Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Aqueous rechargeable aluminum battery – a mini review

Ryohei Moriab
aGreen Science Alliance Co., Ltd, 2-22-11 Obana, Kawanishi city, Hyogo Prefecture 666-0015, Japan. E-mail: moriryohei@fuji-pigment.co.jp; Fax: +81-72-759-9008; Tel: +81-72-759-8501
bFuji Pigment Co., Ltd, 2-23-2 Obana, Kawanishi city, Hyogo Prefecture 666-0015, Japan

Received 30th May 2025 , Accepted 5th August 2025

First published on 8th August 2025


Abstract

Recently, rechargeable aqueous aluminum ion batteries (RAAIBs) have been a promising candidate as the next-generation secondary battery in the rechargeable battery industry owing to its enhanced theoretical specific energy, low cost, and environmental friendliness. The manufacturing cost and price of battery components are very low because they can be prepared in ambient atmosphere and have a simple manufacturing process, which is advantageous compared to other battery types. Furthermore, the raw materials that comprise the battery's components are easily available and not expensive. However, currently its inferior cycle stability precludes real industrial application. In this article, the current progress in development of RAAIBs is briefly summarized based on the type of aluminum salt, including aluminum fluoride, chloride, sulfide, nitride, and others. Additionally, research areas necessary for improving the electrochemical performance of RAAIB will be discussed.


1. Introduction

The rechargeable battery is one of the most important electronic devices for renewable energy systems. Currently, the representative rechargeable battery technology is the lithium ion battery, despite lithium's scarcity in some regions of the world and its ever-increasing price.1 Thus, development of another metal-based rechargeable battery technology is urgently needed. In this regard, aluminum (Al)-based rechargeable batteries are a promising alternative because Al is the third most abundant element on Earth, making Al-based batteries potentially much cheaper than lithium ion batteries. In addition, Al is the most recyclable metal industrially, with a firmly established recycling system.2 The size of aluminum trivalent ions is small (0.535 Å), which is in theory ideal for creating rechargeable batteries with high energy density.3–5 An Al-based redox couple, which involves three electron transfers during electrochemical charge/discharge reactions, provides storage capacity surpassing that of the single electron transfer lithium-ion battery. Compared to its monovalent and divalent counterparts, trivalent Al exhibits the highest theoretical charge capacity per unit volume (8046 mAh cm−3), which is approximately four times that of Li (2042 mAh cm−3, Table 1).6 Although the theoretical gravimetric capacity of Al (2980 mAh g−1) is lower than that of Li (3861 mAh g−1), it is still superior to many other metals. Because of these exceptional characteristics, Al batteries are an attractive candidate for rechargeable batteries in smartphones, computers, electric vehicles (EVs), as well as for smart grids. From an industrial standpoint, large-scale development of Al-based rechargeable batteries offer the possibility of high energy density with low cost, resource stability and safety when coupled with suitable electrolytes and cathode materials.
Table 1 Essential properties of various metal anodes applied in metal ion batteries. Gravimetric and volumetric capacity, abundance in Earth's crust, cost, standard potential and cation radius of lithium (Li), sodium (Na), potassium (K), magnesium (Mg), calcium (Ca), and aluminum (Al) are presented
Element Gravimetric capacity (mAh g−1) Abundance (ppm) Cost (USD per kg) Volumetric capacity (mAh cm−3) Standard potential (V vs. NHE) Cation radius (Å)
Na 1166 22[thin space (1/6-em)]700 3.1 1050 −2.71 1.02
Ca 1340 41[thin space (1/6-em)]000 2.4 2071 −2.87 1
K 685 18[thin space (1/6-em)]400 13.1 609 −2.925 1.38
Mg 2205 23[thin space (1/6-em)]000 2.2 3868 −2.37 0.72
Li 3861 65 19.2 2042 −3.042 0.76
Al 2980 82[thin space (1/6-em)]000 1.9 8046 −1.66 0.535


Large number of research has been studied to develop Al ion batteries (AIBs) and the battery components materials are different from those for lithium ion battery. So far, transition metal oxides such as MnO2, V2O5, TiO2 and WO3 etc… have been investigated as cathode material.7–10 Transition metal sulfides such as MoS2, Co3S4 etc… also have been applied as cathode.11,12 Prussian blue analog,13 organic compound,14 conductive polymer such as polyaniline,15 and carbon based materials such as graphite,16 graphene,17 carbon nanotube18 also have been studied as candidate cathode materials. Apart from AIBs, sulfur also has been investigated as cathode for aluminium sulfur battery.19,20 For anode, not to mention aluminum is the typical anode for AIBs. Aluminum alloy also have been intensively studied and Zn, Cu, Ce, and Sn etc… have been carried out to form intermetallic alloy for AIBs anode.21 Placing a functionalized interfacial layer on the surface of the Al anode is another method to promote stable electrochemical reaction of AIBs. In this regard, amorphous aluminum,22 amorphous carbon,23 Mxene24 etc… were applied as interfacial layer for AIB anode. So far, AlCl3 based electrolyte was especially focused to study for AIBs since AlCl3 demonstrated stable aluminum ion redox reaction. Since AlCl3 possess strong acidity and corrosive nature, molybdenum and niobium were often applied as current collector.25 Furthermore, in general, glass fiber membranes had been used as separator for AIBs.26 In addition, polyacrylonitrile as well as cellulose based separator also have been challenged for AIB separator.27,28 It should be noted that the standard electrode potential of Al3+/Al is −1.67 V vs. SHE.29 The theoretical energy density for AIBs is 1060 Wh kg−1 which is much higher than that of lithium ion battery (406 Wh kg−1).30 Even though, practical capacity ranges up to a little over 400 Ah Kg−1.31

However, certain issues hinder the application of RAAIB. For example, during electrochemical reactions, a thin oxide layer, typically a few nanometer thick, is formed on Al anodes, which behave as a dielectric material (bandgap of 5–8 eV)32 blocking the reversible plating/stripping reaction. To make use of Al as an anode material, reversible aluminum electrodeposition at high coulombic efficiencies is an essential criterion. This can only be achieved in a limited number of electrolyte systems, particularly those employing Al-halide molten salts, including ionic liquids (ILs) and deep eutectic solvent systems. The first report to overcome this problem used an electrolyte composed of AlCl3 and an IL;33 since this discovery, considerable numbers of studies using similar electrolyte systems have been reported. It has also been suggested that battery capacity in many RAAIBs originates from monovalent ions (AlCl4 and Al2Cl7), which do not fully utilize the advantage of Al's ability to exist as a trivalent ion.34 Moreover, AlCl3-based IL electrolytes are highly corrosive and sensitive to moisture, making their application in real batteries challenging.35–42 AlCl3 and IL analogs as deep eutectic solvent-based electrolytes have also been studied for Al-based batteries, although they still possess corrosive characteristics.43,44 On the other hand, organic electrolytes present an appealing choice for the practical application of rechargeable Al batteries. However, a major challenge that has hindered the development of organic electrolytes for rechargeable Al batteries is the lack of a suitable organic system, as most studies using organic electrolytes do not demonstrate high efficiency for practical usage.45,46 Finally, solid state electrolyte-based Al batteries have also been studied although high capacity could not be obtained due to high resistance at the electrode/solid electrolyte interface.47,48

Based on previous research, it appears aqueous-based electrolytes are still the ideal choice. If Al-based, efficient rechargeable batteries with aqueous electrolytes could be developed, it would make the manufacturing process simpler and cheaper compared to that of the lithium ion battery, because these materials can be prepared in an ambient atmosphere and the need for inert gas such as nitrogen or argon is removed. Al-based batteries can even operate in ambient atmosphere unlike lithium ion batteries, providing an alternative choice for use in industrial rechargeable batteries.

In this review, we will explain the basic principles of Al batteries with aqueous-based electrolyte with a few examples classified by the type of aluminum salt in the aqueous electrolyte. In addition, necessary research directions to improve RAAIBs will be discussed.

2. Metal salts in electrolyte

2.1. Aluminum TFSI and (OTF)3

A major problem in RAAIBs is the formation of oxide layers in Al-based rechargeable batteries. To keep charge conduction and avoid oxide layer formation, potentials that exceed the thermodynamic stability of water is necessary, but would result in undesirable hydrogen evolution reaction (HER), thus the electrochemical window of RAAIBs is constrained. Furthermore, applying Al as an anode for rechargeable battery give rise to more issues, including (i) corrosion, (ii) dendrite formation, and (iii) re-passivation (Fig. 1).49
image file: d5ya00148j-f1.tif
Fig. 1 Obstacles at the aluminum and electrolyte interface: illustrated explanation of possible phenomena and chemical species is shown assuming under water based electrolytes with acidic condition.

Until now, Al-based metal salts such as Al(CF3SO3)3 (also known as Al(OTF)3) have been actively investigated due to the fact that they readily dissociate in water to form hexa-coordinated Al-ion complexes.50 Although aqueous electrolytes typically possess narrow electrochemical stability limited by the electrochemical window (thermodynamically as low as 1.23 V) of water, recent studies demonstrated that aqueous electrolytes can stably operate up to >2 V by incorporating highly concentrated aqueous solutions of metal salts containing perfluorinated anions such as TFSI. Molecular dynamics simulations have elucidated that their stability at a high oxidation state is related to water depleted zone formation at the Al anode surface, owing to the accumulation of TFSI anions.51 The anion OTF participates in the Al3+ solvation structure as [Al(H2O)4(OTF)]2+.52,53

Costa et al. investigated the influence of Al(OTF)3 concentration in aqueous electrolyte for RAAIB with carbon as the cathode material.54 As one can see in Fig. 2(a), 5 M of Al(OTF)3 electrolyte results in an unstable charge–discharge profile, with drifting potentials and polarization augmentation (from ∼0.7 V at 20 h to 1.0 V at 180 h), likely due to reduced ionic conductivity and increased viscosity.55 When Al(OTF)3 concentration decreases, polarization values became smaller and charge–discharge curves become more stable. Electrochemical impedance spectroscopy (EIS) was also used to elucidate battery resistance at various reaction times (Fig. 2(b)). Right after the battery has been prepared, inductive loops at low frequencies can be observed, owing to its modification by ion adsorption processes because the boundary between the Al anode and the electrolyte is not well established.56 Resistance dropped as the electrochemical reaction proceeded over 100 hours (Fig. 2c), and increased again with an additional 100 hours of battery operation (Fig. 2d). From these results, it can be deduced that resistance depends heavily on the concentration of Al salt in the electrolyte and electrochemical reaction time. From their study, Al(OTF)3 at 2 M concentration gave the best results with lowest battery resistance, in good agreement with other reports.50,56 It should be noted here that surface modification of the Al anode is also a critical factor that was performed in their study.54 Typical surface modification involves IL treatment of the Al anode, which improves the battery characteristics. We will briefly discuss IL and deep eutectic solvent surface treatment of Al anodes in a later section.


image file: d5ya00148j-f2.tif
Fig. 2 (a) Stability of the electrolyte under charge–discharge current of ±0.01 mA cm−2. (b) Electrochemical impedance spectroscopy (EIS) of the electrolyte before electrochemical reaction. (c) EIS of the electrolyte after 100 h of electrochemical reaction. (d) EIS of the electrolyte after 200 h of electrochemical reaction.

Carbon is a common choice for cathodes in both aqueous and non-aqueous rechargeable Al batteries. Graphite is one of the representative carbon-based materials studied, although their hydrophobicity hinders the access of aqueous electrolyte, resulting in poorer electrochemical performance. To overcome this, graphite modified into reduced graphene oxide (rGO) materials with some hydrophilicity were investigated as a cathode material for aqueous Al ion batteries.57 Fig. 3(a) presents the Raman spectra of rGO in its charged and discharged states; rGO appears to be relatively stable against charge–discharge reactions since no obvious change was observed for the G-band in terms of its frequency, width, and IG/ID (Intensity of G band/D band) values. Presence of the anion OTF in the electrolyte was confirmed by 1251, 1033, and 641 cm−1 bands in Fourier transform infrared spectroscopy (FTIR) measurements (Fig. 3(b)). In addition, the frequency of the C–O stretching mode was red shifted after discharge due to electronic charge transfer.58 The C 1s X-ray photoelectron spectroscopy (XPS) results elucidated the presence of C–O, C[double bond, length as m-dash]O, and O–C[double bond, length as m-dash]O bonding at 286.0, 287.2, and 288.6 eV, respectively. Moreover, the C–Fx bond arising from (OTF)3 can be observed at 292 eV (Fig. 3(c)). The Al 2p XPS spectra presented the binding energy of the Al 2p shell as two peaks, centered at 75.9 eV and 75.0 eV (Fig. 3(d)). The main peak of discharged state is at 75.9 eV, indicating the absence of salt, which can be attributed to the formation of [Al(H2O)6]3+. The band at 75.0 eV is more dominant for the charged state, suggesting a different chemical environment surrounding Al. The 27Al solid-state nuclear magnetic resonance (SS NMR) analysis suggested that the band originates from dimer [Al2(OH)2(H2O)8]4+ and/or trimer [Al3(OH)4(H2O)10]5+.59 Furthermore, it was elucidated that the band at 72.6 eV attributed to metallic Al was absent, indicating no Al deposition and dendrite formation risk on the cathode. In summary, XPS analysis proved that Al ions migrate to the rGO cathode during discharge.


image file: d5ya00148j-f3.tif
Fig. 3 Spectra of rGO cathodes at charged and discharged condition. (a) Raman spectroscopy. (b) FTIR. (c) C 1s XPS. (d) Al 2p XPS. All spectra are normalized by the highest intensity.

To further scrutinize Al ion species in the aqueous state, solid-state magic angle spinning (SS-MAS) NMR analysis was performed.57 In water, the most stable form of Al3+ is [Al(H2O)6]3+. Initially, intact Al(OTF)3 was measured with NMR under Ar atmosphere to compare against Al(OTF)3 exposed to moisture over various durations, as Al(OTF)3 is extremely hygroscopic. The 27Al spectrum of pure Al(OTF)3 presented a clear resonance at −8.3 ppm (Fig. 4a). When Al(OTF)3 was solvated in water by exposure to ambient atmosphere for 30 min, another weak signal at 1.2 ppm was observed. When Al(OTF)3 was exposed to ambient atmosphere for 1 day, a significant peak at 1.4 ppm emerged. Simultaneously, the peak at −8.3 ppm shifted by 0.2 ppm and its intensity was attenuated. These changes are likely due to newly formed species [[Al(OH)n(H2O)6−n](3−n)+] (where n = 0–2). It was suggested that the 1.4 ppm signal is characteristic of the most stable species, [Al(H2O)6]3+, whereas the −8.3 ppm peak originated from Al hexa-coordinated oxygen donor ligands.60 1H SS-MAS NMR was performed for further confirmation (Fig. 4b). It was interesting to see signals from intact Al(OTF)3 because hydrogen is not contained in the chemical formula. A broad peak at 0–15 ppm and two prominent peaks at 5.0 and 7.2 ppm were observed. This was ascribed to residual water and OTF from the synthesis procedure of Al(OTF)3. As solvation proceeded, bands became narrower and both peaks were shifted. At this stage, different complexes of (semi-) hydrated aluminum or hydronium species (H3O+) can be assigned. After 24 hours, one clear peak emerged at 6.0 ppm, which originates from [[Al(OH)n(H2O)6−n](3−n)+] (where n = 0–2) as similarly observed in the 27Al NMR measurement. NMR measurement was also performed on the rGO cathode (Fig. 4c). Before any electrochemical reaction, no peaks were observed in the 27Al NMR measurement for the rGO cathode, confirming no elemental traces of Al on the cathode before any electrochemical reaction. After a charge and discharge sequence, peaks and shoulders at 1.8 and 6.4 ppm were observed. Similarly to Fig. 4(a), peaks at 1.8 ppm can be attributed to the monomer [Al(H2O)6]3+, while shoulders at 6.4 ppm are attributed to the formation of dimers and/or trimers of the type [Al2(OH)2(H2O)8]4+ and [Al3(OH)4(H2O)10]5+.61


image file: d5ya00148j-f4.tif
Fig. 4 SS MAS A(Solid State Magic Angle Spinning) NMR spectra offal(OTF)3, freeze dried and the cycled rGO cathodes. (a) 27Al spectra of the Al(OTF)3 salt. (b) 1H spectra of the Al(OTF)3 salt. (c) 27Al spectra of the rGO at each condition. (d) 1H spectra of the rGO at each condition.

1H NMR spectra demonstrated that the rGO cathode had a broad band extending from −10 to 20 ppm centered at 6.3 ppm, confirming that the rGO cathode contains functional groups such as hydroxyls and carboxyls, as well as bound water retained from the freeze dry process in rGO cathode preparation. Peaks at 2.1 ppm were detected for all intact and electrochemically cycled rGO samples and are suggested to originate from aliphatic carbons at the edge of the graphene basal planes. Furthermore, the 4.4 ppm peak corresponds to the protons in the solvated Al monomer, [Al(H2O)6]3+.

So far, we have investigated the details of Al(OTF)3 solvation during RAAIB operation. This leads to discussion of studies to improve battery performance with various electrolyte mixtures. Li et al. prepared mixed electrolyte composed of 5 M Al(OTF)3 and 0.5 M LiOTF for an aqueous Al battery,62 which was able to reduce corrosion caused by hydrogen evolution. When LiOTF was added to Al(OTF)3, the solvation sphere of Al3+ changes from a fully octahedral structure to partly tetrahedral and octahedral, decreasing the complexes’ hydrogen bonding toward water molecules, which reduces hydrogen evolution. As a result, the stability of the electrochemical window was improved to 3.8 V, with an initial charge capacity of 437 mAh g−1 and 103 mAh g−1 after 100 cycles.

Ran et al. attempted to use Al-based alloys for Al aqueous rechargeable batteries.63 They prepared lamella-nanostructured eutectic Al82Cu18 alloy anodes, improving electrochemical stability by preventing Al dendrite growth during stripping/plating cycles. AlxMnO2 was used for the cathode and the prepared battery demonstrated stable Al stripping/plating for 2000 h with an overpotential smaller than ±53 mV. The initial discharge capacity was ∼400 mAh g−1 and 83% capacity was maintained after 400 cycles. This improvement was achieved based on utilizing the difference in corrosion potential between α-Al lamellas and Al2Cu lamellas to obtain low overpotentials, in which α-Al transferred Al3+ charge carriers to Al2Cu.

Other studies have focused on cathode materials. Wu et al. used AlxMnO2·nH2O synthesized by an in situ electrochemical transformation reaction as a cathode material with Al(OTF)3–H2O electrolyte, which exhibited high specific capacity and energy density of 467 mAh g−1 and 481 Wh kg−1, respectively. To scrutinize the origin of this high performance, combined experiments with different conditions were performed with spinel Mn3O4 or AlxMnO2·nH2O as cathodes, and the three different types of electrolyte: Al(OTF)3–H2O, HOTF (triflic acid), and AlCl3/[BMIM]Cl.64 The HOTF-based electrolyte was examined to clarify the effect of Al3+, and AlCl3/[BMIM]Cl was applied as a comparison to an IL-based electrolyte. In addition, to examine the possibility of H3O+ influencing the RAAIB capacity, coin cell structures of carbon fiber paper (CFP)/HOTF–H2O/Mn3O4 and CFP/HOTF–H2O/AlxMnO2·nH2O were prepared (Fig. 5(a)). The concentration of H3O+ was adjusted to be the same as Al(OTF)3–H2O based electrolyte a precise comparison. Fig. 5b–d show that both CFP/HOTF–H2O/Mn3O4 and CFP/HOTF–H2O/AlxMnO2·nH2O exhibited no discharge capacity, implying that all discharge capacity arose from Al3+ intercalation and H3O+ does not contribute to electrochemical performance at all.


image file: d5ya00148j-f5.tif
Fig. 5 The schematic figure of each experiments. (a) Experiment is performed with AlxMnO2·nH2O cathode and Mn3O4 anode with aqueous and ionic liquid electrolytes. (b) and (d) Charge and discharge curves at each experimental condition. (c) and (e) Discharge capacity t each experimental condition.

2.2. Aluminum chloride

Aluminum chloride (AlCl3) is often used as an aluminum salt for IL-based electrolytes, although they suffer from high material costs as ILs are expensive in general. In this regard, some research based on aqueous electrolytes with AlCl3 have been performed. Das et al. had demonstrated that graphite is capable of reversibly hosting Al3+ ions and was combined with TiO2 in 1 M of AlCl3 aqueous electrolyte (Fig. 6(a and b)).65 The battery demonstrated an average discharge voltage of 1.5 V with over 1000 cycles of charge–discharge stability. Cyclic voltammetry data of the prepared Al-ion battery is shown in Fig. 6(c), showing a pair of cathodic and anodic redox peaks (indicated as A and B) at 0.4 and 0.5 V (vs. Ag/AgCl), respectively. Moreover, Fig. 6(d) shows the charge–discharge curve at an applied current of 0.05 mA cm−2, where a very small charge–discharge plateau at approximately 0.5 and 0.4 V (vs. Ag/AgCl), coinciding with the results of cyclic voltammetry, can be observed.
image file: d5ya00148j-f6.tif
Fig. 6 (a) Schematic image of an aqueous aluminum ion battery with graphite as the cathode and TiO2 as the anode. (b) XRD pattern of graphite cathode. (c) Cyclic voltammetry at a scan rate of 2.5 mV s−1. (d) Charge–discharge curves of the graphite cathode in a 1 M AlCl3 aqueous electrolyte at a specific current of 0.05 mA cm−2.

Since AlCl3 is strongly corrosive, Cakan et al. investigated the influence of current collectors on the Al-ion battery with 1.0 M of AlCl3 aqueous electrolyte. Four types of current collectors, namely nickel foil, titanium foil, graphite plate, and stainless steel, were examined and corrosion tests, charge–discharge, and cyclic voltammetry measurements were performed.66 Results indicate graphite as the optimal choice to minimize corrosion and HER. The anatase phase TiO2 was applied as the cathode material with graphite plate, yielding initial discharge capacity of 249 mAh g−1 at a current density of 3 A g−1.

Besides aluminum, other elements were used in a metal salt for Al-ion battery electrolytes. Tao et al. added iron to AlCl3-based electrolyte to form a hydrolyzed, polymerized Al–Fe hybrid electrolyte (poly aluminum ferric chloride, PAFC), which enhanced electrochemical performance.67 The prepared electrolyte contributes to the capacity through an iron redox reaction, proceeding in parallel with aluminum redox when an Al–Fe alloy was used as an anode. This type of hybrid Al–Fe ion battery demonstrated 35 Ah L−1 under 1.0 mA cm−1 of current density with 90% capacity retention after 500 cycles. This special hydrolyzed PAFC was synthesized by the polymerization reaction of aluminum chloride and iron chloride and applied as an electrolyte. To investigate this hybrid metal redox reaction mechanism, FTIR and ex situ Raman spectra of the electrolyte were measured and analyzed. The electrolyte specimens were collected at different stages of the electrochemical reaction as exhibited in the time–voltage curve (Fig. 7(a)). The peak at 314 cm−1 in the Raman spectra is attributed to Fe ions in Fig. 7(b).68 The narrow blue zone in Fig. 7(b) was expanded for easy recognition of peak change as presented in Fig. 7(c). The peak decreased as the discharge reaction proceeds from 0.9 to 0.2 V, implying reduction of ferric ions into ferrous ions. On the other hand, from 1.10 to 1.65 V during the charging process, the peak profile was similar to that of open-circuit voltage state. This phenomenon proves that oxidation/reduction of iron ions is highly reversible, which is the main reason for strong electrochemical performance. Fig. 7(d) is the ex situ FTIR spectra of the prepared electrolyte. Peaks at approximately 1610–1630 cm−1 are due to H–OH bonding, shifting to frequencies of 590–601 cm−1 and 3460–3490 cm−1, indicating the polymerization of hydroxyl aluminum and hydroxyl iron. This study also paved the way for developing high performance Al-ion batteries by applying hybrid-ion electrolytes.


image file: d5ya00148j-f7.tif
Fig. 7 Ex situ electrochemical measurements of PAFC electrolyte at different charge–discharge stages. (a) Potential–time curve during redox reaction. (b) Raman spectra of the electrolyte at respective charge and discharge voltage conditions. (c) Magnified Raman spectra of the blue zone. (d) FT-IR spectra of various electrolyte including PAFC.

Wills et al. prepared an Al-ion battery with CuHCF (copper-hexacyanoferrate) as the positive electrode and TiO2 as the negative electrode.69 The electrolyte was composed of 1 M of AlCl3 and KCl. Galvanostatic and cyclic voltammetry analysis revealed that battery capacity was approximately 10 mAh g−1 with an average discharge voltage of approximately 1.5 V. Interestingly, the electrode was dark blue–grey in the charged state when observing the TiO2 on Al anode; this subsequently reverts to the original white color of TiO2 in the discharged state. It was suggested that the TiO2 color variation could be ascribed to the reduction of Ti4+ to Ti3+, driven by possible reactions of Al3+ intercalation, Al3+ surface adsorption, or H+ intercalation and H+ surface adsorption.

Rastabi et al. investigated the metallurgical Al anode behavior in a water-in-salt electrolyte.70 AlCl3·6H2O was applied as the metal salt for an aqueous electrolyte and graphite was used as the cathode. The capacity decay from charge–discharge cycling was found to be due to the formation of a detrimental solid–electrolyte interphase (SEI) layer on the Al anode surface. Detailed observation clarified that Al corrosion initiates around the Al3Fe intermetallic phase, as Al3Fe particles allow the electrolyte to break the oxide barrier and penetrate into bulk Al. The Al–Fe phase diagram indicates that iron is a normal impurity element for Al and its solubility is less than 0.05%, often resulting in iron-containing intermetallic phases, such as Al3Fe.71,72 It should be noted that such impurity precipitation is extremely difficult to eliminate and Al3Fe exists even in Al that is over 99.9% pure. In addition, the cost of extremely high purity Al of over 99.99% is considerably more expensive and not realistic for application in Al-ion batteries for industries.73 This study demonstrates that metallurgical study of Al anodes is also important for creating feasible options for aqueous Al-ion batteries.

Xu et al. prepared an RAAIB with the MnO2 cathode with good cycling properties, optimizing Al3+ solvation structures by changing electrolyte concentration to prevent Al3+ intercalation.74 By combining molecular dynamics simulations with results obtained from experimental analysis, it was found that the solvation size of Al3+ in relatively concentrated electrolyte is wider than MnO2 lattice spacing, suggesting that Al3+ cannot enter into the MnO2 cathode material (Fig. 8). Namely, the solvation structure and the size of Al3+ cations are larger in the preferably modified saturated electrolyte (s-Al/Mn electrolyte composed of 4.4 M AlCl3 and 1 M MnCl2 in H2O) than in the diluted electrolyte (d-Al/Mn electrolyte composed of 1 M AlCl3 and 1 M MnCl2 in H2O). As a result, self-intercalation with the MnO2 cathode lattice could be suppressed due to the larger size of Al3+ in the optimized saturated electrolyte. Owing to this optimized electrolyte concentration effect, prepared RAAIBs achieved a specific capacity of ∼493 mAh g−1 over 1000 cycles.


image file: d5ya00148j-f8.tif
Fig. 8 Schematic image presenting Al anode corrosion as electrochemical reaction proceeds. (a) Intact Al anode. (b) After contacting with ionic liquid based electrolyte. One can observe the corrosion around Al3Fe particles. (c) After contacting with aqueous based electrolyte. One can observe the corrosion around Al3Fe particles and solid electrolyte interface layer is formed on unbroken oxide film.

With respect to chloride-based Al complex metal salts, Al(ClO4)3·9H2O has been attempted in electrolytes for Al ion batteries. Even though Al(ClO4)3·9H2O is a chloride-based Al complex, its non-corrosive characteristics compared to AlCl3 is the big advantage for industrial application.75 It should be noted that Al(ClO4)3·9H2O was combined with succinonitrile to form a hydrated eutectic electrolyte and not an aqueous one. Author's group also have applied Al(ClO4)3·9H2O as the metal salt in the electrolyte and successfully prepared an Al ion battery.76,77

2.3. Aluminum nitride, sulfide

There have been some attempts to apply Al(NO3)3 as a metal salt in electrolytes for Al ion batteries. Nanocubic structured potassium cobalt hexacyanoferrate (K2CoFe(CN)6) synthesized by a hydrothermal method was used as a cathode successfully in an RAAIB with a capacity of 50 mAh g−1 at 0.1 A g−1 with 1 M of Al(NO3)3 as an electrolyte.78 Wang et al. had prepared a flexible RAAIB with a MoO3 anode coated with polypyrrole (PPy) and a copper hexacyanoferrate cathode, with 1 M Al(NO3)3 as an electrolyte. The prepared cell demonstrated high stability with capacity retention of 83.2% for 100 cycles, due to high conductivity from deposited PPy on MoO3. This battery demonstrated good safety and flexibility because a separator material composed of a PVA–Al(NO3)3 gel polymer was used. Thus, the battery exhibited high durability against various mechanical stress forces and deformation such as folding, bending, squeezing, twisting, and drilling, which paves the way for application of flexible Al-ion batteries in wearable electronic devices.79 Lahan et al. also applied MoO3 to prepare an aqueous Al-ion battery to examine wether Al3+ ions intercalate into MoO3 in aqueous-based electrolytes.80 It was confirmed that AlCl3 aqueous electrolyte demonstrated better electrochemical performance than Al2(SO4)3 and Al(NO3)3 aqueous electrolytes due to factors such as higher storage capacity of Al3+ ions, retention of battery capacity, and long-term stability. Wu et al. studied the influence of IL surface modification and Al alloying (Al–Cu alloy or Zn–Al alloy) in an aqueous electrolyte with 1 M of Al(OTF)3, AlCl3, and Al(NO3)3. They observed gas evolution and corrosion during the electrochemical redox process. The anode activity of the aluminum and the surface treated aluminum with IL was mainly evoked by chloride anions or solid–electrolyte interface layer residue. It was also found that when an Al–Cu alloy was applied as an anode, nanolamellar Al2Cu worked as the catalyst and maintained hydrogen generation. They also confirmed that this gas evolution effect was even more pronounced in the case of a Zn–Al alloy anode.81

Al2(SO4)3 was applied as an electrolyte and the electrochemical performance was compared to an aluminum bis(trifluoromethanesulfonyl)-imide (Al[TFSI]3) electrolyte. The Zn-supported Al alloy was used as an anode and MnO2 as the cathode. The Zn–Al alloy reduces the detrimental effect of surface passivation, which contributed to accelerating Al3+/Al redox reaction.82 Furthermore, electrodeposited substances on the Zn–Al alloy were mainly composed of uniform nanostructures that tend not to form dendrites, which contributed to excellent cycle performance. However, Al[TFSI]3 exhibited an overall better electrochemical performance compared to the Al2(SO4)3-based electrolyte.82

Zhao et al. had reported that certain types of IL treatment on Al anodes formed SEI, which effectively improved electrochemical performance in aqueous electrolyte. MnO2 was used as a cathode and the prepared Al-ion battery demonstrated specific energy of approximately 500 Wh kg−1.83 The composition and structure of SEI formed by AlCl3-based IL electrolyte were investigated. Fig. 9(a) shows the result of the total reflection ATR-FTIR data of AlCl3-based IL electrolyte-treated electrode (TAI electrode). The presence of a few organic functional groups can be observed. C[double bond, length as m-dash]H groups were confirmed in the pristine ionic liquid, as well as additional C[double bond, length as m-dash]O groups and the enhancement of the C[double bond, length as m-dash]C functional group in the TAI electrode. It should be noted here that infrared vibration modes that arise from the imidazolium ring (1100–600 cm−1) in the IL disappeared after IL treatment, proving that imidazolium was incorporated into SEI composition. Fig. 9(b–d) presents the XPS spectra of Al, TAI, and pure Al, where an obvious peak originates from Al2O3. On the other hand, the Al2p spectra of the TAl substrate shifted towards higher binding energy, which can be ascribed to Al salt, such as AlCl3. Cl2p XPS spectra also support that the 199 eV peak is consistent with a valence of −1 (Fig. 9(c)). The N1s spectra imply that an abundance of nitrogen-containing species exist in the TAI electrode and the 401 eV peak is ascribed to imidazole ions (Fig. 9(d)). Morphology data revealed that surface topography of the TAI electrode is smoother than that of the pure Al surface (Fig. 9(e and f)). The scanning electron microscopy (SEM) cross-section view clearly shows an interfacial layer that is composed of Al, Cl, and N (Fig. 9(g)). It was also demonstrated that even though the aqueous Al(OTF)3)3-based electrolyte demonstrated better electrochemical performance, the Al2(SO4)3 aqueous electrolyte showed rechargeable performance after IL treatment of the Al anode.


image file: d5ya00148j-f9.tif
Fig. 9 Surface analysis of metal Al anode before and after treatment with AlCl3-IL electrolyte. (A) ATR-FTIR spectra of Al and TAI surface. XPS spectra of (B) Al 2p (C) Cl 2p (D) N 1s on Al anode (red line) and TAl anode (blue line). SEM pictures of (E) Al anode (F) TAl anode. The insets of (E) and (F) are digital image of Al and TAl (G) Cross-sectional SEM picture of TAl anode and corresponding EDX mapping of Al, Cl, and N. Al: Intact Al anode, TAI: Al anode surface treated with ionic liquid based electrolyte. a.u., arbitrary units.

Furthermore, Chunshuang et al. succeeded in shifting the reduction potential for aluminum deposition by applying amorphous aluminum anode. It was found that the amorphous structure decreased the nucleation energy barrier of aluminum anode which accelerated the interfacial ion transfer kinetics and obtained enhanced aluminum plating/stripping reaction with Al2(SO4)3 electrolyte.84

2.4. Aluminum iodide, bromide and aluminum PF6, BF4

To the best of our knowledge, there is not much research on RAAIB using aluminum iodide, bromide, PF6, or BF4. Instead of utilizing iodide-based electrolyte, iodine-containing nitrogen-doped microporous carbon polyhedrons have been used as an electrode. The prepared Al-ion battery exhibited high specific capacity and high rate performance of 102.6 mAh g−1 at 8 A g−1, with good cycling stability. LiTFSI- and AlCl3-based aqueous electrolytes were used, although iodide from the cathode participated in the reversible redox reaction between I2 and I, with I3and I5 as intermediates.85 Tian et al. also applied iodide as a cathode material by combining it with PVP, conductive carbon, and polytetrafluoroethylene as the binder, with the prepared battery demonstrating a high capacity of over 200 mAh g−1 at 0.2C based on I3−/I redox chemistry.86

Although not a water-based electrolyte, the electrochemical reaction of Al3+/Al was examined in AlBr3 dissolved in ethylbenzene for rechargeable Al ion batteries. It was found that KBr was necessary as an additive to promote Al deposition/dissolution reactions on glassy carbon and platinum electrodes.87

Another example of a non-aqueous electrolyte uses (Al(PF6)3) dissolved in dimethyl sulfoxide (DMSO) to investigate aluminum deposition and reversibility. Al deposition-stripping was confirmed with SEM and XPS with depth profiling analysis.88 The redox reaction was further accelerated by addition of Et3Al. It was found that continuous cathodic decomposition of DMSO resulted in Al2O3 formation on the electrode surface. To the best of our knowledge, there is no studies on RAAIBs using tetrafluoroborate (BF4)-based electrolytes.

3. Critical research areas

Having discussed above, there still are many issues such as charge–discharge cycling stability that need to be solved before RAAIBs can be applied industrially. At least cycling stability needs to be maintained for over 200 cycles for drones and over 3000 cycles for EV for example.

It seems that fluoride- or chloride-based aqueous electrolytes demonstrate better electrochemical performance overall compared to other types of electrolytes (Table 2). Some research challenges have been explained above although the following are some necessary research topics that need to be investigated to make RAAIB technology feasible.

Table 2 Summary of component materials and their properties discussed in this study. All of electrolyte are aqueous based except ref. 86–88
Electrolyte Anode Cathode Initial capacity Capacity stability Ref.
5 M of Al(OTF)3 Al Expanded graphite 10 mAh g−1 50% after 200 cycles 54
5 M of Al(OTF)3 Al Reduced graphene oxide (rGO) 136 Wh kg−1 60% after 200 cycles 57
5 M Al(OTF)3 and 0.5 M LiOTF Al MnO 437 mAh g−1 103 mAh g−1 after 100 cycles 62
2 M Al(OTF)3 Al82Cu18 alloy AlxMnO 400[thin space (1/6-em)]mAh g−1 83% after 400 cycles 63
5 M of Al(OTF)3 Al AlxMnO2·nH2O 481 Wh kg−1 272[thin space (1/6-em)]mAh g−1 after 60 cycles 64
1–5 M of Al(OTF)3 Al82Cu18 alloy Tungsten   Charge/discharge cycling and a current density of 0.2 mA cm−2 with voltage gap of 2.6 V 81
1 M of AlCl3 TiO2 Graphite 62.5 mAh g−1 25 mAh g−1 after 1000 cycles 65
1 M of AlCl3 Al TiO2 20 mAh g−1 39.5 mAh−1 after 20 cycles 66
Poly aluminium ferric chloride Fe–Al alloy Nickel hexacyanoferrate (NiFe-PBA) 35 Ah L−1 90% capacity retention after 500 cycles 67
1 M of AlCl3 and KCl TiO2 CuHCF (copper-hexacyanoferrate) 15 MWh g−1 70% capacity retention after 1750 cycles 69
2 M of AlCl3·6H2O Al Nano graphite 400 mAh g−1 40 mAh g−1 at cycle 100 70
4.4 M AlCl3 and 1 M MnCl2 in H2O Al MnO2 493 mAh g−1 Stable over 1000 cycles 74
1 M of AlCl3 Al MoO3 680 mAh g−1 170 mAh g−1 after 20th cycle 80
Al(ClO4)3·9H2O in succinonitrile to form a hydrated eutectic electrolyte Al Polyaniline 185 mAh g−1 Stable over 300 cycles 75
1 M of Al(NO3)3 Al Cobalt hexacyanoferrate (K2CoFe(CN)6) 50 mAh g−1 76% Capacity retention after 1600 cycles 78
1 M of Al(NO3)3 Polypyrrole coated MoO3 Copper hexacyanoferrate (CuHCF)   83.2% for 100 cycles 79
2 M of Al2(SO4)3 Zn–Al MnO2 290 mAh g−1   82
1 M of Al2(SO4)3 Ionic Liquid Treated Al MnO2 500 Wh kg−1 Stable capacity retention after 40 cycles 83
0.5 M of Al2(SO4)3 Amorphous Al AlxMnO2 780 mAh g−1 180 mAh g−1 after 80 cycles 84
9 M LiTFSI + 1 M AlCl3 Al MOF derived N-doped microporous carbon polyhedrons 219.8 mAh g−1   85
AlCl3 and 1 ethyl 3 methylimidazolium chloride (EMIC) Al PVP, I2 and conductve carbon 200 mAh g−1 High stability for over 150 cycles 86
1 M AlBr3 and 0.5 M KBr in ethylbenzene Al Glassy carbon 0.4 mAh g−1 Confirmed over 120 cycles 87
0.25 M Al(PF6)3 in dimethyl sulfoxide Al Platinum   Confirmation of Al deposition–stripping 88


3.1. Electrolyte mixture. Water base + solvent, alcohol, ionic liquid

Electrolyte composition and utilization of additives in water-based electrolytes is one of the strategies that can improve the electrochemical performance of RAAIBs.

Lahiri et al. studied how the addition of ILs to water-based electrolytes influences Al solvation dynamics by density functional theory (DFT) simulations as well as by experiments.89 DFT calculations elucidated that the coordination number of solvated Al in aqueous (Al(TfO)3) electrolytes could be lowered by adding 1-ethyl-3-methylimidazolium trifluoromethylsulfonate (EMIMTfO), which would result in change of the Al deposition/stripping behavior on the Zn/Al alloy anode. Interestingly, IL addition reduces strain on the manganese oxide cathode during intercalation/deintercalation reactions, which also contributed to better electrochemical performance. FTIR spectroscopy in Fig. 10(a and b) indicates that IL addition induced a shift of CF3 vibration to lower wavenumbers, which could be related to the change in Al coordination. In the range of between 2000 and 3600 cm−1 as presented in Fig. 10(c), additional peaks originating from the imidazolium cation were observed when ionic liquids are introduced.90 To further investigate the influence of EMIM + TfO ions’ addition to aqueous electrolytes, DFT-based quantum chemical calculations were performed. Payard et al. had previously showed that TfO ions would have monodentate or bidentate coordination with Al3+ ions.91 Other DFT calculations and experimental studies also report that six water molecules exist in proximity of the Al(TfO)3 molecule.92 Thus, the relaxed geometry of Al(TfO)3·6H2O was constructed, in which six water molecules are located around Al(TfO)3, as shown in Fig. 10(d). Formation of hydrogen bonds between water molecules and the TfO ions was also confirmed. The IR spectrum of the relaxed structure of Al(TfO)3·6H2O based on theoretical calculations is presented in Fig. 10(e) and the wavelength of peaks agree with the experimental spectrum. These results suggest that the interaction between TfO ions and Al3+ ions did not change when the six solvating water molecules were introduced to Al(TfO)3. They also studied the introduction of EMIM+ and TfO ions to Al(TfO)3·6H2O (data not shown). It was found that AlOH species are formed by deprotonation when water molecules coordinate with Al3+ ions. Furthermore, the Al3+ solvation environment is changed due to H3O+ species from the addition of EMIM+ and TfO ions, resulting in altered deposition/stripping behavior of Al.


image file: d5ya00148j-f10.tif
Fig. 10 FT-IR spectra of AlTfO at each concentrations of EMIMTfO. (a) 900 and 1400 cm−1. (b) 740 and 780 cm−1. (c) 2000 and 4000 cm−1. (d) Al(TfO)3·6H2O Geometries with all the TfO ions are being bi-coordinated to the Al3+ ion. (e) Comparison of experimental and simulated IR spectra of AlTfO in water. (f) Relaxed geometries of Al(TfO)3·6H2O with EMIMTfO.

3.2. Surface treatment

Surface treatment of the Al anode is also a strategy for improvement. As explained above, SEI formation by ILs or deep eutectic solvents could alter the Al anode surface chemistry, yielding stable cell cycling stability even in aqueous electrolytes.83 Besides ILs, deep eutectic solution can be applied in aluminum anode surface treatment to form SEI.93 SEI is expected to promote ion transfer, eliminate the passivating oxide layer, create a stable solid protective layer, and improve compatibility between the aqueous electrolyte and the Al anode to stabilize electrochemical performance at the interface. The RAAIB prepared with a surface-treated Al anode, an FeFe(CN)6 cathode, and 2 M Al(OTF)3 as an electrolyte demonstrated a discharge capacity of 85 mAh g−1 with a capacity retention of 58 mAh g−1 after 150 cycles. Average coulombic efficiency was 97.1% under a 100 mA g−1 current density. Since surface treatment is a relatively easy process, it can feasibly be applied industrially for Al-ion battery commercialization.

3.3. Al alloys

Use of aluminum alloys is another effective option to enhance RAAIB performance. Two of the main problems of Al-ion batteries is the oxide layer (Al2O3) on the Al anode and side reactions of hydrogen generation. As explained above, a few studies have applied Al–Cu and Zn–Al alloys as anodes for Al-ion batteries.81 It was also found that the Al–Cu alloy lamellar hetero-structures (E-Al82Cu18) enhanced Al-ion electrochemical reversibility by achieving dendrite-free Al deposition during redox cycles.63 Induced by an in situ eutectic solidification reaction, E-Al82Cu18 possess an aligned lamellar structure consisting of intermetallic Al2Cu nanolamellas and monometallic α-Al. These two phases pair with each other to give localized galvanic couples of Al/Al2Cu in periodic structures.

Jia et al. studied Sn as an alloying element owing to its good compatibility with Al. They applied a rolling procedure to increase the contact interface area between Al and Sn to create Al–Sn laminate electrodes. This Al–Sn hetero-structure reduced interface resistance and promoted Al stripping/plating.94 When AlxMnO2 cathode was applied, battery delivered a specific capacity of 177 mAh with a stable discharge voltage plateau of 1.5 V. To further ameliorate the electrochemical properties of Al–Sn laminate electrode, a polymer coating was used to reduce HER. Combining this type of anode with a KNHCF cathode, the prepared Al-ion battery maintained 82% capacity retention even after 700 cycles. The author's group also studied the effect of chemical composition of the Al anode especially respect to magnesium impurity, and succeeded in improving the electrochemical performance of RAAIBs.76

3.4. Hybrid salts for aqueous electrolytes

A considerable amount of research has been conducted by applying hybrid salt-based electrolytes for other types of metal-ion batteries, including lithium-, zinc-, and magnesium-based batteries.95–97

For RAAIBs, a major problem is that the stripping/plating potential of Al3+/Al (−1.68 V) is lower than the hydrogen evolution potential of H+/H2. On top of that, the Al anode can be readily corroded under acidic conditions of the aqueous electrolyte. Thus, searching for electrolytes that can solve these issues is essential. Addition of lithium bis(trifluoromethane sulfonyl)imide (LiTFSI) to AlCl3 electrolytes was effective to suppress Al anode corrosion and hydrogen generation. Al-ion batteries with the electrolyte composition of 0.5 M of AlCl3 and 12 M of LiTFSI with a vanadium potassium cathode was prepared, exhibiting 223 mAh g−1 at 1000 mA g−1 with good stability, maintaining a capacity of 64.6% after 300 cycles.98 Structural and electrochemical analysis also revealed that a reversible Al3+, Li+, and H+ intercalation/de-intercalation reaction occurred in the vanadium potassium cathode during redox reaction.

Gao et al. also investigated the effect of an additional second salt to enhance electrochemical performance and to widen the electrochemical window.99 They attempted a bi-salt aqueous electrolyte based on aluminum-lithium trifluoromethanesulfonate (OTF) dual salt, which resulted in a wide electrochemical window of 4.35 V with a low over-potential of 14.6 mV. AlxMnO2·nH2O was used as the cathode. It was revealed that the bi-salt electrolyte reduces Mn dissolution, which suppresses self-discharge based on the electrode–electrolyte interface concentrated confinement effect. As a result, the prepared Al-ion battery exhibited a capacity of 364 mAh g−1 with a capacity decay rate of 0.37% per cycle over 150 cycles. Furthermore, Yuan et al. developed aluminum ammonium hybrid battery with aqueous electrolyte composed of AlCl3 and NH4Cl and demonstrated 89.3 Wh kg−1 capacity with high stability over 10[thin space (1/6-em)]000 cycles.100 They have applied prussian blue analogue cathode (K1.14FeIII[FeII(CN)6nH2O) which demonstrated excellent stability.

Wen et al. prepared silicon based inorganic/organic hybrid hydrogel by combining 2 M of Al(OTF)3/H2O aqueous solution and TEOS (tetraethyl orthosilicate) by sol–gel polymerization method. With cathode composed of potassium nickel hexacyanoferrate, this hydrogel electrolyte demonstrated a high ionic conductivity of 2.9 × 10−3 S cm−1 and delivered initial discharge capacity of 74.9 mAh g−1 with capacity retention of 90.0% after 200 cycles.101

3.5. Cathode

Another problem that Al ion batteries face is that once Al3+ ions are inserted, extracting them from the cathode host is difficult due to strong electrostatic interactions between Al3+ ions and the cathode lattice. This irreversible reaction would induce volume expansion and unfavorable structural change. Acidic electrolytes may be necessary to extract Al3+ from the cathode, although the acidity would damage and dissolve the cathode's structure. Discussion of cathode choice for aqueous Al-ion batteries can also be accessed in some review studies.102

δ-MnO2 has been reported as a prime candidate for cathode material in aqueous Al-ion batteries owing to its layered crystalline structure combined with large interlayer distance. However, they tend to collapse when high ionic potential multivalent ions such as Al3+ intercalates into those layers. In this regard, Chen et al. revealed that when heterogeneous metal dopants are introduced into layered crystals, dopants bond to oxygen to make the δ-MnO2 structure stronger. When δ-MnO2 is doped with vanadium, the Al-ion battery using 2 M Al(OTF)3 electrolyte demonstrated a specific capacity of 518 mAh g−1 at 200 mA g−1 with stable cell cycling for 400 cycles (Fig. 11(a)).103 Electrochemical impedance spectroscopy (EIS) analyses were performed to examine the resistance of δ-MnO2 cathodes (Fig. 11(b)). Necessary battery resistance values can be obtained by fitting semicircles at the high frequency region and slope lines in the low frequency region, together with the equivalent circuit (Fig. 11(b) inset). R1 in the circuit is the resistance of the electrolyte and R2 is the charge transfer resistance as (Rct). Fig. 11(c) indicates the Rct fitting values, with vanadium-doped δ-MnO2 cathodes showing the lowest resistance. It was suggested that heterogeneous metal doping in δ-MnO2 can enhance electrical conductivity and reduce Rct by creating additional pathways for Al3+ ion conduction and active sites.104 Rate capacity performance is shown in Fig. 11(d) and vanadium-doped δ-MnO2 demonstrated the strongest rate performance even against high current densities. Cycling stability is presented in Fig. 11(e and f), with vanadium-doped δ-MnO2 delivering the favorable cycling performance at both current densities of 200 and 500 mA g−1.


image file: d5ya00148j-f11.tif
Fig. 11 Various electrochemical measurements of intact and doped δ-MnO2 cathodes. (a) 1st charge–discharge curves at a current density of 200 mg−1. (b) Cole–Cole plots of V-δ-MnO2 and δ-MnO2 electrodes (Inset presents the equivalent circuit). (c) The charge transfer resistance of intact and doped δ-MnO2 cathodes. (d) Rate performances of each δ-MnO2 cathodes. (e) and (f) Cycling performances under 200 and 500 mA g−1, respectively.

Besides MnO2, nanobelt-structured MoO3 was synthesized by a hydrothermal method and used as a cathode. The high number of active sites of MoO3 promoted Al3+ intercalation/de-intercalation reactions and increased the Al3+ diffusion rate. This MoO3 nanobelt cathode-based Al-ion battery displayed 267.61 mAh g−1 at 0.2 Ag−1 and capacity was maintained even after 100 cycles.105 Nanostructured Mn3O4 has also been applied as cathodes for aqueous Al-ion batteries. Using AlCl3 aqueous electrolytes, Mn3O4 cathodes demonstrated 271 mAh g−1 under a current density of 0.5 A g−1, yielding a capacity retention of 98% over 1000 cycles.106 Phase transition from a Mn3O4 spinel to a MnO2-layered structure is hypothesized to be one of the reasons for high cycling stability. The author's group also investigated the potential use of AlFeO3 cathodes, demonstrated high cell performance.107 There are many cathode candidates besides these materials such as carbon, graphite, and sulfur, which should be discussed in a forthcoming paper.

4. Conclusion and future prospects

As explained above, numerous studies have taken on the challenge to overcome the major obstacles of RAAIBs, which includes Al anode corrosion, dendrite formation, and passivation of the inert Al2O3 layer. Intercalation/de-intercalation of Al3+ ions from the cathode material is also a major problem for RAAIBs. So far, aluminum chloride- and fluoride-based aqueous electrolytes seem to be the most promising. Furthermore, addition of ILs, deep eutectic solvents into water-based electrolytes, surface treatment of aluminum anodes, alloying of aluminum anodes, hybrid salt application, and cathode optimization are all necessary directions for future research topics to make RAAIBs feasible for industrial rechargeable batteries.

Conflicts of interest

There are no conflicts to declare.

Data availability

No new data were created or analyzed in this study. This review is based on existing literatures and previously published datasets, as cited in the reference list.

Acknowledgements

The author wishes to express thanks to Dr. Hideki Yoshioka, Ms. Maiko Mori and Ms. Ayako Mori for their helpful discussions.

References

  1. S. Orangi, N. Manjong, D. P. Clos, L. Usai, O. S. Burheim and A. H. Strømman, J. Energy Storage, 2024, 76, 109800–109813 CrossRef .
  2. W. Q. Chen, J. Ind. Ecol., 2013, 17, 926–938 CrossRef .
  3. R. Mori, RSC Adv., 2017, 7, 6389–6395 RSC .
  4. R. Mori, Sustainable Energy Fuels, 2017, 1, 1082–1089 RSC .
  5. R. Mori, Phys. Chem. Chem. Phys., 2018, 20, 29983–29988 RSC .
  6. Q. Liu, H. Wang, C. Jiang and Y. Tang, Energy Storage Mater., 2019, 23, 566–586 CrossRef .
  7. M. H. Alfaruqi, S. Islam, J. Lee, J. Jo, V. Mathew and J. Kim, J. Mater. Chem. A, 2019, 7, 26966–26974 RSC .
  8. A. M. Diem, J. Bill and Z. Burghard, ACS Appl. Energy Mater., 2020, 3, 4033–4042 CrossRef CAS .
  9. S. Wang, et al., ACS Appl. Nano Mater., 2019, 2, 6428–6435 CrossRef CAS .
  10. J. Tu, H. Lei, Z. Yu and S. Jiao, Chem. Commun., 2018, 54, 1343–1346 RSC .
  11. H. Li, H. Yang, Z. Sun, Y. Shi, H.-M. Cheng and F. Li, Nano Energy, 2019, 56, 100–108 CrossRef CAS .
  12. S. Guo, H. Yang, M. Liu, X. Feng, H. Xu, Y. Bai and C. Wu, ACS Appl. Energy Mater., 2021, 4, 7064–7072 CrossRef CAS .
  13. S. Liu, G. L. Pan, G. R. Li and X. P. Gao, J. Mater. Chem. A, 2015, 3, 959–962 RSC .
  14. K. Q. Qin, J. H. Huang, K. Holguin and C. Luo, Energy Environ. Sci., 2020, 13, 3950–3992 RSC .
  15. S. Wang, S. Huang, M. Yao, Y. Zhang and Z. Niu, Angew. Chem., Int. Ed., 2020, 59, 11800–11807 CrossRef CAS .
  16. P. R. Gifford and J. B. Palmisano, J. Electrochem. Soc., 1987, 134, 610–614 CrossRef CAS .
  17. X. Z. Yu, B. Wang, D. C. Gong, Z. Xu and B. G. Lu, Adv. Mater., 2017, 29, 1604118 CrossRef .
  18. E. Zhang, J. Wang, B. Wang, X. Yu, H. Yang and B. Lu, Energy Storage Mater., 2019, 23, 72–78 CrossRef .
  19. J. Zhang, R. He, L. Jia, C. You, Y. Zhang, M. Liu, N. Tian, H. Lin and J. Wang, Adv. Funct. Mater., 2023, 33, 2305674 CrossRef CAS .
  20. S. Licht and D. Peramunage, J. Electrochem. Soc., 1993, 140, L4–L7 CrossRef CAS .
  21. J. Zhao, et al., Green Chem., 2025, 27, 352 RSC .
  22. C. Yan, et al., J. Am. Chem. Soc., 2022, 144, 11444–11455 CrossRef CAS PubMed .
  23. S. Peng, et al., ACS Sustainable Chem. Eng., 2021, 9, 3710–3717 CrossRef CAS .
  24. Q. Ran, S. P. Zeng, M. H. Zhu, W. B. Wan, H. Meng, H. Shi, Z. Wen, X. Y. Lang and Q. Jiang, Adv. Funct. Mater., 2022, 33, 2211271 CrossRef .
  25. E. Zemlyanushin, B. Schwarz and S. Dsoke, J. Power Sources, 2025, 633(30), 236458 CrossRef CAS .
  26. M. Angell, C. J. Pan, Y. Rong, C. Yuan, M. C. Lin, B. J. Hwang and H. Dai, Proc. Natl. Acad. Sci. U. S. A., 2017, 114(5), 834–839 CrossRef CAS .
  27. G. A. Elia, O. B. Ducros, D. Sotta, V. Delhorbe, A. Brun, K. Marquardt and R. Hahn, ACS Appl. Mater. Interfaces, 2017, 9(44), 38381–38389 CrossRef CAS PubMed .
  28. Y. Liu, et al., Nano Energy, 2024, 130, 110140 CrossRef CAS .
  29. P. Meng, Z. Yang, J. Zhang, M. Jiang, Y. Wang, X. Zhang, J. Luo and C. Fu, Energy Storage Mater., 2023, 63, 102953 CrossRef .
  30. J. Verma and D. Kumar, Nanoscale Adv., 2021, 3, 3384 RSC .
  31. Y. Guo, F. Sun, T. Zhu, C. Zhang, K. Feng, X. Wang, G. Feng, F. You and C. Liang, Electrochim. Acta, 2024, 507(10), 145196 CrossRef CAS .
  32. E. O. Filatova and A. S. Konashuk, J. Phys. Chem. C, 2015, 119, 20755–20761 CrossRef CAS .
  33. J. S. Wilkes, J. A. Levisky, R. A. Wilson and C. L. Hussey, Inorg. Chem., 1982, 21, 1263–1264 CrossRef CAS .
  34. J. Smajic, S. Wee, F. R. F. Simoes, M. N. Hedhili, N. Wehbe, E. Abou-Hamad and P. M. F. J. Costa, ACS Appl. Energy Mater., 2020, 3, 6805–6814 CrossRef CAS .
  35. D. Y. Wang, et al., Nat. Commun., 2017, 8, 14283 CrossRef CAS .
  36. M. Angell, C. J. Pan, Y. Rong, C. Yuan, M. C. Lin, B. J. Hwang and H. Dai, Proc. Natl. Acad. Sci. U. S. A., 2017, 114(5), 834–839 CrossRef CAS .
  37. Z. Zhang, A. Kitada, S. Gao, K. Fukami, N. Tsuji, Z. Yao and K. Murase, ACS Appl. Mater. Interfaces, 2020, 12, 43289–43298 CrossRef CAS .
  38. K. V. Kravchyk and M. V. Kovalenko, Commun. Chem., 2020, 3, 120 CrossRef .
  39. C. Xie, et al., Adv. Funct. Mater., 2024, 34(48), 2408296 CrossRef CAS .
  40. Z. Yuan, Q. Lin, Y. Li, W. Han and L. Wang, Adv. Mater., 2023, 35(17), 2211527 CrossRef CAS .
  41. H. Wang, et al., Energy Environ. Sci., 2022, 15(1), 311 RSC .
  42. E. Faegh, B. Ng, D. Hayman and W. E. Mustain, Nat. Energy, 2021, 6(1), 21 CrossRef CAS .
  43. C. Xu, T. Diemant, X. Liu and S. Passerini, Adv. Mater., 2024, 36(24), 2400263 CrossRef CAS PubMed .
  44. C. Wang, J. Li, H. Jiao, J. Tu and S. Jiao, RSC Adv., 2017, 7, 32288–32292 RSC .
  45. X. Shen, et al., Nat. Commun., 2021, 12, 820–831 CrossRef CAS PubMed .
  46. L. D. Reed, S. N. Ortiz, M. Xiong and E. J. Menke, Chem. Commun., 2015, 51, 14397–14400 RSC .
  47. A. Mohammad, T. Köhler, S. B. H. Stöcker and D. C. Meyer, ACS Appl. Energy Mater., 2023, 6(5), 2914–2923 CrossRef CAS .
  48. Z. Yu, Y. Xie, W. Wang, J. Hong and J. Ge, Front. Chem., 2023, 11, 1190102 CrossRef CAS PubMed .
  49. J. Smajicz and P. M. F. J. Costa, J. Electrochem. Soc., 2022, 169, 040549 CrossRef .
  50. W. Tang, L. Deng, L. Guo, S. Zhou, Q. Jiang and J. Luo, Green Energy Environ., 2024, 9, 1183–1191 CrossRef CAS .
  51. J. Vatamanu and O. Borodin, J. Phys. Chem. Lett., 2017, 8, 4362–4367 CrossRef CAS PubMed .
  52. Y. Li, L. Liu, Y. Lu, R. Shi, Y. Ma, Z. Yan, K. Zhang and J. Chen, Adv. Funct. Mater., 2021, 31, 2102063 CrossRef CAS .
  53. J. F. Stoddart, D. J. Kim, J. W. Choi and D. J. Yoo, Angew. Chem., Int. Ed., 2021, 60, 5794 CrossRef PubMed .
  54. J. Smajic, A. Alazmi, N. Wehbe and P. M. F. J. Costa, Nanomaterials, 2021, 11, 3235–3244 CrossRef CAS .
  55. L. N. Xu, J. Y. Zhu, M. X. Lu, L. Zhang and W. Chang, Int. J. Miner., Metall. Mater., 2015, 22, 500–508 CrossRef CAS .
  56. B. A. Boukamp, Solid State Ionics, 2001, 143, 47–55 CrossRef CAS .
  57. M. F. J. Costa, Adv. Mater. Interfaces, 2021, 2101733 Search PubMed .
  58. C. D. Wagner, A. V. Naumkin, A. Kraut-Vass, J. W. Allison, C. J. Powell and J. R. Rumble Jr., NIST standard reference database, National Institute of Standards and Technology, Gaithersburg, Maryland, USA 2003.
  59. J. Chen, Q. Zhu, L. Jiang, R. Liu, Y. Yang, M. Tang, J. Wang, H. Wang and L. Guo, Angew. Chem., 2021, 60, 5794 CrossRef CAS PubMed .
  60. B. Wang, A. J. Fielding and R. A. Dryfe, J. Phys. Chem. A, 2019, 123, 22556 CAS .
  61. M. J. Park, H. Y. Asl and A. Manthiram, ACS Energy Lett., 2020, 5, 2367 CrossRef CAS .
  62. Q. Sun, L. Chai, S. Chen, W. Zhang, H. Y. Yang and Z. Li, ACS Appl. Mater. Interfaces, 2024, 16(8), 10061–10069 CrossRef CAS PubMed .
  63. Q. Ran, et al., Nat. Commun., 2022, 13, 576 CrossRef CAS PubMed .
  64. C. Wu, et al., Nat. Commun., 2019, 10, 73 CrossRef CAS PubMed .
  65. S. Nandi, H. Lahan and S. K. Das, Bull. Mater. Sci., 2020, 43, 26 CrossRef CAS .
  66. B. Unal, O. Sel and R. D. Cakan, J. Appl. Electrochem., 2024, 54, 1425 CrossRef CAS .
  67. R. Tao, C. Gao, E. Xie, B. Wang and B. Lu, Chem. Sci., 2022, 13, 10066 RSC .
  68. C. Gao, et al., Adv. Funct. Mater., 2018, 28, 1803129 CrossRef .
  69. A. Holland, R. D. Mckerracher, A. Cruden and R. G. A. Wills, et al., J. Appl. Electrochem., 2018, 48, 243 CrossRef CAS .
  70. S. A. Rastabi, G. Razaz, M. Hummelgård, T. Carlberg, N. Blomquist, J. Ortegren and H. Olin, J. Power Sources, 2022, 523, 231066 CrossRef .
  71. J. R. Davis, ASM International Handbook, 1993, p. 547 Search PubMed .
  72. G. Razaz and T. Carlberg, Metall. Mater. Trans., 2019, A50(4), 1873 CrossRef .
  73. Y. J. Cho, I. J. Park, H. J. Lee and J. G. Kim, J. Power Sources, 2015, 277, 370 CrossRef CAS .
  74. Y. Xu, et al., Energy Storage Mater., 2022, 47, 113 CrossRef .
  75. P. Meng, J. Huang, Z. Yang, F. Wang, T. Lv, J. Zhang, C. Fu and W. Xiao, Adv. Mater., 2022, 34(8), 2106511 CrossRef CAS PubMed .
  76. R. Mori, Japan Patent Application Number, 2025-055734.
  77. R. Mori, Japan Patent Application Number, 2025-055735.
  78. Y. Ru, S. Zheng, H. Xue and H. Pang, Chem. Eng. J., 2020, 382, 122853 CrossRef CAS .
  79. P. Wang, et al., Chem. Eng. J., 2019, 373, 580 CrossRef CAS .
  80. H. Lahan and S. K. Das, J. Power Sources, 2019, 413, 134 CrossRef CAS .
  81. Y. H. Wu, W. J. Lin, T. H. Tsai and M. C. Lin, ACS Appl. Energy Mater., 2024, 7, 3957 CrossRef CAS .
  82. A. Ejigu, L. W. Le Fevre, A. Elgendy, B. F. Spencer, C. Bawn and R. A. W. Dryfe, ACS Appl. Mater. Interfaces, 2022, 14, 25232 CrossRef CAS PubMed .
  83. Q. Zhao, M. J. Zachman, W. I. A. Sadat, J. Zheng, L. F. Kourkoutis and L. Archer, Sci. Adv., 2018, 4, eaau8131 CrossRef CAS PubMed .
  84. Y. Chunshuang, et al., J. Am. Chem. Soc., 2022, 144(25), 11444 CrossRef PubMed .
  85. S. Yang, C. Li, H. Lv, X. Guo, Y. Wang, C. Han, C. Zhi and H. Li, Small Methods, 2021, 5(10), 2100611 CrossRef CAS PubMed .
  86. H. Tian, S. Zhang, Z. Meng, W. He and W. Q. Han, ACS Energy Lett., 2017, 2(5), 1170 CrossRef CAS .
  87. M. Chiku, H. akeda, Y. Yamaguchi, E. Higuchi and H. Inoue, Int. J. Chem., 2013, 5, 4 Search PubMed .
  88. X. Wen, et al., J. Phys. Chem. Lett., 2021, 12(25), 5903 CrossRef CAS PubMed .
  89. A. Lahiri, S. Guan and A. Chutia, ACS Appl. Energy Mater., 2023, 6, 11874 CrossRef CAS PubMed .
  90. V. Balland, M. Mateos, A. Singh, K. D. Harris, C. L. Robust and B. Limoges, Small, 2021, 17, 2101515 CrossRef CAS PubMed .
  91. P. A. Parayad, et al., Eur. J. Inorg. Chem., 2018, 14146 Search PubMed .
  92. G. R. Pastel, et al., Energy Environ. Sci., 2022, 15, 2460 RSC .
  93. R. Bai, J. Yang, G. Li, J. Luo and W. Tang, Energy Storage Mater., 2021, 41, 41 CrossRef .
  94. B. E. Jia, et al., Energy Storage Mater., 2024, 65, 103141 CrossRef .
  95. F. Wang, O. Borodin, T. Gao, X. Fan, W. Sun and F. Han, et al., Nat. Mater., 2018, 17, 543 CrossRef CAS PubMed .
  96. L. Suo, et al., Angew. Chem., Int. Ed., 2016, 55, 7136 CrossRef CAS PubMed .
  97. T. Yang, et al., Appl. Surf. Sci., 2024, 660, 159995 CrossRef CAS .
  98. X. Li, et al., J. Mater. Chem. A, 2022, 10, 4739 RSC .
  99. Y. Gao, Y. Li, H. Yang, L. Zheng, Y. Bai and C. Wu, J. Energy Chem., 2022, 67, 613 CrossRef CAS .
  100. X. Yuan, et al., Adv. Energy Mater., 2025, 15, 16 Search PubMed .
  101. Z. Wen, et al., Adv. Mater., 2025, 37(15), 2500695 CrossRef CAS PubMed .
  102. H. Yang, H. Li, J. Li, Z. Sun, K. He, H. M. Cheng and F. Li, Angew. Chem., Int. Ed., 2019, 58, 11978 CrossRef CAS PubMed .
  103. S. Chen, Y. Kong, C. Tang, N. A. Gadelhak, A. K. Nanjundan, A. Du, C. Yu and X. Huang, Small, 2024, 20, 2312229 CrossRef CAS PubMed .
  104. J. L. Yang, W. B. Gong and F. X. Geng, Adv. Funct. Mater., 2023, 33, 2301202 CrossRef CAS .
  105. L. Chen, C. Lv, H. Gu, W. Zhang and Z. Li, J. Colloid Interface Sci., 2025, 677, 1045 CrossRef CAS PubMed .
  106. P. De, L. Bharti, J. Halder, S. Priya and A. Chandra, Electrochim. Acta, 2023, 469(20), 143248 CrossRef CAS .
  107. R. Mori, Japan Patent Application Number, 2025- 008535.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.