, high ionic conductivity, and high chemical stability [163, 164].
As early as 2005, ionic liquids based on imidazolium cations with trihalides used as anions have been explored [165]. Studies have found that trihalide-based ionic liquids ([Bmim] [Br3], [Bmim] [ICl2], and [Bmim] [IBr2]) provide lower melting point, lower viscosity, and higher density and hydrophobicity, compared to some other common imidazolium ionic liquids, such as chloride, bromide, or iodide. It is worth noting that the size of the anions makes a great influence on the potential for forming ion aggregates. And these aggregates significantly influences the physical-chemical properties of ionic liquids, such as refractive index, electronic polarizability, ion conductivity, and viscosity. Particularly, [Bmim] [IBr2] exhibits the highest conductivity, 40 mS cm-1, while the conductivity of [Bmim] [Br3] and [Bmim] [IBr2] is lower, and the viscosity is much higher. Subsequently, in early 2010, Plashnitsa et al. have reported an imidazolium-based ionic liquid electrolyte consisting of 1-ethyl-3-methylimidazolium tetrafluoroborate EMIm-BF4 and 0.4 M NaBF4. And the ionic liquid-based electrolyte is applied to a symmetrical battery with Na3V2(PO4)3 (the cathode and anode are both Na3V2(PO4)3). This battery shows better cycle life due to the lower reactivity of the ionic liquid electrolyte compared to conventional organic electrolytes. In addition, it possesses higher thermal stability at high temperatures. Notably, as shown in Figure 11(a), the Na3V2(PO4)3 electrode coupled with the ionic liquid-based electrolyte still exhibits good thermal stability at temperatures up to 400°C [166]. Recently, Wu et al. also report EMIBF4-based ionic liquids with different concentrations of NaBF4 as nonflammable electrolytes for SIBs. Particularly, when the concentration of NaBF4 is 0.1 M, the ionic liquid-based electrolyte shows excellent ionic conductivity of S cm-1 at 20°C. And in the voltage range of 1-5 V, the ionic liquid-based electrolyte remains electrochemically stable. In addition, in Figure 11(b), the electrolyte is proven to possess high thermal stability, and thermal decomposition does not occur until 380°C [167].
Monti et al. investigate the physical and chemical properties of imidazolium-based ionic liquids (1-ethyl-3-methylimidazolium TFSI (EMITFSI)) and 1-butyl-3-methylimidazolium TFSI (BMITFSI) coupled with NaTFSI salts. Taking Na0.1EMIm0.9TFSI (equivalent to 0.4 M NaTFSI) as an example, as shown in Figure 11(c), its ionic conductivity at room temperature is 5.3 mS cm-1 and it can maintain thermal stability in the temperature range of -86 to 150°C [168]. By DFT calculation and Raman spectroscopy, it is concluded that the Na charge carrier is the [Na(TFSI)3]2- complex, which plays a major role in ionic conductivity. Then, the viscosity, diffusion coefficient, and conductivity of the ionic liquid electrolyte based on Emim-TFSI are first calculated by MD simulation [169]. Recently, Wu et al. report NaPF6/BMIm-TFSI electrolytes for SIBs. The self-extinguishing time (SET) and thermogravimetric tests are used to study the flammability and thermal stability of the electrolyte, which are shown in Figures 11(d) and 11(e) [170]. NaPF6/BMImTFSI electrolyte shows a short SET of 25 s g-1, which is much lower than that of the 1.0 M NaPF6/PC electrolyte (210 s g-1). A shorter SET indicates that ionic liquid-based electrolyte is nonflammable. In the TG measurement, the NaPF6/BMIm-TFSI electrolyte does not show significant weight loss until 350°C. In contrast, the 1.0 M NaPF6/PC electrolyte shows a 90% weight loss at 60°C. Therefore, NaPF6/BMIm-TFSI electrolyte shows good thermal stability.
Recently, an energy storage method using an ionic liquid as a combined ion conductive medium and a redox-active catholyte material has been widely concerned [171]. The schematic diagram and working mechanism of the battery are displayed in Figure 11(f). Earth-rich iron ions in the form of a low melting point NaFeCl4 are mixed with the ionic liquid ethylmethylimidazole tetrachloroaluminate as an electrolyte. The electrolyte maintains the highest ionic conductivity. Moreover, the battery with the ionic liquid electrolyte and liquid sodium anode exhibits a high voltage exceeding 3.2 V, and its cyclability is confirmed. When operating at 180°C, it shows a super high energy efficiency. This provides a new insight for ionic liquids as catholytes to further develop high energy density SIBs.
6.4. Pyrrolidinium-Based Ionic Liquid Electrolytes
Pyrrolidinium-based ionic liquid electrolytes including N-propyl N-methylpyrrolidinium (Pyr13-) and N-butyl-N-methylpyrrolidinium (Pyr14- or BMP-) are very attractive for SIB applications [172]. At present, there are mainly two kinds of pyrrolium-based ionic liquid electrolyte systems used in SIBs, namely, NaFSI/Pyr13FSI and NaTFSI/BMPTFSI.
In 2013, Ding et al. investigate the physical and chemical properties of NaFSI/Pyr13FSI ionic liquids that acted as a new electrolyte for SIBs with a wide temperature range [173]. Studies have shown that the viscosity and ionic conductivity of the ionic liquid-based electrolyte as shown in Figure 12(a). This ionic liquid-based electrolyte exhibits a wide electrochemical window, which is shown in Figure 12(b). The battery assembled by using the ionic liquid-based electrolyte with NaCrO2 cathode exhibits a specific capacity of 92 and 106 mAh g-1 at 298 and 353 K at 20 mA g-1, respectively. It is verified that NaFSI/Pyr13FSI is a promising electrolyte for SIBs operating in a wide temperature range. Then, the group then investigates the effects of Na ion concentration and operating temperature on the viscosity, ionic conductivity, and charge-discharge performance of Na/NaFSI/Pyr13FSI/NaCrO2 batteries. As depicted in Figure 12(c), the optimal concentrations are confirmed for the batteries operating [174].
Subsequently, Fukunaga et al. demonstrate the electrochemical performance of the SIB with NaFSI/Pyr13FSI as the electrolyte and hard carbon (HC) as the anode in Figure 12(d) [175]. This Na/NaFSI-Pyr13FSI/HC battery shows a reversible capacity of 260 mAh g-1 at the current density of 50 mA g-1 at 363 K and remains 95.5% after 50 cycles. The battery also shows a high rate capacity of 211 mAh g-1 at the current density of 1000 mA g-1, which is equivalent to a charging rate of about 4C. Therefore, it can be concluded that this combination of HC anode and NaFSI/Pyr13FSI electrolyte can provide a safe, high-rate performance SIB. In addition, Wang et al. report the electrochemical performance of Na0.44MnO2/HC full battery based on NaFSI-Pyr13FSI. In Figure 12(e), this battery shows a higher capacity (117 mAh g-1) and higher stability compared to conventional carbonate-based electrolyte because of the SEI layer formed by IL electrolyte on the HC electrode (Figure 12(f)) [176]. Still for the NaFSI/Pyr13FSI electrolyte system, Forsyth et al. find that high concentrations of sodium salts can increase the number of Na+ transfers in ionic liquids, thereby improving the interfacial characteristics of Na metal anode, achieving low polarization potential and long cycle life [177]. Recently, Kim et al. report a new electrolyte based on a combination of two ionic liquids (Pyr13FSI; Pyr13TFSI), sodium salts (NaTFSI), and ethylene carbonate (EC) additives for seawater batteries. Adding 5% EC can promote the formation of SEI, then stabilize the electrolyte/anode (hard carbon) interface. In addition, this battery shows better energy efficiency (voltage efficiency) and recyclability than conventional organic carbonate-based electrolytes [178].
In addition to NaFSI/Pyr13FSI, there are studies on NaTFSI/BMPTFSI electrolytes in SIBs as early as 2013. Noor et al. conclude that sodium metal can be reversibly plated/stripped from the ionic liquid-based electrolyte by cyclic voltammetry [179]. And the electrochemical properties, thermal properties, density, viscosity, and conductivity of the electrolyte have been systematically investigated. Surprisingly, the ionic conductivity is up to 8 mS cm-1. Due to the increase in the viscosity and density, their ionic conductivity slowly decreases with the increase of the salt content.
Subsequently, as depicted in Figure 13(b), the effect of NaTFSI concentration on the electrochemical performance of Na/NaTFSI-BMPTFSI/NaFePO4 batteries has been studied [180]. When the NaTFSI concentration is about 0.5 M, a compromise between the number of working ions and the electrolyte conductivity can be achieved to optimize battery capacity and high rate capacity. In addition, the battery also has excellent performance at high temperatures. At 50°C, the Na/NaTFSI-BMPTFSI/NaFePO4 battery capacity in ionic liquid mixed with 0.5 M NaTFSI is 125 mAh g-1 (at 0.05 C). Moreover, the nonflammability property of the IL electrolytes is ideal for high-safety batteries (Figure 13(a)).
Not only the concentration of the sodium salt, the composition of the sodium salt also results in different electrolyte performance. Based on the BMPTFSI electrolyte, the effects of different sodium salts (such as NaBF4, NaClO4, NaTFSI, and NaPF6) on the performance of Na/NaTFSI-BMPTFSI/Na0.44MnO2 batteries are reported [181]. From Figure 13(c), the results indicate that NaClO4 is the most suitable salt for this system. From another perspective, Na/NaTFSI-BMPTFSI/Na0.44MnO2 batteries with different Na salts (NaBF4, NaClO4, NaPF6, and NaN(CN)2) are investigated [182]. In Figures 13(d) and 13(e), NaBF4-based electrolytes exhibit the highest ionic conductivity (1.9 mS cm-1 at room temperature), the lowest viscosity, and the most stable cycle performance. Recently, the BMPTFSI-NaTFSI electrolyte is applied in a full battery (P2-Na0.6Ni0.22Fe0.11Mn0.66O2/Sb-C), and the results show a high initial capacity [183].
6.5. Ammonium-Based Ionic Liquid Electrolytes
Ammonium-based ionic liquid electrolytes have also received widespread attention for their high solubility in Na salts. Generally, the increase in sodium salt content can result in high viscosity and glass transition temperature (Tg), as well as low ionic conductivity and diffusion coefficient. To solve this problem, a ternary electrolyte is prepared, in which NaBF4 with poly(ethylene glycol) dimethyl ether (PEGDME) as a ligand dissolved in DEMEBF4 [184]. The ionic conductivity, viscosity, and thermal stability of this new ternary ionic liquid electrolyte are investigated. The highest ionic conductivity of 1.2 mS cm-1 is obtained at 25°C when the molar ratio of PEGDME : NaBF4 : DEMEBF4 is kept at 8 : 1 : 2. In addition, the ternary electrolyte containing NaClO4, PC, and DEMETFSI has been reported to be used in SIBs [185]. The electrolyte exhibits a highly reversible insertion and extraction behavior of sodium ions with an optimized volume content of 70% DEMETFSI.
To further improve the solubility of ammonium-based ionic liquids for sodium salts, researchers design large volumes of ionic liquids with ether functional groups on quaternary ammonium cations. Because the ether group features with good sodium ion chelating properties, resulting in the excellent solvating ability of the electrolyte, it is reported that the solubility of NaTFSI in the modified ionic liquid electrolyte can be increased to 2.0 mol kg-1 [186]. A similar phenomenon exists in lithium systems [187].
6.6. Challenges and Benefits of Ionic Liquid-Based Electrolytes for SIBs
For most ionic liquids, cost is undoubtedly a serious issue. Battery-level ionic liquid must be very pure, which can lead to further cost increases. Such pure ionic liquid has not been widely synthesized at the industrial level, and therefore, it is difficult to estimate the price even for lithium-ion batteries. Generally, quaternary ammonium cations are more expensive than that of ternary, as well as fluorine-containing anions are also more expensive than that of non-fluorine-containing. Alkylammonium and alkylphosph are more economical and effective than dialkylimidazolium [188]. In contrast to TFSI- salt, FSI- salt can be synthesized without high cost electrofluorination. Therefore, by further improving the synthesis process and mass production in the near future, the price can be potentially reduced [189, 190]. Second, the high viscosity of ionic liquid (especially at low temperatures) can cause many follow-up problems including the porous separators and electrode impregnation [148]. Third, ionic liquids are not always good for the environment, and the decomposition products of some ionic liquids are harmful to the environment [191]. Finally, due to its high thermal stability, compared to traditional organic electrolytes, ionic liquids consume a lot of energy in the disposal process after the use of batteries.
Moreover, the mobility of metal cations in ionic liquid-based electrolytes is lower, which, together with higher viscosities, usually limits the rate performance of Na batteries. A high salt concentration ionic liquid electrolyte (also known as a “solvent in salt”) has been made to obtain good rate performance. Na mobility increases significantly with increasing salt concentration (0.3 when the salt content is 50 mol% and <0.1 when it is less than 15 mol%). Despite the higher viscosity, the sodium mobility is still improved [177, 192]. MD simulations show that the key to “unlocking” this conduction mechanism may be related to the coordination of anions with Na cations. These electrolytes make a balance between transference and viscosity, resulting in the perfect performance.
In summary, the imidazolium-based ionic liquids have attracted widespread attention in practical applications, because they feature with relatively low viscosity, high ionic conductivity, and high chemical stability. Pyrrolidinium-based ionic liquid electrolyte, which acted as an emerging system for SIB applications, shows a wide electrochemical window. The assembled batteries exhibit a specific capacity with a wide temperature range. However, the effects of Na ion concentration and operating temperature on the viscosity and their ionic conductivities need in-depth investigations. Moreover, the ammonium-based ionic liquid electrolytes can alleviate the problems of high viscosity and glass transition temperature (Tg), as well as low ionic conductivity and diffusion coefficient due to their high solubility in Na salts. The design of large volumes of ionic liquids with ether functional groups on quaternary ammonium cations is expected. Although there are many problems in the industrial application of ionic liquids, there still exists a great potential in using ionic liquids as SIB electrolytes. First, ionic liquid extends the temperature limit and cut-off voltage range of electrochemical measurements. The test condition flexibility not only facilitates the well-developed material assessment but also helps determine their optimal working environment. Second, ionic liquid-based electrolytes show improved cycling performance among wide range temperatures [170, 182, 193–195]. It is widely recognized that ionic liquid-based electrolytes can provide a uniform and strong SEI layer on active Na metals. In addition, due to the improved interface properties and the ionic conductivity, using ionic liquid at the high temperature can significantly increase the rate capability and power density. Finally, at present, although the use of full battery research to confirm the practical feasibility of SIBs with ionic liquid electrolytes is still in the preliminary stage, there have been several reports on full battery testing of ionic liquids using coin cells [153, 176, 183, 196–198]. There is also a report on prismatic batteries [199].
7. Conclusions and Outlook
Ionic liquids have found applications in almost all “postlithium” battery chemistry. In this review, we mainly introduce the basic properties of ionic liquid-based electrolyte and discuss their applications in aluminum-ion batteries, magnesium-ion batteries, and sodium-ion batteries. Then, we list the types of ionic liquid-based electrolytes that have been applied and analyze the existing advantages and limitations. In addition, the development directions of ionic liquid-based electrolytes in these three battery systems will be displayed.
In terms of electrolytes, the choice of aluminum ion batteries is very limited. Most studies are focused on ionic liquids containing corrosive and moisture sensitive chloroaluminate. Typical AlCl3-containing imidazolium-based ionic liquids are confronted with several problems (expensive, highly corrosive and sensitive, and low electrochemical windows), so they are far from commercial applications. Therefore, considering the practical application of AIBs, it is necessary to develop new, safe, nontoxic, and stable electrolytes. Generally, Al metal is very stable in nature because of its natural oxides. Thus, an ideal electrolyte should both act as a corrosive to dissolve the Al2O3 but also a corrosion inhibitor for the anode, polymer binder, current collector, and battery case. The development of advanced aluminum batteries must stem from enhancements and advancements in proper electrolyte systems. The discovery of new cathode materials can certainly be promoted by investigating appropriate electrolytes. So far, much work remains to do to develop the ideal aluminum-ion battery electrolyte.
Although ionic liquid-based electrolytes have also been explored for Mg batteries, the results are still quite limited. In ionic liquid-based electrolyte MIBs, the most commonly used magnesium salt is Mg(TFSI)2. However, the TFSI- anions are reductively unstable when coordinated with Mg2+. The performance of MIB with Mg(TFSI)2 as magnesium salts is poor. In addition, the layer formed on the Mg metal by the ionic liquid electrolyte decomposition prevents the transmission of Mg ions. However, reversible Mg deposition and dissolution in ionic liquids with TFSI- anions can be promoted by introducing oligomeric ether additives or using ether-functionalized ionic liquid cations. In these electrolytes, ether oxygen can replace the TFSI- anion in the Mg2+ coordination domain. In addition, in rechargeable magnesium batteries, ionic liquids can be effective additives to improve the electrochemical performance of electrolytes, such as ionic conductivity and anode stability.
At present, there are some ionic liquid-based electrolytes for SIBs with a variety of electrode materials that have been reported to exhibit better capacity, cycle stability, safety, and high-temperature performance than traditional solvent electrolytes. However, there are still many obstacles to realize a commercial SIB coupled with the ionic liquid electrolytes. For example, the high viscosity and density of ionic liquids may cause processing and wetting problems during the manufacturing process and may also decrease the specific energy of the battery. In addition, the battery rate performance must also be improved. We can expect further functionalization of cations and anions to improve the solvation and mobility of charge carriers to overcome above difficulties. Examples include the use of phosphonium systems, functionalized cations, and mixed anion electrolytes. In addition, the rate performance of SIBs can be improved by increasing the salt content to “superconcentrate” (>50 mol%). And the relative required ionic liquid will be greatly reduced leading to lower cost of the electrolyte. On the other hand, mixing ionic liquid-based materials with conventional “cheap” materials is more promising than pure materials to improve economic feasibility or some other disadvantages of ionic liquid. Ionic liquid-molecular solvent mixed electrolytes and ionic gels are all important strategies in this regard. If the economies of scale of these materials can be achieved, then these electrolytes seem to be a viable future technology for medium temperature applications.
The challenges and strategies of the three battery systems are illustrated in Figure 14. The research discussed here demonstrates that it is necessary to think outside the framework established by previous lithium-ion battery research when conducting research on ionic liquid electrolytes for aluminum-ion batteries, magnesium-ion batteries, and sodium-ion batteries, because these different metal elements possess inherent differences in chemical properties. In order to develop novel high-performance ionic liquid electrolytes, we need in-depth understanding and research on their fundamentals, paving the way for designing next-generation products.
Conflicts of Interest
The authors declare no competing financial interest.
Authors’ Contributions
Na Zhu and Kun Zhang contributed equally to this work.
Acknowledgments
This work was supported by the National Basic Research Program of China (Grant No. 2015CB251100) and the National Natural Science Foundation of China (Grant Nos. 21975026 and 22075028).
References
- M. Armand, F. Endres, D. R. MacFarlane, H. Ohno, and B. Scrosati, “Ionic-liquid materials for the electrochemical challenges of the future,” Nature Materials, vol. 8, no. 8, pp. 621–629, 2009. View at: Google Scholar
- Q. Yang, Z. Zhang, X. G. Sun, Y. S. Hu, H. Xing, and S. Dai, “Ionic liquids and derived materials for lithium and sodium batteries,” Chemical Society Reviews, vol. 47, no. 6, pp. 2020–2064, 2018. View at: Google Scholar
- J. Tarascon and M. Armand, “Issues and challenges facing rechargeable lithium batteries,” Nature, vol. 414, no. 6861, pp. 359–367, 2001. View at: Google Scholar
- F. Wu, H. Yang, Y. Bai, and C. Wu, “Paving the path toward reliable cathode materials for aluminum-ion batteries,” Advanced Materials, vol. 31, no. 16, p. 1806510, 2019. View at: Google Scholar
- G. A. Giffin, “Ionic liquid-based electrolytes for “beyond lithium” battery technologies,” Journal of Materials Chemistry A, vol. 4, no. 35, pp. 13378–13389, 2016. View at: Google Scholar
- J. Ma, F. Li, J. He et al., “Gradient solid electrolyte interphase and lithium ion solvation regulated by bisfluoroacetamide for stable lithium metal batteries,” Angewandte Chemie International Edition, 2020. View at: Publisher Site | Google Scholar
- S. Qi, H. Wang, J. He et al., “Electrolytes enriched by potassium perfluorinated sulfonates for lithium metal batteries,” Science Bulletin, 2020. View at: Publisher Site | Google Scholar
- K. Xu, “Nonaqueous liquid electrolytes for lithium-based rechargeable batteries,” Chemical Reviews, vol. 104, no. 10, pp. 4303–4418, 2004. View at: Publisher Site | Google Scholar
- V. Etacheri, R. Marom, R. Elazari, G. Salitra, and D. Aurbach, “Challenges in the development of advanced Li-ion batteries: a review,” Energy & Environmental Science, vol. 4, no. 9, pp. 3243–3262, 2011. View at: Google Scholar
- J. P. Hallett and T. Welton, “Room-temperature ionic liquid: slovents for synthesis and catalysis,” Chemical Reviews, vol. 111, no. 5, pp. 3508–3576, 2011. View at: Google Scholar
- Y. Wang, K. Zaghib, A. Guerfi, F. F. Bazito, R. M. Torresi, and J. R. Dahn, “Accelerating rate calorimetry studies of the reactions between ionic liquids and charged lithium ion battery electrode materials,” Electrochimica Acta, vol. 52, no. 22, pp. 6346–6352, 2007. View at: Google Scholar
- S. Zhang, N. Sun, X. He, X. Lu, and X. Zhang, “Physical properties of ionic liquids: database and evaluation,” Journal of Physical and Chemical Reference Data, vol. 35, no. 4, pp. 1475–1517, 2006. View at: Google Scholar
- J. D. Holbrey and R. D. Rogers, “Green chemistry and ionic liquids: synergies and ironies,” Cheminform, vol. 33, no. 48, pp. 243–243, 2002. View at: Google Scholar
- M. F. DR, M. Forsyth, P. C. Howlett et al., “Ionic liquids and their solid-state analogues as materials for energy generation and storage,” Nature Reviews Materials, vol. 1, no. 2, article 15005, 2016. View at: Google Scholar
- M. F. DR, N. Tachikawa, M. Forsyth et al., “Energy applications of ionic liquids,” Energy & Environmental Science, vol. 7, no. 1, pp. 232–250, 2014. View at: Google Scholar
- A. Ponrouch, D. Monti, A. Boschin, B. Steen, P. Johansson, and M. R. Palacín, “Non-aqueous electrolytes for sodium-ion batteries,” Journal of Materials Chemistry A, vol. 3, no. 1, pp. 22–42, 2015. View at: Google Scholar
- K. Xu, “Electrolytes and interphases in Li-ion batteries and beyond,” Chemical Reviews, vol. 114, no. 23, pp. 11503–11618, 2014. View at: Google Scholar
- H. Wang, S. Gu, Y. Bai et al., “Anion-effects on electrochemical properties of ionic liquid electrolytes for rechargeable aluminum batteries,” Journal of Materials Chemistry A, vol. 3, no. 45, pp. 22677–22686, 2015. View at: Google Scholar
- L. D. Reed, A. Arteaga, and E. J. Menke, “A combined experimental and computational study of an aluminum triflate/diglyme electrolyte,” The Journal of Physical Chemistry B, vol. 119, no. 39, pp. 12677–12681, 2015. View at: Google Scholar
- X. Zhang and T. M. Devine, “Passivation of aluminum in lithium-ion battery electrolytes with LiBOB,” Journal of the Electrochemical Society, vol. 153, no. 9, p. B365, 2006. View at: Google Scholar
- H. Yang, H. Li, J. Li et al., “The rechargeable aluminum battery: opportunities and challenges,” Angewandte Chemie International Edition, vol. 58, no. 35, pp. 11978–11996, 2019. View at: Google Scholar
- A. P. Karpinski, S. J. Russell, J. R. Serenyi, and J. P. Murphy, “Silver based batteries for high power applications,” Journal of Power Sources, vol. 91, no. 1, pp. 77–82, 2000. View at: Google Scholar
- O. Hasvold, K. H. Johansen, O. Mollestad, S. Forseth, and N. Storkersen, “The alkaline aluminium/hydrogen peroxide power source in the huginII unmanned underwater vehicle,” Journal of Power Sources, vol. 80, no. 1-2, pp. 254–260, 1999. View at: Google Scholar
- S. Licht, R. Tenne, H. Flaisher, and J. Manassen, “A pronounced cation effect on performance and stability of Cd-chalcogenide/polysulfide photoelectrochemical cells,” Journal of the Electrochemical Society, vol. 131, no. 4, pp. 950-951, 1984. View at: Google Scholar
- Y. Zhang, S. Liu, Y. Ji, J. Ma, and H. Yu, “Emerging nonaqueous aluminum-ion batteries: challenges, status, and perspectives,” Advanced Materials, vol. 30, no. 38, p. 1706310, 2018. View at: Google Scholar
- G. Torsi and G. Mamantov, “Acid-base properties of the systems AlCl3-MCl (M = Li, Na, K, Cs),” Inorganic Chemistry, vol. 11, no. 6, p. 1439, 1972. View at: Google Scholar
- G. D. Braunstein, L. E. Reichert, E. V. VanHall, J. L. Vaitukaitis, and G. T. Ross, “The effects of desialylation on the biologic and immunologic activity of human pituitary luteinizine hormone,” Biochemical and Biophysical Research Communications, vol. 42, no. 5, pp. 962–967, 1971. View at: Google Scholar
- G. L. Holleck, “The reduction of chlorine on carbon in AICI3-KCI-NaCI melts,” Journal of the Electrochemical Society, vol. 119, pp. 1158–1161, 1972. View at: Google Scholar
- N. Koura, “A preliminary investigation for an AI/AICI -NaCI/FeS secondary cell,” Journal of the Electrochemical Society, vol. 127, pp. 1529–1531, 1980. View at: Google Scholar
- H. A. Hjuler, “A novel inorganic low melting electrolyte for secondary aluminum-nickel sulfide batteries,” Journal of the Electrochemical Society, vol. 1987-7, pp. 657–668, 1989. View at: Google Scholar
- R. W. Berg, S. V. Winbush, and N. Bjerrum, “Negative oxidation states of the chalcogens in molten salts. 1. Raman spectroscopic studies on aluminum chlorosulfides formed in chloride and chloroaluminate melts and some related solid and dissolved compounds,” Inorganic Chemistry, vol. 19, pp. 2688–2698, 1980. View at: Google Scholar
- G. A. Elia, K. Marquardt, K. Hoeppner et al., “An overview and future perspectives of aluminum batteries,” Advanced Materials, vol. 28, no. 35, pp. 7564–7579, 2016. View at: Google Scholar
- Y. Zheng, K. Dong, Q. Wang, J. Zhang, and X. Lu, “Density, viscosity, and conductivity of lewis acidic 1-butyl- and 1-hydrogen-3-methylimidazolium chloroaluminate ionic liquids,” Journal of Chemical & Engineering Data, vol. 58, no. 1, pp. 32–42, 2012. View at: Google Scholar
- P. R. Gifford and J. B. Palmisano, “A substituted imidazolium chloroaluminate molten salt possessing an increased electrochemical window,” Journal of the Electrochemical Society, vol. 134, pp. 610–614, 1987. View at: Google Scholar
- C. M. Lang, K. Kim, L. Guerra, and P. A. Kohl, “Cation electrochemical stability in chloroaluminate ionic liquids,” Journal of Physical Chemistry B, vol. 109, pp. 19454–19462, 2005. View at: Google Scholar
- T. J. Melton and J. Joyce, “Electrochemical studies of sodium chloride as a lewis buffer for room temperature chloroaluminate molten salts,” Journal of the Electrochemical Society, vol. 137, pp. 3865–3869, 1990. View at: Google Scholar
- B. Vestergaard, N. J. Bjerrum, I. Petrushina, H. A. Hjuler, R. W. Berg, and M. Begtrup, “Molten triazolium chloride systems as new aluminum battery electrolytes,” Journal of the Electrochemical Society, vol. 140, pp. 3108–3113, 1993. View at: Google Scholar
- A. A. Fannin, L. A. King, J. A. Levisky, and J. S. Wilkes, “Properties of 1,3-dialkylimidazoIium chloride-aluminum chloride ionic liquids. 1. Ion interactions by nuclear magnetic resonance spectroscopy,” The journal of Physical Chemistry, vol. 88, pp. 2609–2614, 1984. View at: Google Scholar
- H. Yang, L. Yin, J. Liang et al., “An aluminum–sulfur battery with a fast kinetic response,” Angewandte Chemie International Edition, vol. 57, no. 7, pp. 1898–1902, 2018. View at: Google Scholar
- H. Chen, H. Xu, B. Zheng et al., “Oxide film efficiently suppresses dendrite growth in aluminum-ion battery,” ACS applied materials & interfaces, vol. 9, no. 27, pp. 22628–22634, 2017. View at: Google Scholar
- S. Choi, H. Go, G. Lee, and Y. Tak, “Electrochemical properties of aluminum anode in ionic liquid electrolyte for rechargeable aluminum-ion battery,” Physical Chemistry Chemical Physics, vol. 19, no. 13, pp. 8653–8656, 2013. View at: Google Scholar
- D. Lee, G. Lee, and Y. Tak, “Hypostatic instability of aluminum anode in acidic ionic liquid for aluminum-ion battery,” Nanotechnology, vol. 29, no. 36, article 36LT01, 2018. View at: Google Scholar
- F. Wu, N. Zhu, Y. Bai, Y. Gao, and C. Wu, “An interface-reconstruction effect for rechargeable aluminum battery in ionic liquid electrolyte to enhance cycling performances,” Green Energy & Environment, vol. 3, no. 1, pp. 71–77, 2018. View at: Google Scholar
- H. Wang, S. Gu, Y. Bai, S. Chen, F. Wu, and C. Wu, “A high-voltage and non-corrosive ionic liquid electrolyte used in rechargeable aluminum battery,” ACS applied materials & interfaces, vol. 8, pp. 27444–27448, 2016. View at: Google Scholar
- P. J. Eng, T. P. Trainor, G. E. Brown Jr. et al., “Structure of the hydrated a-Al2O3 (0001) surface,” Science, vol. 288, pp. 1029–1033, 2000. View at: Google Scholar
- N. Mortazavi, C. Geers, V. Esmaily et al., “Interplay of water and eeactive elements in oxidation of alumina-forming alloys,” Nature Materials, vol. 17, pp. 610–617, 2018. View at: Google Scholar
- X. Dong, H. Xu, H. Chen et al., “Commercial expanded graphite as high-performance cathode for low-cost aluminum-ion battery,” Carbon, vol. 148, pp. 134–140, 2019. View at: Google Scholar
- C. Zhang, R. He, J. Zhang, Y. Hu, Z. Wang, and X. Jin, “Amorphous carbon-derived nanosheet-bricked porous graphite as high-performance cathode for aluminum-ion batteries,” ACS applied materials & interfaces, vol. 10, no. 31, pp. 26510–26516, 2018. View at: Google Scholar
- J. Wei, W. Chen, D. Chen, and K. Yang, “An amorphous carbon-graphite composite cathode for long cycle life rechargeable aluminum ion batteries,” Journal of Materials Science & Technology, vol. 34, no. 6, pp. 983–989, 2018. View at: Google Scholar
- Z. Liu, J. Wang, H. Ding, S. Chen, X. Yu, and B. Lu, “Carbon nanoscrolls for aluminum battery,” ACS Nano, vol. 12, no. 8, pp. 8456–8466, 2018. View at: Google Scholar
- J. Smajic, A. Alazmi, N. Batra, T. Palanisamy, D. H. Anjum, and P. Costa, “Mesoporous reduced graphene oxide as a high capacity cathode for aluminum batteries,” Small, vol. 14, no. 51, p. 1803584, 2018. View at: Google Scholar
- D. Y. Wang, C. Y. Wei, M. C. Lin et al., “Advanced rechargeable aluminium ion battery with a high-quality natural graphite cathode,” Nature communications, vol. 8, p. 14283, 2017. View at: Google Scholar
- K. V. Kravchyk, S. Wang, L. Piveteau, and M. V. Kovalenko, “Efficient aluminum chloride–natural graphite battery,” Chemistry of Materials, vol. 29, no. 10, pp. 4484–4492, 2017. View at: Google Scholar
- X. Huang, Y. Liu, H. Zhang, J. Zhang, O. Noonan, and C. Yu, “Free-standing monolithic nanoporous graphene foam as a high performance aluminum-ion battery cathode,” Journal of Materials Chemistry A, vol. 5, no. 36, pp. 19416–19421, 2017. View at: Google Scholar
- G. A. Elia, I. Hasa, G. Greco et al., “Insights into the reversibility of aluminum graphite batteries,” Journal of Materials Chemistry A, vol. 5, no. 20, pp. 9682–9690, 2017. View at: Google Scholar
- L. Zhang, L. Chen, H. Luo, X. Zhou, and Z. Liu, “Large-sized few-layer graphene enables an ultrafast and long-life aluminum-ion battery,” Advanced Energy Materials, vol. 7, no. 15, p. 1700034, 2017. View at: Google Scholar
- L. Yang, Z. Wang, Y. Feng et al., “Flexible composite solid electrolyte facilitating highly stable “soft contacting” Li–electrolyte interface for solid state lithium-ion batteries,” Advanced Energy Materials, vol. 7, no. 22, p. 1701437, 2017. View at: Google Scholar
- H. Chen, H. Xu, S. Wang et al., “Ultrafast all-climate aluminum-graphene battery with quarter-million cycle life,” Science advances, vol. 3, no. 12, article EAAO7233, 2017. View at: Google Scholar
- Y. Wu, M. Gong, M. C. Lin et al., “3D graphitic foams derived from chloroaluminate anion intercalation for ultrafast aluminum-ion battery,” Advanced Materials, vol. 28, no. 41, pp. 9218–9222, 2016. View at: Google Scholar
- M. Chiku, H. Takeda, S. Matsumura, E. Higuchi, and H. Inoue, “Amorphous vanadium oxide/carbon composite positive electrode for rechargeable aluminum battery,” ACS applied materials & interfaces, vol. 7, no. 44, pp. 24385–24389, 2015. View at: Google Scholar
- H. Sun, W. Wang, Z. Yu, Y. Yuan, S. Wang, and S. Jiao, “A new aluminium-ion battery with high voltage, high safety and low cost,” Chemical Communications, vol. 51, no. 59, pp. 11892–11895, 2015. View at: Google Scholar
- M. C. Lin, M. Gong, B. Lu et al., “An ultrafast rechargeable aluminium-ion battery,” Nature, vol. 520, no. 7547, pp. 325–328, 2015. View at: Google Scholar
- J. Tu, H. Lei, Z. Yu, and S. Jiao, “Ordered WO3-x nanorods: facile synthesis and their electrochemical properties for aluminum-ion batteries,” Chemical Communications, vol. 54, no. 11, pp. 1343–1346, 2018. View at: Google Scholar
- X. Zhang, G. Zhang, S. Wang, S. Li, and S. Jiao, “Porous CuO microsphere architectures as high-performance cathode materials for aluminum-ion batteries,” Journal of Materials Chemistry A, vol. 6, no. 7, pp. 3084–3090, 2018. View at: Google Scholar
- S. Gu, H. Wang, C. Wu, Y. Bai, H. Li, and F. Wu, “Confirming reversible Al3+ storage mechanism through intercalation of Al3+ into V2O5 nanowires in a rechargeable aluminum battery,” Energy Storage Materials, vol. 6, pp. 9–17, 2017. View at: Google Scholar
- H. Wang, X. Bi, Y. Bai et al., “Open-structured V2O5·nH2O nanoflakes as highly reversible cathode material for monovalent and multivalent intercalation batteries,” Advanced Energy Materials, vol. 7, no. 14, p. 1602720, 2017. View at: Google Scholar
- K. Suto, A. Nakata, H. Murayama, T. Hirai, J.-I. Yamaki, and Z. Ogumi, “Electrochemical properties of Al/vanadium chloride batteries with AlCl3-1-Ethyl-3-methylimidazolium chloride electrolyte,” Journal of the Electrochemical Society, vol. 163, no. 5, pp. A742–A747, 2016. View at: Google Scholar
- H. Wang, Y. Bai, S. Chen et al., “Binder-free V2O5 cathode for greener rechargeable aluminum battery,” ACS applied materials & interfaces, vol. 7, no. 1, pp. 80–84, 2015. View at: Google Scholar
- H. Li, H. Yang, Z. Sun, Y. Shi, H.-M. Cheng, and F. Li, “A highly reversible Co3S4 microsphere cathode material for aluminum-ion batteries,” Nano Energy, vol. 56, pp. 100–108, 2019. View at: Google Scholar
- Y. Hu, D. Ye, B. Luo et al., “A binder-free and free-standing cobalt sulfide@carbon nanotube cathode material for aluminum-ion batteries,” Advanced Materials, vol. 30, no. 2, p. 1703824, 2018. View at: Google Scholar
- Z. Li, B. Niu, J. Liu, J. Li, and F. Kang, “Rechargeable aluminum-ion battery based on MoS2 microsphere cathode,” ACS applied materials & interfaces, vol. 10, no. 11, pp. 9451–9459, 2018. View at: Google Scholar
- S. Wang, S. Jiao, J. Wang et al., “High-performance aluminum-ion battery with CuS@C microsphere composite cathode,” ACS Nano, vol. 11, no. 1, pp. 469–477, 2017. View at: Google Scholar
- Y. Hu, B. Luo, D. Ye, X. Zhu, M. Lyu, and L. Wang, “An innovative freeze-dried reduced graphene oxide supported SnS2 cathode active material for aluminum-ion batteries,” Advanced Materials, vol. 29, no. 48, p. 1606132, 2017. View at: Google Scholar
- T. Mori, Y. Orikasa, K. Nakanishi et al., “Discharge/charge reaction mechanisms of FeS2 cathode material for aluminum rechargeable batteries at 55°C,” Journal of Power Sources, vol. 313, pp. 9–14, 2016. View at: Google Scholar
- B. Lee, H. R. Lee, T. Yim et al., “Investigation on the structural evolutions during the insertion of aluminum ions into Mo6S8 Chevrel phase,” Journal of the Electrochemical Society, vol. 163, no. 6, pp. A1070–A1076, 2016. View at: Google Scholar
- L. Geng, G. Lv, X. Xing, and J. Guo, “Reversible electrochemical intercalation of aluminum in Mo6S8,” Chemistry of Materials, vol. 27, no. 14, pp. 4926–4929, 2015. View at: Google Scholar
- J. Jiang, H. Li, T. Fu, B. J. Hwang, X. Li, and J. Zhao, “One-dimensional Cu2-xSe nanorods as the cathode material for high-performance aluminum-ion battery,” ACS applied materials & interfaces, vol. 10, no. 21, pp. 17942–17949, 2018. View at: Google Scholar
- X. Huang, Y. Liu, C. Liu, J. Zhang, O. Noonan, and C. Yu, “Rechargeable aluminum-selenium batteries with high capacity,” Chemical science, vol. 9, no. 23, pp. 5178–5182, 2018. View at: Google Scholar
- T. Cai, L. Zhao, H. Hu et al., “Stable CoSe2/carbon nanodice@reduced graphene oxide composites for high-performance rechargeable aluminum-ion batteries,” Energy & Environmental Science, vol. 11, no. 9, pp. 2341–2347, 2018. View at: Google Scholar
- M. Walter, K. V. Kravchyk, C. Bofer, R. Widmer, and M. V. Kovalenko, “Polypyrenes as high-performance cathode materials for aluminum batteries,” Advanced Materials, vol. 30, no. 15, p. 1705644, 2018. View at: Google Scholar
- S. Liu, G. L. Pan, G. R. Li, and X. P. Gao, “Copper hexacyanoferrate nanoparticles as cathode material for aqueous Al-ion batteries,” Journal of Materials Chemistry A, vol. 3, no. 3, pp. 959–962, 2015. View at: Google Scholar
- L. D. Reed, S. N. Ortiz, M. Xiong, and E. J. Menke, “A rechargeable aluminum-ion battery utilizing a copper hexacyanoferrate cathode in an organic electrolyte,” Chemical Communications, vol. 51, no. 76, pp. 14397–14400, 2015. View at: Google Scholar
- Z. Li, K. Xiang, W. Xing, W. C. Carter, and Y.-M. Chiang, “Reversible aluminum-ion intercalation in prussian blue analogs and demonstration of a high-power aluminum-ion asymmetric capacitor,” Advanced Energy Materials, vol. 5, no. 5, p. 1401410, 2015. View at: Google Scholar
- X. Yu and A. Manthiram, “Electrochemical energy storage with a reversible nonaqueous room-temperature aluminum-sulfur chemistry,” Advanced Energy Materials, vol. 7, no. 18, p. 1700561, 2017. View at: Google Scholar
- G. Cohn, L. Ma, and L. A. Archer, “A novel non-aqueous aluminum sulfur battery,” Journal of Power Sources, vol. 283, pp. 416–422, 2015. View at: Google Scholar
- X. Yu, M. J. Boyer, G. S. Hwang, and A. Manthiram, “Room-temperature aluminum-sulfur batteries with a lithium-ion-mediated ionic liquid electrolyte,” Chem, vol. 4, no. 3, pp. 586–598, 2018. View at: Google Scholar
- K. Zhang, T. H. Lee, J. H. Cha et al., “Two-dimensional boron nitride as a sulfur fixer for high performance rechargeable aluminum-sulfur batteries,” Scientific reports, vol. 9, no. 1, p. 13573, 2019. View at: Google Scholar
- S. Takahashi, L. A. Curtiss, D. Gosztola, N. Koura, and M. Saboungi, “Molecular orbital calculations and raman measurements for 1-ethyl-3-methylimidazolium chloroaluminates,” Inorganic Chemistry, vol. 34, no. 11, pp. 2990–2993, 1995. View at: Google Scholar
- R. Perez, I. Antonio, and X. Ji, “Anion hosting cathodes in dual-ion batteries,” ACS Energy Letters, vol. 2, pp. 1762–1770, 2017. View at: Google Scholar
- L. Geng, J. P. Scheifers, C. Fu, J. Zhang, B. P. T. Fokwa, and J. Guo, “Titanium sulfides as intercalation-type cathode materials for rechargeable aluminum batteries,” ACS applied materials & interfaces, vol. 9, no. 25, pp. 21251–21257, 2017. View at: Google Scholar
- J. Shi, J. Zhang, and J. Guo, “Avoiding pitfalls in rechargeable aluminum batteries research,” ACS Energy Letters, vol. 4, no. 9, pp. 2124–2129, 2019. View at: Google Scholar
- P. R. Gifford and J. B. Palmisano, “An aluminum chlorine rechargeable cell employing a room-temperature molten salt electrolyte,” Journal of the Electrochemical Society, vol. 135, no. 3, pp. 650–654, 1988. View at: Google Scholar
- C. Yang, J. Chen, X. Ji et al., “Aqueous Li-ion battery enabled by halogen conversion-intercalation chemistry in graphite,” Nature, vol. 569, no. 7755, pp. 245–250, 2019. View at: Google Scholar
- B. Wang, L. Qin, T. Mu, Z. Xue, and G. Gao, “Are ionic liquids chemically stable?” Chemical Reviews, vol. 117, no. 10, pp. 7113–7131, 2017. View at: Google Scholar
- S. Guo, S. Zhang, L. Wu, and S. Sun, “Co/CoO nanoparticles assembled on graphene for electrochemical reduction of oxygen,” Angewandte Chemie International Edition, vol. 51, no. 47, pp. 11770–11773, 2012. View at: Google Scholar
- Y. Tan, C. Xu, G. Chen, X. Fang, N. Zheng, and Q. Xie, “Facile synthesis of manganese-oxide-containing mesoporous nitrogen-doped carbon for efficient oxygen reduction,” Advanced Functional Materials, vol. 22, no. 21, pp. 4584–4591, 2012. View at: Google Scholar
- Y. Ito and T. Nohira, “Non-conventional electrolytes for electrochemical applications,” Electrochimica Acta, vol. 45, pp. 2611–2622, 2000. View at: Google Scholar
- T. Jiang, M. J. Chollier Brym, G. Dubé, A. Lasia, and G. M. Brisard, “Electrodeposition of aluminium from ionic liquids: part I—electrodeposition and surface morphology of aluminium from aluminium chloride (AlCl3)–1-ethyl-3-methylimidazolium chloride ([EMIm]Cl) ionic liquids,” Surface and Coatings Technology, vol. 201, no. 1-2, pp. 1–9, 2006. View at: Google Scholar
- L. D. Reed and E. Menke, “The roles of V2O5 and stainless steel in rechargeable Al-ion batteries,” Journal of the Electrochemical Society, vol. 160, no. 6, pp. A915–A917, 2013. View at: Google Scholar
- P. C. Lin, I. W. Sun, J. K. Chang, C.-J. Su, and J.-C. Lin, “Corrosion characteristics of nickel, copper, and stainless steel in a lewis neutral chloroaluminate ionic liquid,” Corrosion Science, vol. 53, no. 12, pp. 4318–4323, 2011. View at: Google Scholar
- S. Wang, Z. Yu, J. Tu et al., “A novel aluminum-ion battery: Al/AlCl3-[EMIm]Cl/Ni3S2@graphene,” Advanced Energy Materials, vol. 6, no. 13, p. 1600137, 2016. View at: Google Scholar
- C. H. Tseng, J. K. Chang, J. R. Chen, W. T. Tsai, M.-J. Deng, and I. W. Sun, “Corrosion behaviors of materials in aluminum chloride–1-ethyl-3-methylimidazolium chloride ionic liquid,” Electrochemistry Communications, vol. 12, no. 8, pp. 1091–1094, 2010. View at: Google Scholar
- A. Kitada, K. Nakamura, K. Fukami, and K. Murase, “Electrochemically active species in aluminum electrodeposition baths of AlCl3/glyme solutions,” Electrochimica Acta, vol. 211, pp. 561–567, 2016. View at: Google Scholar
- E. Peled and E. Gileadi, “The electrodeposition of aluminum from aromatic hydrocarbon I. Composition of baths and the effect of additives,” Journal of the Electrochemical Society, vol. 123, no. 1, pp. 15–19, 1976. View at: Google Scholar
- P. Eiden, Q. Liu, S. Zein El Abedin, F. Endres, and I. Krossing, “An experimental and theoretical study of the aluminium species present in mixtures of AlCl3 with the ionic liquids [BMP]Tf2N and [EMIm]Tf2N,” Chemistry, vol. 15, no. 14, pp. 3426–3434, 2009. View at: Google Scholar
- S. Licht, G. Levitin, R. Tel-Vered, and C. Yarnitzky, “The effect of water on the anodic dissolution of aluminum in non-aqueous electrolytes,” Electrochemistry Communications, vol. 2, pp. 329–333, 2000. View at: Google Scholar
- H. Yang, F. Wu, Y. Bai, and C. Wu, “Toward better electrode/electrolyte interfaces in the ionic-liquid-based rechargeable aluminum batteries,” Journal of Energy Chemistry, vol. 45, pp. 98–102, 2020. View at: Google Scholar
- R. J. Gale and R. A. Osteryoung, “Dissociative chlorination of nitrogen oxides and oxy anions in molten sodium chloride-aluminum chloride solvents,” Inorganic Chemistry, vol. 14, no. 6, pp. 1232–1236, 1975. View at: Google Scholar
- F. Endres, O. Hofft, N. Borisenko et al., “Do solvation layers of ionic liquids influence electrochemical reactions?” Physical Chemistry Chemical Physics, vol. 12, pp. 1724–1732, 2010. View at: Google Scholar
- A. P. Abbott, R. C. Harris, Y. T. Hsieh, K. S. Ryder, and I. W. Sun, “Aluminium electrodeposition under ambient conditions,” Physical Chemistry Chemical Physics, vol. 16, pp. 14675–14681, 2014. View at: Google Scholar
- D. Aurbach, Z. Lu, A. Schechter et al., “Prototype systems for rechargeable magnesium batteries,” Nature, vol. 407, no. 6805, pp. 724–727, 2000. View at: Google Scholar
- Z. Lu, A. Schechter, M. Moshkovich, and D. Aurbach, “On the electrochemical behavior of magnesium electrodes in polar aprotic electrolyte solutions,” Journal of Electroanalytical Chemistry, vol. 466, pp. 203–217, 1999. View at: Google Scholar
- J. D. Genders and D. Pletcher, “Studies using microelectrodes of the Mg(II)/Mg couple in tetrahydrofuran and propylene carbonate,” Journal of Electroanalytical Chemistry, vol. 199, no. 1, pp. 93–100, 1986. View at: Google Scholar
- O. R. Brown and R. Mcintyre, “The magnesium and magnesium amalgam electrodes in aprotic organic solvents a kinetic study,” Electrochimica Acta, vol. 30, no. 5, pp. 627–633, 1985. View at: Google Scholar
- P. Novak, R. Tuhof, and O. Haas, “Magnesium insertion electrodes of rechargeable nonaqueous batteries. A competitive alternative to lithium?” Electrochimica Acta, vol. 45, pp. 351–367, 1999. View at: Google Scholar
- R. Deivanayagam, B. J. Ingram, and R. Shahbazian-Yassar, “Progress in development of electrolytes for magnesium batteries,” Energy Storage Materials, vol. 21, pp. 136–153, 2019. View at: Google Scholar
- D. Aurbach, Y. Gofer, Z. Lu et al., “A short review on the comparison between Li battery systems and rechargeable magnesium battery technology,” Journal of Power Sources, vol. 97-98, pp. 28–32, 2001. View at: Google Scholar
- C. B. Bucur, T. Gregory, A. G. Oliver, and J. Muldoon, “Confession of a magnesium battery,” The journal of physical chemistry letters, vol. 6, no. 18, pp. 3578–3591, 2015. View at: Google Scholar
- D. Aurbach, H. Gizbar, A. Schechter et al., “Electrolyte solutions for rechargeable magnesium batteries based on organomagnesium chloroaluminate complexes,” Journal of the Electrochemical Society, vol. 149, p. A115, 2002. View at: Google Scholar
- J. Muldoon, C. B. Bucur, A. G. Oliver et al., “Electrolyte roadblocks to a magnesium rechargeable battery,” Energy & Environmental Science, vol. 5, no. 3, p. 5941, 2012. View at: Google Scholar
- H. S. Kim, T. S. Arthur, G. D. Allred et al., “Structure and compatibility of a magnesium electrolyte with a sulphur cathode,” Nature communications, vol. 2, p. 427, 2011. View at: Google Scholar
- J. Muldoon, C. B. Bucur, and T. Gregory, “Quest for nonaqueous multivalent secondary batteries: magnesium and beyond,” Chemical Reviews, vol. 114, no. 23, pp. 11683–11720, 2014. View at: Google Scholar
- Y. NuLi, J. Yang, and R. Wu, “Reversible deposition and dissolution of magnesium from BMIMBF4 ionic liquid,” Electrochemistry Communications, vol. 7, pp. 1105–1110, 2005. View at: Google Scholar
- Y. NuLi, J. Yang, J. Wang, J. Xu, and P. Wang, “Electrochemical magnesium deposition and dissolution with high efficiency in ionic liquid,” Electrochemical and Solid-State Letters, vol. 8, no. 11, p. C166, 2005. View at: Google Scholar
- P. Wang, Y. NuLi, J. Yang, and Z. Feng, “Mixed ionic liquids as electrolyte for reversible deposition and dissolution of magnesium,” Surface and Coatings Technology, vol. 201, pp. 3783–3787, 2006. View at: Google Scholar
- M. Forsyth, P. C. Howlett, S. K. Tan, D. R. MacFarlane, and N. Birbilis, “An ionic liquid surface treatment for corrosion protection of magnesium alloy AZ31,” Electrochemical and Solid-State Letters, vol. 9, no. 11, p. B52, 2006. View at: Google Scholar
- N. Amir, Y. Vestfrid, O. Chusid, Y. Gofer, and D. Aurbach, “Progress in nonaqueous magnesium electrochemistry,” Journal of Power Sources, vol. 174, no. 2, pp. 1234–1240, 2007. View at: Google Scholar
- G. Vardar, A. E. Sleightholme, J. Naruse, H. Hiramatsu, D. J. Siegel, and C. W. Monroe, “Electrochemistry of magnesium electrolytes in ionic liquids for secondary batteries,” ACS applied materials & interfaces, vol. 6, no. 20, pp. 18033–18039, 2014. View at: Google Scholar
- P. Saha, M. K. Datta, O. I. Velikokhatnyi, A. Manivannan, D. Alman, and P. N. Kumta, “Rechargeable magnesium battery: current status and key challenges for the future,” Progress in Materials Science, vol. 66, pp. 1–86, 2014. View at: Google Scholar
- N. Yoshimoto, Y. Tomonaga, M. Ishikawa, and M. Morita, “Ionic conductance of polymeric electrolytes consisting of magnesium salts dissolved in cross-linked polymer matrix with linear polyether,” Electrochimica Acta, vol. 46, pp. 1195–1200, 2001. View at: Google Scholar
- N. Yoshimoto, S. Yakushiji, M. Ishikawa, and M. Morita, “Ionic conductance behavior of polymeric electrolytes containing magnesium salts and their application to rechargeable batteries,” Solid State Ionics, vol. 152-153, pp. 259–266, 2002. View at: Google Scholar
- M. Morita, T. Shirai, N. Yoshimoto, and M. Ishikawa, “Ionic conductance behavior of polymeric gel electrolyte containing ionic liquid mixed with magnesium salt,” Journal of Power Sources, vol. 139, no. 1-2, pp. 351–355, 2005. View at: Google Scholar
- N. Yoshimoto, T. Shirai, and M. Morita, “A novel polymeric gel electrolyte systems containing magnesium salt with ionic liquid,” Electrochimica Acta, vol. 50, no. 19, pp. 3866–3871, 2005. View at: Google Scholar
- G. T. Cheek, W. E. O’Grady, S. Z. El Abedin, E. M. Moustafa, and F. Endres, “Studies on the electrodeposition of magnesium in ionic liquids,” Journal of the Electrochemical Society, vol. 155, no. 1, p. D91, 2008. View at: Google Scholar
- N. Yoshimoto, M. Matsumoto, M. Egashia, and M. Morita, “Mixed electrolyte consisting of ethylmagnesiumbromide with ionic liquid for rechargeable magnesium electrode,” Journal of Power Sources, vol. 195, no. 7, pp. 2096–2098, 2010. View at: Google Scholar
- A. Kitada, Y. Kang, K. Matsumoto, K. Fukami, R. Hagiwara, and K. Murase, “Room temperature magnesium electrodeposition from glyme-coordinated ammonium amide electrolytes,” Journal of the Electrochemical Society, vol. 162, no. 8, pp. D389–D396, 2015. View at: Google Scholar
- G. A. Giffin, A. Moretti, S. Jeong, and S. Passerini, “Complex nature of ionic coordination in magnesium ionic liquid-based electrolytes: solvates with mobile Mg2+ cations,” The Journal of Physical Chemistry C, vol. 118, no. 19, pp. 9966–9973, 2014. View at: Google Scholar
- T. Watkins and D. A. Buttry, “Determination of Mg(2+) speciation in a TFSI(-)-based ionic liquid with and without chelating ethers using Raman spectroscopy,” The Journal of Physical Chemistry B, vol. 119, no. 23, pp. 7003–7014, 2015. View at: Google Scholar
- G. A. Giffin, J. Tannert, S. Jeong, W. Uhl, and S. Passerini, “Crystalline complexes of Pyr12O1TFSI-based ionic liquid electrolytes,” The Journal of Physical Chemistry C, vol. 119, no. 11, pp. 5878–5887, 2015. View at: Google Scholar
- T. Kakibe, N. Yoshimoto, M. Egashira, and M. Morita, “Optimization of cation structure of imidazolium-based ionic liquids as ionic solvents for rechargeable magnesium batteries,” Electrochemistry Communications, vol. 12, no. 11, pp. 1630–1633, 2010. View at: Google Scholar
- T. Kakibe, J.-Y. Hishii, N. Yoshimoto, M. Egashira, and M. Morita, “Binary ionic liquid electrolytes containing organo-magnesium complex for rechargeable magnesium batteries,” Journal of Power Sources, vol. 203, pp. 195–200, 2012. View at: Google Scholar
- X. Gao, A. Mariani, S. Jeong et al., “Prototype rechargeable magnesium batteries using ionic liquid electrolytes,” Journal of Power Sources, vol. 423, pp. 52–59, 2019. View at: Google Scholar
- M. M. Huie, C. A. Cama, P. F. Smith et al., “Ionic liquid hybrids: progress toward non-corrosive electrolytes with high-voltage oxidation stability for magnesium-ion based batteries,” Electrochimica Acta, vol. 219, pp. 267–276, 2016. View at: Google Scholar
- M. Kar, Z. Ma, L. M. Azofra, K. Chen, M. Forsyth, and D. R. MacFarlane, “Ionic liquid electrolytes for reversible magnesium electrochemistry,” Chemical communications, vol. 52, pp. 4033–4036, 2016. View at: Google Scholar
- S. Su, Y. NuLi, N. Wang, D. Yusipu, J. Yang, and J. Wang, “Magnesium borohydride-based electrolytes containing 1-butyl-1-methylpiperidinium bis(trifluoromethyl sulfonyl)imide ionic liquid for rechargeable magnesium batteries,” Journal of the Electrochemical Society, vol. 163, no. 13, pp. D682–D688, 2016. View at: Google Scholar
- J. H. Cho, J. H. Ha, S. H. Lee et al., “Effect of 1-allyl-1-methylpyrrolidinium chloride addition to ethylmagnesium bromide electrolyte on a rechargeable magnesium battery,” Electrochimica Acta, vol. 231, pp. 379–385, 2017. View at: Google Scholar
- B. Pan, K.-C. Lau, J. T. Vaughey, L. Zhang, Z. Zhang, and C. Liao, “Ionic liquid as an effective additive for rechargeable magnesium batteries,” Journal of the Electrochemical Society, vol. 164, no. 4, pp. A902–A906, 2017. View at: Google Scholar
- K. Matsumoto, J. Hwang, S. Kaushik, C.-Y. Chen, and R. Hagiwara, “Advances in sodium secondary batteries utilizing ionic liquid electrolytes,” Energy & Environmental Science, vol. 12, no. 11, pp. 3247–3287, 2019. View at: Google Scholar
- G. L. Xu, R. Amine, A. Abouimrane et al., “Challenges in developing electrodes, electrolytes, and diagnostics tools to understand and advance sodium-ion batteries,” Advanced Energy Materials, vol. 8, no. 14, p. 1702403, 2018. View at: Google Scholar
- M. Okoshi, Y. Yamada, A. Yamada, and H. Nakai, “Theoretical analysis on de-solvation of lithium, sodium, and magnesium cations to organic electrolyte solvents,” Journal of the Electrochemical Society, vol. 160, no. 11, pp. A2160–A2165, 2013. View at: Google Scholar
- J. B. Goodenough and Y. Kim, “Challenges for rechargeable Li batteries,” Chemistry of Materials, vol. 22, no. 3, pp. 587–603, 2010. View at: Google Scholar
- H. Pan, Y. S. Hu, and L. Chen, “Room-temperature stationary sodium-ion batteries for large-scale electric energy storage,” Energy & Environmental Science, vol. 6, no. 8, pp. 2338–2360, 2013. View at: Google Scholar
- X. Wang, Z. Shang, A. Yang et al., “Combining quinone cathode and ionic liquid electrolyte for organic sodium-ion batteries,” Chem, vol. 5, no. 2, pp. 364–375, 2019. View at: Google Scholar
- S. Brutti, M. A. Navarra, G. Maresca et al., “Ionic liquid electrolytes for room temperature sodium battery systems,” Electrochimica Acta, vol. 306, pp. 317–326, 2019. View at: Google Scholar
- B. R. Long, S. G. Rinaldo, K. G. Gallagher et al., “Enabling high-energy, high-voltage lithium-ion cells: standardization of coin-cell assembly, electrochemical testing, and evaluation of full cells,” Journal of the Electrochemical Society, vol. 163, no. 14, pp. A2999–A3009, 2016. View at: Google Scholar
- D. Larcher and J. M. Tarascon, “Towards greener and more sustainable batteries for electrical energy storage,” Nature chemistry, vol. 7, no. 1, pp. 19–29, 2015. View at: Google Scholar
- G. T. Kim, S. S. Jeong, M. Joost et al., “Use of natural binders and ionic liquid electrolytes for greener and safer lithium-ion batteries,” Journal of Power Sources, vol. 196, no. 4, pp. 2187–2194, 2011. View at: Google Scholar
- H. Che, S. Chen, Y. Xie et al., “Electrolyte design strategies and research progress for room-temperature sodium-ion batteries,” Energy & Environmental Science, vol. 10, no. 5, pp. 1075–1101, 2017. View at: Google Scholar
- Y. Huang, L. Zhao, L. Li, M. Xie, F. Wu, and R. Chen, “Electrolytes and electrolyte/electrode interfaces in sodium-ion batteries: from scientific research to practical application,” Advanced materials, vol. 31, no. 21, p. 1808393, 2019. View at: Google Scholar
- J. S. Wilkes and M. J. Zaworotko, “Air and water stable 1-ethyl-3-methylimidazolium based ionic liquids,” Journal of the Chemical Society, Chemical Communications, vol. 23, no. 43, pp. 965–965, 1992. View at: Google Scholar
- J. Fuller, R. T. Carlin, H. D. Long, and D. Haworth, “Structure of 1-ethyl-3-methylimidazolium hexafluorophosphate: model for room temperature molten salts,” Journal of the Chemical Society, Chemical Communications, vol. 3, no. 3, pp. 299-300, 1994. View at: Google Scholar
- Z. Xue, L. Qin, J. Jiang, T. Mu, and G. Gao, “Thermal, electrochemical and radiolytic stabilities of ionic liquids,” Physical Chemistry Chemical Physics, vol. 20, no. 13, pp. 8382–8402, 2018. View at: Google Scholar
- J. F. Wishart, “Energy applications of ionic liquids,” Energy & Environmental Science, vol. 2, no. 9, pp. 956–961, 2009. View at: Google Scholar
- H. Srour, L. Chancelier, E. Bolimowska et al., “Ionic liquid-based electrolytes for lithium-ion batteries: review of performances of various electrode systems,” Journal of Applied Electrochemistry, vol. 46, no. 2, pp. 149–155, 2015. View at: Google Scholar
- A. Bagno, C. Butts, C. Chiappe et al., “The effect of the anion on the physical properties of trihalide-based N,N-dialkylimidazolium ionic liquids,” Organic& Biomolecular chemistry, vol. 3, no. 9, pp. 1624–1630, 2005. View at: Google Scholar
- L. S. Plashnitsa, E. Kobayashi, Y. Noguchi, S. Okada, and J.-I. Yamaki, “Performance of NASICON symmetric cell with ionic liquid electrolyte,” Journal of the Electrochemical Society, vol. 157, no. 4, p. A536, 2010. View at: Google Scholar
- F. Wu, N. Zhu, Y. Bai, L. Liu, H. Zhou, and C. Wu, “Highly safe ionic liquid electrolytes for sodium-ion battery: wide electrochemical window and good thermal stability,” ACS applied materials & interfaces, vol. 8, no. 33, pp. 21381–21386, 2016. View at: Google Scholar
- D. Monti, E. Jónsson, M. R. Palacín, and P. Johansson, “Ionic liquid-based electrolytes for sodium-ion batteries: Na+ solvation and ionic conductivity,” Journal of Power Sources, vol. 245, pp. 630–636, 2014. View at: Google Scholar
- P. Kubisiak and A. Eilmes, “Molecular dynamics simulations of ionic liquid-based electrolytes for Na-ion batteries: effects of force field,” The journal of Physical Chemistry B, vol. 121, no. 42, pp. 9957–9968, 2017. View at: Google Scholar
- F. Wu, N. Zhu, Y. Bai et al., “Unveil the mechanism of solid electrolyte interphase on Na3V2(PO4)3 formed by a novel NaPF6/BMITFSI ionic liquid electrolyte,” Nano Energy, vol. 51, pp. 524–532, 2018. View at: Google Scholar
- L. Xue, T. G. Tucker, and C. A. Angell, “Ionic liquid redox catholyte for high energy efficiency, low-cost energy storage,” Advanced Energy Materials, vol. 5, no. 12, p. 1500271, 2015. View at: Google Scholar
- Y. Gao, G. Chen, X. Wang et al., “PY13FSI-infiltrated SBA-15 as nonflammable and high ion-conductive ionogel electrolytes for quasi-solid-state sodium-ion batteries,” ACS applied materials & interfaces, vol. 12, no. 20, pp. 22981–22991, 2020. View at: Google Scholar
- C. Ding, T. Nohira, K. Kuroda et al., “NaFSA–C1C3pyrFSA ionic liquids for sodium secondary battery operating over a wide temperature range,” Journal of Power Sources, vol. 238, pp. 296–300, 2013. View at: Google Scholar
- C. Ding, T. Nohira, R. Hagiwara et al., “Na[FSA]-[C3C1pyrr][FSA] ionic liquids as electrolytes for sodium secondary batteries: effects of Na ion concentration and operation temperature,” Journal of Power Sources, vol. 269, pp. 124–128, 2014. View at: Google Scholar
- A. Fukunaga, T. Nohira, R. Hagiwara et al., “A safe and high-rate negative electrode for sodium-ion batteries: hard carbon in NaFSA-C1C3pyrFSA ionic liquid at 363 K,” Journal of Power Sources, vol. 246, pp. 387–391, 2014. View at: Google Scholar
- C. H. Wang, C. H. Yang, and J. K. Chang, “Suitability of ionic liquid electrolytes for room-temperature sodium-ion battery applications,” Chemical Communications, vol. 52, no. 72, pp. 10890–10893, 2016. View at: Google Scholar
- M. Forsyth, H. Yoon, F. Chen et al., “Novel Na+ ion diffusion mechanism in mixed organic–inorganic ionic liquid electrolyte leading to high Na+ transference number and stable, high rate electrochemical cycling of sodium cells,” The Journal of Physical Chemistry C, vol. 120, no. 8, pp. 4276–4286, 2016. View at: Google Scholar
- Y. Kim, G.-T. Kim, S. Jeong et al., “Large-scale stationary energy storage: seawater batteries with high rate and reversible performance,” Energy Storage Materials, vol. 16, pp. 56–64, 2019. View at: Google Scholar
- S. A. Mohd Noor, P. C. Howlett, D. R. MacFarlane, and M. Forsyth, “Properties of sodium-based ionic liquid electrolytes for sodium secondary battery applications,” Electrochimica Acta, vol. 114, pp. 766–771, 2013. View at: Google Scholar
- N. Wongittharom, T. C. Lee, C. H. Wang, Y.-C. Wang, and J.-K. Chang, “Electrochemical performance of Na/NaFePO4 sodium-ion batteries with ionic liquid electrolytes,” Journal of Materials Chemistry A, vol. 2, no. 16, p. 5655, 2014. View at: Google Scholar
- C.-H. Wang, Y.-W. Yeh, N. Wongittharom et al., “Rechargeable Na/Na0.44MnO2 cells with ionic liquid electrolytes containing various sodium solutes,” Journal of Power Sources, vol. 274, pp. 1016–1023, 2015. View at: Google Scholar
- N. Wongittharom, C. H. Wang, Y. C. Wang, C. H. Yang, and J. K. Chang, “Ionic liquid electrolytes with various sodium solutes for rechargeable Na/NaFePO4 batteries operated at elevated temperatures,” ACS applied materials & interfaces, vol. 6, no. 20, pp. 17564–17570, 2014. View at: Google Scholar
- I. Hasa, S. Passerini, and J. Hassoun, “Characteristics of an ionic liquid electrolyte for sodium-ion batteries,” Journal of Power Sources, vol. 303, pp. 203–207, 2016. View at: Google Scholar
- M. Egashira, T. Asai, N. Yoshimoto, and M. Morita, “Ionic conductivity of ternary electrolyte containing sodium salt and ionic liquid,” Electrochimica Acta, vol. 58, pp. 95–98, 2011. View at: Google Scholar
- M. Egashira, T. Tanaka, N. Yoshimoto, and M. Morita, “Influence of ionic liquid species in non-aqueous electrolyte on sodium insertion into hard carbon,” Electrochemistry, vol. 80, no. 10, pp. 755–758, 2012. View at: Google Scholar
- C. R. Pope, M. Kar, D. R. MacFarlane, M. Armand, M. Forsyth, and L. A. O'Dell, “Ion dynamics in a mixed-cation alkoxy-ammonium ionic liquid electrolyte for sodium device applications,” ChemPhysChem, vol. 17, no. 20, pp. 3187–3195, 2016. View at: Google Scholar
- T. Tamura, K. Yoshida, T. Hachida et al., “Physicochemical properties of glyme–Li salt complexes as a new family of room-temperature ionic liquids,” Chemistry Letters, vol. 39, no. 7, pp. 753–755, 2010. View at: Google Scholar
- Z. P. Visak, “Some aspects of ionic liquids as diverse and versatile sustainable solvents,” Journal of Solution Chemistry, vol. 41, no. 10, pp. 1673–1695, 2012. View at: Google Scholar
- N. V. Plechkova and K. R. Seddon, “Applications of ionic liquids in the chemical industry,” Chemical Society Reviews, vol. 37, no. 1, pp. 123–150, 2008. View at: Google Scholar
- M. O'Meara, A. Alemany, M. Maase, U. Vagt, and I. Malkowsky, “Deposition of aluminum using ionic liquids,” Metal Finishing, vol. 107, pp. 38-39, 2009. View at: Google Scholar
- R. P. Swatloski, J. D. Holbrey, and R. D. Rogers, “Ionic liquids are not always green: hydrolysis of 1-butyl-3-methylimidazolium hexafluorophosphate,” Green Chemistry, vol. 5, no. 4, pp. 361–363, 2003. View at: Google Scholar
- K. Matsumoto, Y. Okamoto, T. Nohira, and R. Hagiwara, “Thermal and transport properties of Na[N(SO2F)2]–[N-Methyl-N-propylpyrrolidinium][N(SO2F)2] ionic liquids for Na secondary batteries,” The Journal of Physical Chemistry C, vol. 119, no. 14, pp. 7648–7655, 2015. View at: Google Scholar
- J. Hwang, K. Matsumoto, and R. Hagiwara, “Na3V2(PO4)3/C positive electrodes with high energy and power densities for sodium secondary batteries with ionic liquid electrolytes that operate across wide temperature ranges,” Advanced Sustainable Systems, vol. 2, no. 5, p. 1700171, 2018. View at: Google Scholar
- C. Y. Chen, T. Kiko, T. Hosokawa, K. Matsumoto, T. Nohira, and R. Hagiwara, “Ionic liquid electrolytes with high sodium ion fraction for high-rate and long-life sodium secondary batteries,” Journal of Power Sources, vol. 332, pp. 51–59, 2016. View at: Google Scholar
- L. G. Chagas, D. Buchholz, L. Wu, B. Vortmann, and S. Passerini, “Unexpected performance of layered sodium-ion cathode material in ionic liquid-based electrolyte,” Journal of Power Sources, vol. 247, pp. 377–383, 2014. View at: Google Scholar
- T. Vogl, C. Vaalma, D. Buchholz et al., “The use of protic ionic liquids with cathodes for sodium-ion batteries,” Journal of Materials Chemistry A, vol. 4, pp. 10472–10478, 2016. View at: Google Scholar
- C. V. Manohar, T. C. Mendes, M. Kar et al., “Ionic liquid electrolytes supporting high energy density in sodium-ion batteries based on sodium vanadium phosphate composites,” Chemical communications, vol. 54, no. 28, pp. 3500–3503, 2018. View at: Google Scholar
- M. P. Do, P. J. Fischer, A. Nagasubramanian, J. Geder, F. E. Kühn, and M. Srinivasan, “Investigation of the electrochemical and thermal stability of an ionic liquid-based Na0.6Co0.1Mn0.9O2/Na2.55V6O16 sodium-ion full-cell,” Journal of the Electrochemical Society, vol. 166, no. 6, pp. A944–A952, 2019. View at: Google Scholar
- J. Song, B. Xiao, Y. Lin, K. Xu, and X. Li, “Interphases in sodium-ion batteries,” Advanced Energy Materials, vol. 8, no. 17, p. 1703082, 2018. View at: Google Scholar
Copyright
Copyright © 2021 Na Zhu et al. Exclusive Licensee Beijing Institute of Technology Press. Distributed under a Creative Commons Attribution License (CC BY 4.0).