Open Access Article
Hui Shi
ab,
Anastasia A. Teck
b,
Laura Landerc and
Maria-Magdalena Titirici
*bd
aSchool of Mechanical Engineering, Chengdu University, Chengdu 610106, China
bDepartment of Chemical Engineering, Imperial College London, London SW7 2AZ, UK. E-mail: m.titirici@imperial.ac.uk
cDepartment of Engineering, King's College London, London WC2R 2LS, UK
dAdvanced Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan
First published on 11th November 2025
Aluminum–graphite dual-ion batteries (AGDIBs) are emerging as a promising alternative in electrochemical energy storage due to aluminum's abundance, low cost, intrinsic safety, high power density, and excellent performance across a wide range of temperatures. Unlike single-ion rocking-chair batteries, AGDIBs operate via a dual-ion mechanism: Al plating/stripping at the Al anode and AlCl4− anions intercalation/deintercalation in the graphitic carbon cathode. Recent years have witnessed significant progress in the development of AGDIBs, particularly in anode interface engineering, graphite cathode optimization, and the formulation of novel electrolytes. Despite these advancements, critical challenges persist, including the low specific capacity of the graphitic carbon cathode and severe electrolyte-induced corrosion. This review offers a comprehensive and up-to-date analysis of recent advancements and persistent challenges in the development of AGDIBs. It systematically examines innovations across all components—including anodes, cathodes, electrolytes, and others—highlighting breakthroughs in materials design and performance optimization. Beyond summarizing progress, the review critically identifies unresolved issues and knowledge gaps, offering forward-looking insights to guide future research efforts.
Green foundation1. We review the recent breakthroughs in aluminum–graphite dual-ion batteries (AGDIBs), emphasizing the use of Earth-abundant materials (aluminum, graphite), non-flammable ionic liquid electrolytes, and extended temperature stability. These developments reduce reliance on critical metals and improve safety and robustness in battery systems.2. AGDIBs align with circular economy and sustainable energy goals by offering lower environmental impact, enhanced safety, and reduced resource risk compared to lithium-ion systems. As demand for green energy storage grows, AGDIBs present an attractive alternative, addressing both technical and policy imperatives. 3. We identify key challenges such as limited cathode capacity, electrolyte corrosion, and lifecycle integration. We propose targeted directions—e.g. novel cathode architectures, corrosion-resistant interfaces, and system-level design strategies—that can guide future R&D toward truly sustainable, high-performance AGDIB systems. |
![]() | ||
| Fig. 1 (a) Ragone plot of AGDIBs reported on the literature4,5,18–20 in comparison with lithium-ion batteries (LIBs), electrochemical capacitors, capacitors, sodium-ion batteries (NIBs), supercapacitors, and redox flow batteries. (b) Schematic charge storage mechanisms of dual-ion and rocking-chair batteries during the charge/discharge process. The different colors used to represent Al in Al2Cl7− and AlCl4− are purely for visual distinction and do not reflect any actual difference in chemical identity or oxidation state. | ||
Understanding how AGDIBs function requires examining their distinct charge storage mechanism, which differs fundamentally from traditional AIBs. Unlike rocking-chair AIBs, which struggle with reversible Al3+ ion storage, AGDIBs operate via a different mechanism. In rocking-chair systems,9–11 the energy storage relies on the reversible transfer of Al3+ ions and metallic aluminum between both electrodes (Fig. 1b). However, it often suffers from poor reversibility due to the trivalent ion's high charge density and sluggish kinetics. In contrast, AGDIBs belong to a class of dual-ion systems where energy storage is enabled by the shuttling of complex aluminum chloride anions between the electrolyte and the electrodes during charge and discharge. This anion-shuttling mechanism avoids the challenges associated with Al3+ intercalation. In AGDIBs, the electrolyte assumes a dual function: it not only mediates ion transport between the electrodes but also provides the electroactive complexes (e.g., AlCl4− and Al2Cl7−) that directly participate in charge storage at both the anode and cathode. Currently, AlCl3/1-ethyl-3-methyl-imidazolium chloride ([EMIm]Cl) is the main electrolyte applied in AGDIBs. The cell operates within 0 to 2.5 V, limited by the electrochemical stability of the chloroaluminate electrolyte and the redox potentials of the electrodes. The onset potential for Al deposition was identified at −0.25 to −0.12 V vs. Al/Al3+, while the stripping current drops to zero between 0.8 and 1.5 V due to passivation of the aluminum electrode. Both deposition and stripping behaviors are strongly influenced by the specific electrolyte formulation.12–14 Concurrently, at the cathode, chloroaluminate anions (AlCl4−) intercalate into graphitic carbon at an average voltage of 2.01–2.3 V,15 leading to the formation of graphite intercalation compounds (GICs). The principle can be simply described as following equations16 and Fig. 1b:
Anode:
| 4Al2Cl7− + 3e− ↔ Al + 7AlCl4− |
Cathode:
| Cn + AlCl4− ↔ Cn[AlCl4] + e− |
Research on AGDIBs has mainly emerged within the past decade, with a limited number of studies making notable contributions; nevertheless, the field remains in its infancy and requires further exploration. AGDIBs have undergone notable advancements in key components such as anode, electrolyte, and cathode to improve their overall performance. Fig. 2 summarizes the key characteristics of the battery components and the associated challenges. The anode, typically made of Al, offers high abundance, excellent power density, and comparable energy density. However, it faces challenges such as corrosion, dendrite growth, and the formation of a passivation layer. In addition, the actual capacity of the battery is primarily limited by either the cathode materials or the electrolyte composition. Most cathodes in AGDIBs are made of graphite and few-layer graphene, which offer good rate performance but struggle with limited intercalation capacity. Some studies have explored hard carbon, which demonstrates high capacity but at low current densities. At this stage, there is a lack of understanding of the storage mechanism and the impact of structure and morphology on electrochemical performance hindering the further optimization of AGDIBs. Chloroaluminate-based ionic liquids, such as AlCl3/[EMIm]Cl, are widely employed due to their negligible volatility, low flammability, and, most importantly, their unique electrochemical behavior. These electrolytes enable both reversible Al plating and anion intercalation, thanks to the presence of two distinct chloroaluminate species. This dual functionality has not been achieved with many other solvents, particularly those that are chloride-free. The anolyte exhibits a specific capacity ranging from 21 to 46 mAh g−1, depending strongly on the AlCl3
:
[EMIm]Cl molar ratio (ranging from 1.3 to 2.0) and the applied current density.21 However, this electrolyte is highly corrosive to other cell components, posing challenges in selecting suitable materials for the battery. Overall, research has mainly targeted optimization of graphitic cathodes to enhance battery performance, alongside efforts to mitigate Al corrosion through anode modification or the use of less corrosive electrolytes. However, limited understanding of the Al–electrolyte interface, conflicting views on cathode intercalation mechanisms, and severe corrosion in imidazolium-based electrolytes highlight critical gaps and opportunities for further investigation. In this review, recent advances in the development of AGDIBs are summarized, with an emphasis on key findings from recent literature. From the perspective of practical application, major challenges and potential solutions related to various cell components are critically discussed. Finally, a summary and outlook are provided to guide future research on high-performance component materials and to support the advancement of practical, next-generation AGDIBs.
Currently, common solutions to corrosion are limited to surface modifications that create an artificial solid-electrolyte interphase (SEI). This surface layer affects not only corrosion but also Al plating/stripping. Coating an SEI layer on Al anode as an inert physical barrier allows the electrolyte to homogeneously infiltrate the surface of the Al anode, which makes the corrosion on the Al anode surface more uniform. Consequently, it generates a uniform Al plating/stripping rather than dendrite growth. Graphite was constructed on Al foil by simple blade coating method as an SEI layer.30 After 40 h cycling at 0.1 mA cm−2 current density, the pristine Al shows big, corroded areas (15 μm), while no obvious corrosion can be observed on the surface of the Al anode underneath the SEI layer. However, the electrochemical performance was not outstanding, and no analysis on the possible participation of the graphite layer in the storage mechanism was provided. Corrosion inhibitors were used in electrolytes to create dense and uniform barrier SEI layers on electrodes, which not only mitigate the corrosion but also dendrite growth. An ab initio molecular dynamics (AIMD) simulation suggests that proper additives can preferentially form stable SEI layers.31 Nonionic surfactant F12732 and cetyltrimethyl-ammonium chloride (CTAC)33 were used as corrosion inhibitors in AlCl3/Et3NHCl (1.5
:
1) electrolytes, respectively. The consequent barrier layers mitigate the corrosion and effectively induce uniform Al plating/stripping on anode (Fig. 4a). The intercalation mechanism into graphite was not affected, while the cells with additives illustrate higher specific capacities at a current density of 500 mA g−1 (Fig. 4b and c). A pretreatment (alkali-assisted boiling water treatment) followed by an anodizing process to the Al foil was reported.34 This pretreatment generates an artificial SEI (a hydrated oxide film) which then transforms into a denser anodic oxide film (391 nm thickness of Al2O3) and a flake-like outer layer (Al2O3·H2O). This process has been shown to enhance the corrosion resistance of the Al anode. In brief, the creation of an artificial SEI or the addition of corrosion inhibitors can effectively mitigate Al corrosion.
![]() | ||
Fig. 4 (a) The SEM images of Al foil soaked in the AlCl3/Et3NHCl (1.5 : 1) electrolyte and the ones with different additives: F127, dodecyltrimethylammonium chloride (DTAC), cetyltrimethyl-ammonium chloride (CTAC), stearyltrimethylammonium chloride (STAC), tetrabutylammonium chloride (TBAC), and benzyldimethylhexadecylammonium chloride (HDBAC). (b) CV curves and (c) charge/discharge profiles of Al//FG full batteries using electrolyte with/without F127 and CTAC additives. Adapted with permission from Elsevier.32 Copyright 2024. Adapted with permission from John Wiley and Sons.33 Copyright 2024. | ||
However, the role of the native Al2O3 film on Al anode remains a subject of considerable debate. Metallic Al is naturally covered with an Al2O3 layer owing to the high reactivity of pristine Al. While this surface oxide is intrinsically both ionically and electronically insulating, its presence has been reported to exert either protective or detrimental effects, depending on its thickness and uniformity. The native oxide layer on the Al surface can be attacked and dissolved by chloroaluminate anions (AlCl4− and Al2Cl7−) of the corrosive AlCl3/[EMIm]Cl electrolyte, exposing a fresh Al surface directly to the electrolyte for Al dissolution and deposition reactions. This process prevents complete passivation and enables reversible Al plating/stripping; however, it simultaneously promotes progressive surface degradation. A dipping experiment of Al metal in AlCl3/[EMIm]Cl electrolyte indicates severe pitting corrosion, which becomes more pronounced in more acidic compositions—i.e., at higher AlCl3 molar ratios—where the concentration of Al2Cl7− species is elevated.29 The pitting corrosion and deposition of the newly formed Al oxide on the defective sites results in growing cracks and fragments on the Al metal surface.
Conflicting findings have been reported regarding the influence of oxide thickness. A thicker Al2O3 layer can better shield the Al anode from aggressive electrolytes, mitigating corrosion effects.35,36 Yet, excessive thickening imposes kinetic limitations by hindering electron transfer and increasing the nucleation barrier for Al deposition, thereby elevating polarization, especially at high current densities.37 In contrast, thinner oxide films (<50 nm) have been shown to suppress dendritic growth, and their removal was correlated with markedly poorer cycling stability in Al//graphite cells.38 Electrochemical pretreatments such as electropolishing, which leave behind an ultrathin oxide layer, result in a larger electrochemically active surface area. Although such surfaces exhibit more evident localized corrosion, the corresponding full cells display enhanced capacity and cycling stability.14,39
Recent studies further suggest that the passivation layer is chemically more complex than a simple Al2O3 film. Instead, cycling generates a stratified passive film consisting of an outer inorganic/organic layer (including carbonates, chlorides, and nitrides) and an inner oxide-rich region.40 This structural and compositional heterogeneity highlights the multifaceted role of surface passivation. While it may partially protect the Al anode against corrosive electrolytes, it can also aggravate surface degradation under certain conditions, thereby complicating interfacial processes and electrochemical performance. Current understanding of the formation, dissolution, and structural evolution of this oxide layer remains limited, and systematic studies—combining advanced structural characterization with theoretical modeling—are required to elucidate its governing mechanisms and to guide rational interface engineering strategies.
:
1) electrolyte.41 In this study, two primary rate-limiting steps have been proposed during Al plating. The first is a chemical step involving the sequential release of chloride ions (Cl−) from aluminum chloride complexes, represented as AlClm− → AlCl(m−1) + Cl−, m = 3, 2, 1. The second involves an electrochemical step where Al undergoes a stepwise reduction from its trivalent (Al3+) to monovalent (Al+) oxidation state by accepting electrons, exhibited as Alm+ + e− → Al(m−1)+, m = 3, 2, 1. These processes are governed by the kinetics of Al speciation and complexation in the electrolyte. Conversely, during stripping, the rate-determining mechanism involves a sequence of three single-electron transfer steps, coupled with chemical reactions that release chloride ions back into the electrolyte. The efficiency and reversibility of Al plating/stripping are highly dependent on the dynamics of Al-based complex ions, particularly AlCl4−, Al2Cl7−, and other chloroaluminate species. The speciation and interconversion of these anions influence not only the deposition rate but also the morphology and uniformity of the Al layer. Furthermore, side reactions involving intermediate species at the Al surface, as well as the electrolyte concentration and composition, play critical roles in the evolution of the anode surface. These factors must be carefully considered when analyzing the electrochemical behavior and long-term stability of Al anodes in AGDIB systems.
Al is often considered a dendrite-free metal; however, various electrodeposition morphologies resembling dendrites have been observed and are commonly referred to as “dendrites” in the literature. Craig et al.42 reported that Al's face-centered cubic crystal structure, characterized by a high coordination number of 12, promotes dense and uniform metal deposition due to its energetically favorable atomic reorganization. This behavior contrasts with metals possessing lower coordination numbers. Dendrites are expected to only form at high current densities above 100 mA cm−2,13 which is significantly above the typical operational range (up to 10 mA cm−2)1 in AGDIBs. Thus, a key advantage of Al is its dendrite-free behavior at moderate current densities when plated onto suitable substrates such as aluminum, tungsten, or planar vitreous carbon. Elia et al.43 observed extensive pitting and irregular deposit morphology after 500 cycles but reported no evidence of protruding dendritic structures. Nevertheless, several studies have documented dendrite formation in AGDIBs, indicating that this issue remains under investigation. In general, the electrodeposition process initiates with nucleation, followed by the growth of the deposited layer. Once a crystal structure is established, further growth becomes limited by either diffusion or interfacial kinetics, depending on which process is slower. Dendrite formation is typically associated with inhomogeneous electrodeposition during repeated Al plating and stripping cycles. In AGDIBs, dendritic growth on Al foil is often attributed to uneven surface corrosion and the presence of impurities at the anode–electrolyte interface (Fig. 5). These dendritic structures are typically less electrochemically active than pristine Al, resulting in uneven current distribution and localized charge accumulation. Consequently, Al tends to deposit preferentially at regions of higher electric field intensity and reduced ion diffusion distance, promoting dendrite growth—particularly under high current densities.1,13,42,43
![]() | ||
| Fig. 5 Schematic illustration of dendrite formation on the Al anode, despite its intended role as a dendrite-free material. | ||
In fact, the formation of dendrites is influenced by several factors, including crystal orientation, surface roughness, current density, and so on. Compared to the dense, continuous, and well-adherent Al electrodeposition shown in Fig. 6a, various other morphologies have been reported. Flake-shaped deposits, which cover the Al substrate homogeneously, were observed in an AlCl3
:
[BMIm]Cl (2
:
1) electrolyte (Fig. 6b).44 As shown in Fig. 6b, these deposits are planar, thin, and large, vertically oriented to form walls with an average width of 20 μm and a thickness of less than 3 μm. Wang et al.45 reported that Al tends to form (111) crystal planes during electrodeposition in AlCl3/Et3NHCl (1.5
:
1) electrolyte. The charge transfer resistance of (111) Al electrodes is significantly lower than (200) Al, (220) Al, and (311) anodes, indicating its faster Al plating/stripping reaction kinetics. As a result, choosing Al with preferred (111) crystal plane as anode could obtain high lattice matching between the substrate and deposited metal, allowing uniform deposition of the Al (111) plane on the Al surface, as shown in Fig. 6c. On the other side, it found that the Al deposition displays a preferential crystallographic orientation along the (200) plane in both AlCl3/Et3NHCl
12 and AlCl3/[EMIm]Cl
46 electrolytes; however, no in-depth explanation to why this is the case was presented. Besides, deposition potential was also found to affect the deposits morphology (Fig. 6d).12 At a deposition potential of −0.3 V, the Al deposition layer appears relatively uniform and dense. However, when the potential is further decreased to −0.5 V, the layer shows the formation of nanowire structures with an approximate diameter of 367 nm. Additionally, the thickness of the deposition layer increases significantly, from 40.47 μm to 127.3 μm, accompanied by a rise in surface roughness. She et al.37 found that dendrite growth involves a series of complex processes, including nucleation, growth, and stripping. The density of dendrites increased with increasing current densities between 1 and 10 mA cm−2, and the corresponding morphological evolution varied from needle-like to spherical shape (Fig. 6e). The dendrites formed during the initial cycles dissolve over time due to contact with a large excess of acidic electrolyte.46 Overall, dendrite formation has been reported in the literature, and it can be mitigated by controlling factors such as the crystallographic orientation of the anode, surface morphology, current density, and other experimental conditions.
![]() | ||
Fig. 6 SEM images of Al surface with electrodeposited Al: (a) faceted structure after 20 min at 60 °C and −0.30 V. Reproduced with permission from Elsevier.13 Copyright 2006. (b) Homogeneous coverage with flake-shaped deposits. Reproduced with permission from Elsevier.44 Copyright 2018. (c) Varying morphologies due to the primary crystal orientations after 50 cycles. Scale bar: 10 μm. Reproduced with permission from Springer Nature.45 Copyright 2024. (d) SEM image of the electrodeposited Al on Cu substrate at 313 K for 1 h under potential of −0.2 V to −0.5 V, using a three-electrode system with AlCl3/Et3NHCl (1.7 : 1) electrolyte. Reproduced with permission from Springer Nature.12 Copyright 2024. (e) SEM images of Al surface with electrodeposited Al with increasing current density between 1 and 10 mA cm−2 using a symmetric battery with AlCl3/[EMIm]Cl (1.3 : 1) electrolyte. Reproduced with permission from IOP Publishing Limited.37 Copyright 2020. | ||
Various strategies have been proposed to suppress dendrite formation on aluminum anodes, with most efforts centering on surface modification techniques, particularly the introduction of an Al2O3 layer to regulate deposition behavior. The strong oxygen-mediated bonding between Al deposits and the substrate has been reported to facilitate uniform planar growth along the interface.47 For instance, immersion in boiling water followed by phosphoric acid treatment yields a dense ∼100 nm aluminum phosphate layer that allows Al deposits to form without significant dendrite growth.48 Interestingly, some studies reported that a uniform Al2O3 layer on the surface could efficiently restrict the growth of crystalline Al dendrites and thus improve the cycling stability of the Al anode.38 Other approaches combine surface modification with substrate design, such as fabricating a micro-porous Al substrate with a partial Al2O3 coverage and leaving oxide-free regions to direct localized deposition.1,49 However, the long-term viability of these strategies remains questionable. In highly acidic electrolytes, the Al2O3 film readily dissolves and cannot function as a persistent barrier, while its pitting corrosion often introduces surface inhomogeneities that accelerate dendrite growth. Consequently, the role of Al2O3 layers in dendrite suppression remains controversial, warranting further systematic investigation.
Table 1 summarizes the reported modifications of Al anodes in dual-ion batteries with graphite cathodes. Early studies mainly investigated Al volume changes in LiPF6/EMC-based Al//graphite systems, which are essentially lithium-ion batteries.50–53 In contrast, research on AGDIBs—typically employing AlCl3/Et3NHCl or AlCl3/[EMIm]Cl electrolytes—has focused primarily on surface treatments, protective coatings, and composite designs to suppress corrosion and promote uniform electrodeposition. However, only a few studies have extended these investigations to demonstrate meaningful improvements in full-cell cycling performance. More recent reports suggest that surface engineering strategies such as electropolishing, controlled oxide layer formation, and carbon-based coatings have enabled stable operation over extended lifetimes, with reports of up to 25
000 cycles. Despite these advances, most work has emphasized interfacial stabilization rather than practical battery metrics, and fundamental issues such as the volume change of Al electrodes in AGDIBs remain unexplored, representing a key knowledge gap in the field.
| Anode|cathode | Electrolyte | Specific capacity (mAh g−1)/current density (mA g−1) | Long-term cycling | Year Ref. |
|||
|---|---|---|---|---|---|---|---|
| Specific capacity (mAh g−1) | Current density (mA g−1) | Capacity retention | Cycle number | ||||
| a (111) Al: Al anode with (111) preferred crystal plane. BPA@Al: Al anode prepared by a simple boiling water treatment followed by phosphoric acid immersion. 2P-Al2O3/Al: A microporous Al coated with aluminum oxide, with the bottom intentionally left uncoated. P-Al2O3/Al: A microporous Al with the upper surface coated with an aluminum oxide layer. EMC/VC: ethyl methyl carbonate with an addition of 5 wt% vinylene carbonate. nAl@C: core/shell aluminum@carbon nanospheres. CHA-Al: carbon-coated hollow aluminum nanospheres on Al foil. 3D pAl/C: 3D porous Al coated by carbon layer. | |||||||
(111) Al a|three-dimensional graphite |
AlCl3/Et3NHCl 1.5 : 1 |
102/1000 | 80 | 5000 | 80.0% | 25 000 |
2024 45 |
BPA@Al |three-dimensional graphite |
AlCl3/Et3NHCl 1.5 : 1 |
112/500 | 96.7 | 5000 | 97.8% | 20 000 |
2024 48 |
2P-Al2O3 /Al|graphite |
AlCl3/Et3NHCl 1.5 : 1 |
80/20 | 78.8 | 100 | 95% | 1000 | 2023 49 |
P-Al2O3 /Al|flake graphite |
AlCl3/Et3NHCl 1.5 : 1 |
79.2/1C rate | 82.4 | 2C rate | 94.3% | 500 | 2023 1 |
| Graphite coated Al|flake graphite | AlCl3/[EMIm]Cl 1.3 : 1 |
— | 98 | 100 | 97% | 740 | 2023 30 |
| Electropolished Al|graphite | AlCl3/[EMIm]Cl 1.3 : 1 |
90/1000 | 53 | 5000 | 72% | 10 000 |
2022 14 |
| 3D Al foam|graphite | 1 M LiPF6 in EMC/VC* | — | 105 | 100 | 96% | 600 | 2020 50 |
nAl@C |graphite |
4 M LiPF6 in EMC/VC | 91/200 | 88 | 1500 | 94.6% | 1000 | 2018 51 |
CHA-Al |expanded graphite |
4 M LiPF6 in EMC | 98/5C rate | 100 | 2C rate | 99% | 1500 | 2017 52 |
3D pAl/C |graphite |
4 M LiPF6 in EMC/VC | 85/20C rate | 93 | 2C rate | 89.4% | 1000 | 2016 53 |
Beyond interfacial stability and electrochemical performance, the sustainability dimension of AGDIBs—particularly the environmental and economic implications of aluminum extraction and recycling—remains insufficiently addressed. Despite aluminum's abundance, its extraction is highly unsustainable due to energy-intensive smelting and associated environmental damage. In 2021, global primary Al smelting capacity reached 81 million tons per year, generating total emissions of 651 million metric tons of CO2 equivalent, with 82% of these emissions stemming from electricity use and 18% from the smelting process itself.54 The process also releases other pollutants, including perfluorocarbons, and causes land degradation and water pollution, particularly from bauxite mining. Efforts to reduce these impacts include increasing Al recycling, which cuts CO2 emissions by 95%, and adopting cleaner energy in production.55 Though both Al mining and recycling have environmental impacts, recycling is undoubtedly more sustainable than primary production. Ideally, efforts should focus on reducing the demand for primary Al and improving the efficiency of recycling systems. While Al-based batteries may contribute to increased overall Al demand, they offer notable advantages over lithium-ion batteries, such as greater supply security, material abundance, and recyclability. At present, however, high-purity Al foils are required to prevent side reactions caused by trace impurities, significantly increasing production costs. As of 2024, the prices of high-purity aluminum were exceptionally high: 4N (99.99%) at €261.95 per kg, 5N (99.999%) at €475.95 per kg, and 6N (99.9999%) at €950.95 per kg.56 Maximizing the use of recycled Al in battery manufacturing will therefore be essential to balance performance with sustainability.
In conclusion, the Al anode in AGDIBs presents several key advantages, including low cost, high volumetric capacity, excellent recyclability, and the potential for dendrite-free plating/stripping under moderate current densities. Nonetheless, key challenges remain:
• Corrosion in aggressive electrolytes undermines interfacial stability and long-term performance.
• Dendrite formation, though less severe than in Li systems, arises from impurities, current-density inhomogeneities, and complex multi-step plating/stripping kinetics.
• Al2O3 surface films show inconsistent effects, leaving interfacial reactions poorly understood.
Future progress will depend on elucidating interfacial mechanisms, optimizing electrolytes, and developing deposition-regulation strategies to ensure durable anode operation.
000 cycles. Compared to pyrolytic graphite (PG) and synthetic graphite (SG), NG showed a larger cyclic voltammetry (CV) area in the 1.80–2.25 V range, indicating easier dilute-stage intercalation.58 CV measurements with NG cathodes also revealed five sharp, well-defined reversible peaks (Fig. 9a),59 suggesting a uniform multistage intercalation process. A free-standing NG film was attached to Ni bar current collectors using conductive carbon tape, which delivered a significantly higher specific capacity than PG free-standing cathode due to better structural reversibility.60 In situ Raman spectroscopy and ex situ XRD confirmed NG's highly reversible structural changes during charge/discharge. In contrast, the (002) XRD peak of PG did not fully return to its original state after discharge, indicating residual strain and structural disorder in the PG material. Overall, the flaky morphology and well-ordered atomic structure of NG facilitate efficient AlCl4− intercalation and highly reversible structural changes.
In addition to intercalation, adsorption within pores may also contribute to a higher capacity. Kish graphite (KG), a synthetic flake-shaped graphite, is available on a large scale as a byproduct of steel manufacturing. Besides its flake structure, KG possesses macro-structural defects such as holes and cracks. As cathode, the intercalation of AlCl4− ions occurs with a significant interlayer space expansion by up to 8.85 Å or 12.1 Å at intercalation stage 4 or 3 GICs, respectively.19 Large-sized kish graphite flakes with high structural perfection (Raman spectra ID/IG = 0.05) showed exceptional rate capability, yielding a high power density of 4363 W kg−1. High-quality Kish expanded graphite (KEG) cathode materials were synthesized from KG through a simple and rapid expansion process.17 As shown in Fig. 7a, the CV curves of KEG demonstrate excellent reversibility and stability in anion storage, which involves both diffusion-controlled intercalation and a significant contribution from adsorption-controlled processes. At higher current densities, AlCl4− anion storage becomes increasingly dominated by surface adsorption. Compared to unexpanded KG (Fig. 7a),61 the increased proportion of capacitive, adsorption-controlled behavior in KEG can be attributed to its fluffy micro-nano interlayer pore structure, which is formed by the cross-linking of single or few graphene layers. This porous architecture enhances anion accessibility and adsorption, leading to a notable improvement in specific capacity—from approximately 110 mAh g−1 to 140 mAh g−1 at 50 mA g−1 (Fig. 7b). Thus, the introduction of hierarchical pores effectively promotes additional adsorption sites, thereby boosting the overall storage capacity.
![]() | ||
Fig. 7 (a) CV curves of cells with natural graphite (NG),59 Kish graphite (KG)61 and kish-based expanded graphite (KEG)17 cathode and AlCl3/[EMIM]Cl (1.3 : 1) electrolyte, at a scan rate of 0.5 mV s−1. Insert is the contribution ratio of these three graphite cathodes. (b) Charging and discharging profiles of cells with various graphitic materials and AlCl3/[EMIM]Cl (1.3 : 1) electrolyte at 50 mA g−1. These graphite materials are KG61 and KEG,17 and kish graphite flakes (KGF, 2 mm), natural large graphite flakes (NGF, 1 mm) and acetylene black (AB, 26 μm).19 (c) CV curves at a scan rate of 10 mV s−1 and (d) charging/discharging profiles at a current density of 2 A g−1 of cells with NG59 (only used in CV), CB (XC-72) and CB-G (carbon black-derived nanosheet-bricked porous graphite, NSPG-120)66 cathode and AlCl3/[EMIM]Cl (1.3 : 1) electrolyte. Adapted with permission from Elsivier.17 Copyright 2024. Adapted with permission from author.19 Copyright 2017. Adapted with permission from Springer Nature.59 Copyright 2024. Adapted with permission from Elsivier.61 Copyright 2023. Adapted with permission from author.66 Copyright 2018. | ||
By constructing a highly porous three-dimensional graphitic foam that allows fast ion diffusion/intercalation can facilitate more rapid battery operation. A pioneering work using a 3D graphitic foam cathode, deposited on a nickel foam template by chemical vapor deposition, was published by Lin et al.18 in 2015. The Al//graphitic-foam pouch cell demonstrated a charging capability at a current density up to 5 A g−1, about 75 times higher than the Al//pyrolytic graphite cell, while maintaining a similar voltage profile and discharge capacity (∼60 mAh g−1). The cell sustained over 7500 cycles without capacity loss and achieved energy and power densities of 40 Wh kg−1 and 3000 W kg−1, respectively. They further prepared a 3D graphitic foam containing vertically aligned few-layered graphene sheets as cathode materials.62 This shows that a low density of defects and oxygen groups can increase accessible spaces for AlCl4− intercalation, leading to a higher capacity and a lower cutoff charge voltage.
It was found that only highly graphitic materials can intercalate AlCl4− anions, whilst the amorphous part of the carbon hinders the accessibility of AlCl4−.63 In Fig. 7b, acetylene black (AB) shows only a slope without a plateau, whereas natural graphite flakes (NGF) and kish graphite flakes (KGF) display typical graphite intercalation storage curves, characterized by both slope and plateau. Thus, to achieve high-quality graphite materials, not only abundant layers, porous morphologies (to facilitate diffusion) and adjustable interlayer spacing (for reversible anion intercalation and deintercalation), but also a high degree of graphitization (for enhanced conductivity) is required. Through a catalyst-free electrochemical graphitization, a flake graphite feature with high graphitization degree and crystallinity is easy to obtain from certain amorphous carbons.64,65 The graphitization was carried out in molten CaCl2 showing advantages of high-efficiency, low-temperature, environmentally friendliness and catalyst-free features. By this method, a nanosheet-bricked porous graphite (CB-G) was derived from carbon black (CB), presenting high crystallinity with an ID/IG of 0.11 and 37–38 layers of graphitic lattice planes.66 Even at a high current density of 100 A g−1, the specific capacity of CB-G maintains 90 mAh g−1. In Fig. 7c, the CV curve of CB shows no distinct peaks, whereas CB-G, although not as pronounced as NG, exhibits oxidation and reduction peaks. The cell charge/discharge profile of CB and CB-G (Fig. 7d) illustrates intercalation capacity only in CB-G, which is reflected in the plateau. This study suggests that AlCl4− anions are unable to intercalate into amorphous carbon but become detectable as the degree of graphitization increases. Simply increasing the carbonization temperature (e.g. 3000 °C (ref. 67)) could also achieve a significant improvement in rate capability.
Overall, graphite cathodes with a high degree of graphitization deliver superior capacity and rate performance in AGDIBs. From a sustainability perspective, liquid-phase ultrasonic exfoliation has been employed to recycle graphite for use as cathodes, where the resulting foam-like morphology with open structures and intrinsic defects facilitates efficient AlCl4− intercalation/deintercalation.59 Furthermore, spent graphitic cathodes can be regenerated through simple electrolysis,68 underscoring the recyclability and environmental friendliness of graphite-based cathode materials for AGDIBs.
Though edge-few/rich few-layer graphene or graphite, made from a low temperature (600 °C) process,72 plasma-etching,73 and N-doping,74–76 have been studied. The presence of structural defects in the graphene layer can deteriorate the charge/discharge performance of the cells by either decreasing the electrical connectivity or hindering AlCl4− intercalation. As seen in the Raman spectra (Fig. 8a), after N-doping, the few-layer graphene cathode failed to charge properly due to poor intercalation, as evidenced by the peaks remaining in the middle stage and not fully shifting to 1634 cm−1, as observed in pristine few-layer graphene. The charging process could not be completed, as the cell failed before reaching the maximum voltage. To take the most advantage of the excellent properties of graphene, a defect-free graphene sheet as large as possible is suggested for better electrochemical performance. Chen et al. prepared a defect-free few-layer graphene cathode with slightly stacked few-layer graphene sheets (4–10 layers) and an ID/IG value of almost 0 (GA 3000, Fig. 8b).77 The achievable specific capacity increases significantly with a higher graphitization degree. By changing the charging cutoff voltage, they found that structure decomposition (defects converted from the original sp2 carbon at a destructive charging cutoff voltage) caused the decline of cathodic capacity. They also prepared a defect-free graphene micro flower (ID/IG 0.0276) presenting a flower-like corrugated and fluffy morphology with numerous nanosized channels.78 It delivered a stable capacity of 100 mAh g−1 over 1000 cycles at both current densities of 10 and 20 A g−1. Later, a concept of “trihigh tricontinuous (3H3C)” was proposed to prepare graphite cathode.4 It requires an ideal cathode to have high quality, orientation, and channeling (3H) for local structures and a continuous electron-conducting matrix, ion-diffusion highway, and electro active mass (3C) for the electrode. The 3H3C cathode exhibited remarkable flexibility and impressive electronic conductivity (270
000 S m−1). Its vertical infiltration fissures, horizontal micrometer-sized gasbags, and nano-sized intervals collectively form interconnected 3D infiltration channels. These features contribute to a specific capacity of 111 mAh g−1 at 100 A g−1 over an impressive 250
000 cycles, maintaining a stable average discharge voltage and exhibiting low hysteresis across current densities ranging from 10 to 200 A g−1. The full cell exhibits stable performances at a wide temperature range (0 to 120 °C). In short, the structural integrity and electrical connectivity of a few-layer graphene cathode must be maintained. Chemical defects, such as heteroatom-dopants, lead to poor performance.
![]() | ||
| Fig. 8 (a) Part of the in situ Raman spectra of the pristine and N-doped few-layer graphene cathode during charge. Adapted with permission from Elsevier.75 Copyright 2019. (b) Charge and discharge profiles at a current density of 5 A g−1 of the cells with few-layer graphene aerogel (GA) cathodes which were carbonized under 2000 °C, 2500 °C and 3000 °C, respectively. The ID/IG values of the GA samples are 0.039 (GA 2000), 0.01 (GA 2500) and approximately 0 (GA 3000). Adapted with permission from John Wiley and Sons.77 Copyright 2017. | ||
:
1) electrolyte when tested in coin cells at 500 mA g−1. Interestingly, from the following literature, heteroatom doping has been found to unexpectedly enhance the electrochemical performance in AGDIBs, leading to significantly improved capacities. After N-doping, the electrical conductivity of the hard carbon xerogel (CX) decreased from 83 S m−1 to 62 S m−1 in the N-doped hard carbon xerogel (N-CX).79 N-CX delivers a remarkably high specific capacity of over 400 mAh g−1 at 100 mA g−1, whereas undoped CX exhibits only 10 mAh g−1 (Fig. 9a). Additionally, the CV curve of N-CX does not show distinct intercalation stages, unlike those seen in graphite materials. When the current density goes up to 500 mA g−1, the specific capacity of N-XC experienced a quick decay to around 60 mAh g−1. This suggests that the charge storage mechanism in these materials may differ fundamentally from graphite cathode.
![]() | ||
Fig. 9 (a) Charge/discharge profiles of carbon xerogel (CX) and N-CX in the 10th cycle at 100 mA g−1 were measured in an AlCl3/[EMIm]Cl (1.5 : 1) electrolyte using a Swagelok cell; insert: CV of N-CX at a scan rate of 1 mV s−1 reveals redox behavior without distinct intercalation stages. Adapted with permission from Wiley.79 Copyright 2024. (b) Charge/discharge profiles of cells with N-doped carbon (NC) at the 10th cycle at a current density of 500 mA g−1 in AlCl3/[EMIm]Cl (1.3 : 1) electrolyte using a Swagelok cell; insert: CV at 1 mV s−1. Adapted with permission from Royal Society of Chemistry.80 Copyright 2024. Charge/discharge profiles of cells employing N,S-doped carbon (N,S–C) cathode in the 100th cycle at a current density of 500 mA g−1 in AlCl3/[EMIm]Cl electrolyte using a Swagelok cell. Adapted from Li, J. et al., Energy Environ. Mater.81 Licensed under CC BY 3.0. ©2018 Wiley. | ||
Furthermore, an N-doped hard carbon (NC) cathode, generated from pomegranate peel (carbon base) and melamine (N source), delivered a high capacity of 323 mAh g−1 at 500 mA g−1 at the 1st cycle and 149 mAh g−1 after 200 cycles.80 The charge/discharge profiles of NC exhibit plateau-like regions (Fig. 9b), suggesting possible intercalation processes. This observation is supported by the CV curve, which reveals subtle but discernible intercalation-stage characteristics. Interestingly, during the discharge process, AlCl4− forms a bond with the pyridine nitrogen at the edge, and during the charging process, AlCl4− leaves, and Cl− forms a bond with the pyridine nitrogen at the time. This is also supported by similar results for N-XC,79 where a strong interaction between pyridinic/graphitic nitrogen and AlCl4− was observed. A study on AGDIBs with C3N4-derived layered N,S-doped carbon (N,S–C) cathode demonstrates that dual heteroatom doping significantly enhances electrochemical performance. DFT calculations reveal that N,S-co-doping reduces both the AlCl4− adsorption energy and the work function value of the carbon matrix, thereby facilitating the reactions with AlCl4− species at the electrode/electrolyte interface.81 Consequently, the cell exhibited a very high capacity of 410 mAh g−1 at a current density of 500 mA g−1 and maintained 329 mAh g−1 after 500 cycles. The charge/discharge profiles for the cells employing NC and N,S–C cathodes exhibit similar shapes, indicating a comparable intercalation/adsorption mechanism. But the capacitive contribution of N,S–C at a sweep rate of 1 mV s−1 is 56.7%, which is much lower than 96% of NC.
Overall, undoped hard carbon shows poor electrochemical performance in AGDIBs. However, certain doped hard carbon materials exhibit unexpectedly high capacities—surpassing those of graphite and few-layer graphene—though only at lower current densities. This improvement is primarily attributed to the lower adsorption energy introduced by heteroatom doping, which enhances anion storage. Additionally, the large surface area and high porosity of the doped hard carbon may contribute by promoting efficient electrolyte absorption and facilitating faster anion transport.82 Nonetheless, these factors alone cannot fully explain the significant capacity increase, suggesting that the energy storage mechanism of hard carbon is fundamentally different from that of graphite. However, the exact mechanism has not yet been clearly elucidated in current studies and remains to be further explored.
Accordingly, the electrochemical performance of AGDIBs is highly dependent on the reversible AlCl4− intercalation/deintercalation in cathode materials. Ex/In situ and operando characterizations such as Raman spectroscopy, X-ray diffraction, solid-state NMR spectroscopy, combined with theoretical modeling (e.g. density functional theory (DFT) and classical molecular dynamics (MD) simulations) have been applied to investigate AlCl4− intercalation.87–90 The interlayer space of graphite (0.335 nm) appears relatively narrow for the large ionic radius of AlCl4− (5.28 Å). However, due to its highly tunable interlayer spacing, graphite can accommodate AlCl4− anions, forming graphite intercalation compounds (GICs) through a unique staging mechanism that depends on the applied potential. GICs can be characterized by changes in d-spacing, which refers to the interplanar distance or gallery height between graphene layers. The intercalation stage index is defined as the number of graphene layers separating two adjacent intercalated layers91,92 (Fig. 10). In the initial stage, anion intercalation progresses through highly ordered, coordinated phases as more anions are inserted into the graphite layers. Once fully charged, with the stage reaching lower numbers, the degree of order is significantly higher, suggesting that the intercalated species likely rearranged themselves further. Notably, one stage does not completely disappear before the next one begins to form.
![]() | ||
| Fig. 10 Schematic image of the intercalation stage index showing the number of graphene sheets between two adjacent intercalated layers. | ||
Angell et al.93 studied the intercalation stage through in situ Raman spectra. Fig. 11a shows part of the Raman spectra. From the very beginning of the charging process, between 1.71 and 1.94 V, the G band of pristine graphite (1584 cm−1) split by 20 cm−1. This splitting arises from the rearrangement of positive charges—i.e., holes formed due to electron withdrawal from the graphite layers during intercalation. These holes alter the electronic environment within the graphite, leading to vibrational changes. As a result, the original band became two different E2g peaks, inner (i) and outer (b). The intercalation stage (n > 2) was then calculated based on the ratio of the intensities of these two peaks on following equation:
![]() | ||
Fig. 11 (a) Selected in situ Raman spectra of the graphite electrode recorded during charge/discharge at 50 mA g−1.93 (b) Selected SAXS (left) and WAXS (right) profiles of the first cycle along with charge/discharge at 50 mA g−1.92 (c) The relative specific capacity across intercalation stages during the first cycle based on publication: galvanostatic charging curve (25 mA g−1) of Swagelok type cell89 and PEEK self-made cell92 with AlCl3/[EMIM]Cl (1.5 : 1) electrolyte, galvanostatic charging curve (50 mA g−1) of pouch cell with AlCl3/[EMIM]Cl (1.7 : 1) at −10 °C,91 and galvanostatic charging curve (50 mA g−1) of pouch cell with AlCl3/urea (1.3 : 1) electrolyte.93 Adapted from Elia, G. A. et al., Adv. Funct. Mater.89 Licensed under CC BY. ©2020 Wiley. Adapted with permission from author. Copyright 2018.91 Adapted from Greco, G. et al., Small Methods.92 Licensed under CC BY 3.0. ©2021 Wiley. Adapted with permission from author.93 Copyright 2017. | ||
In contrast, Greco et al.86 calculated a stage 4 GIC in the fully charged state at 2.4 V through ex situ measurements using the following equation:
| Δd = Ic − 0.335n = di − 0.335(n + 1) |
Overall, there is a lack of consistency in the reported stage number of fully charged intercalation compounds across various studies, as shown in Fig. 11c. One key factor contributing to this discrepancy is the preferential intercalation that occurs at the surface of the cathode, rather than throughout the bulk. Surface-sensitive techniques, such as Raman spectroscopy, tend to detect higher degrees of intercalation, often corresponding to lower stage indices (e.g., stage 1 or 2 GICs) under fully charged conditions. In contrast, bulk-sensitive methods like XRD, which probe deeper into the material, typically observe lower intercalation levels, corresponding to higher stage indices (e.g., stage 3 or 4 GICs). This discrepancy underscores the importance of combining both surface and bulk characterization methods to achieve a more comprehensive and accurate understanding of the intercalation behavior in these materials. Beyond the choice of technique, factors such as porosity, applied voltage, temperature, and electrolyte composition also influence the degree of anion intercalation into graphite. Higher porosity creates more pathways for ion diffusion, while the applied voltage governs the electrochemical driving force for intercalation, with higher voltages promoting greater ion insertion but potentially causing instability at extreme levels. Both lower and potentially higher temperatures were found to improve performance in specific systems, but the underlying mechanisms need to be carefully optimized for stability and capacity. Additionally, electrolyte composition affects the availability and mobility of ions, with different electrolytes influencing the efficiency and stability of the intercalation process. Furthermore, meaningful comparisons of GIC levels across studies are hindered by variations in both the types of materials and the volumes employed.
A survey of recent work on graphitic carbon and hard carbon cathodes of AGDIBs is listed in Table 2. The best electrochemical performance in each year is compared. Generally, the specific capacities obtained for graphitic materials are mostly around 100 mAh g−1 and not more than 150 mAh g−1, but high rates and ultra-long-term cycling life are achieved. In hard carbon, the specific capacity reaches above 400 mAh g−1; however, only under lower current densities. Graphitization degree, defects, pore structures, and morphology play crucial roles in affecting electrochemical performance. Fast electron transportation is based on a high graphitization degree, while hetero atoms doping offers more active sites for surface adsorption of AlCl4−. Thus, defect-free few-layer graphene demonstrated the best high-rate performance. Hard carbon, especially containing heteroatoms, displayed the highest specific capacities. The pore size as well as the interlayer space need to be big enough for the intercalation of AlCl4− anions but avoid a large accumulation of ions in too big spaces. Particularly, taking advantage of unique morphologies (flakes/microflowers/open nanostructures) may result in a higher capacity and rate performance.
| Cathode | Electrolyte | Specific capacity (mAh g−1)/current density (mA g−1) | Long-term cycling | Year Ref. |
|||
|---|---|---|---|---|---|---|---|
| Specific capacity (mAh g−1) | Current density (mA g−1) | Capacity retention | Cycle number | ||||
| Graphite/few-layer graphene | |||||||
| Reduced graphene oxide | AlCl3/[EMIm]Cl 1.3 : 1 |
149/500 | 94.5 | 5000 | 95.8% | 10 000 |
2025 96 |
| Recycled graphite | AlCl3/[EMIm]Cl 1.3 : 1 |
138/2000 | 127 | 10 000 |
97.0% | 2000 | 2024 59 |
| Kish-based expanded graphite | AlCl3/[EMIm]Cl no ratio | 141/50 | 134 | 1000 | 99.1% | 10 000 |
2024 17 |
| Modified expanded graphite | AlCl3/[EMIm]Cl 1.3 : 1 |
118/1000 | 102 | 5000 | 100.0% | 10 000 |
2023 97 |
| Free-standing expanded graphite film | AlCl3/Et3NHCl 1.5 : 1 |
110/1000 | 100 | 5000 | — | 27 500 |
2022 67 |
| Graphene nanoplatelets | AlCl3/TMAHCl 1.7 : 1 |
134/2000 | 134 | 2000 | 98% | 3000 | 2020 98 |
| Natural graphite powder | AlCl3/[EMIm]Cl no ratio | 100/1000 | ∼60 | 20 000 |
70.0% | 500 000 |
2019 5 |
| Nanosheet-bricked porous graphite | AlCl3/[EMIm]Cl 1.3 : 1 |
103/2000 | 104 | 10 000 |
— | 3000 | 2018 66 |
| Carbon nanoscrolls | AlCl3/[EMIm]Cl 1.3 : 1 |
104/1000 | 101 | 50 000 |
100.0% | 55 000 |
2018 69 |
| Edge-rich few-layer graphene | AlCl3/[EMIm]Cl 1.3 : 1 |
128/2000 | 90 | 8000 | 99.2% | 20 000 |
2018 72 |
| Defect-free few-layer graphene | AlCl3/[EMIm]Cl 1.3 : 1 |
100/5000 | 97 | 50 000 |
97.0% | 25 000 |
2017 77 |
| “3H3C” graphite | AlCl3/Et3NHCl 1.5 : 1 |
116/300 000 |
120 | 400 000 |
91.7% | 250 000 |
2017 4 |
| Hard carbon | |||||||
| N-doped hard carbon | AlCl3/[EMIm]Cl 1.3 : 1 |
323/500 | 109 | 1000 | — | 1000 | 2024 80 |
| Layered N,S-doped carbon | AlCl3/[EMIm]Cl no ratio | 330/500 | 90 | 5000 | 97.0% | 10 000 |
2024 81 |
| Pomegranate peel-based carbon | AlCl3/[EMIm]Cl 1.3 : 1 |
125/200 | 82 | 150 | — | 500 | 2023 99 |
| Coconut shell-based porous carbon | AlCl3/[EMIm]Cl 1.3 : 1 |
150/100 | ∼80 | 1000 | 90.0% | 1500 | 2022 100 |
| N,P-doped carbon | AlCl3/[EMIm]Cl 1.3 : 1 |
110/500 | 98 | 5000 | ∼100% | 2500 | 2018 82 |
Overall, it requires a high graphitization degree, defect-free planes, high porosity, and high stability for graphitic materials to deliver optimal performance. However, it is important to note that the current cathode capacity of such materials may be approaching its maximum, limiting the potential for significant increases in energy density. Hard carbon exhibits a notably high specific capacity but suffers from limited rate capability, likely due to either (i) slow charge storage kinetics or (ii) different storage mechanisms compared to graphitic materials. This performance difference may stem from its unique properties, such as an amorphous microstructure, abundant defect sites, and complex pore networks that simultaneously enable high capacity while restricting rapid ion transport. To advance the development of high-performance AGDIBs, future studies should focus on two key objectives:
(1) Optimization of graphitic carbon cathodes for high-power applications
• Design strategies to enhance fast ion transport kinetics.
• Investigate electrolyte formulations that enable rapid anion intercalation.
(2) Fundamental understanding of hard carbon storage mechanisms
• Utilize advanced in situ/operando characterization techniques to uncover the dominant charge storage processes. A combined approach using both surface-sensitive (e.g., XPS, Raman) and bulk-sensitive (e.g., XRD, neutron scattering) methods is essential to capture the complex interfacial and structural evolution occurring during cycling.
• Establish correlations between material properties (porosity, heteroatom doping, defect density) and electrochemical performance.
• Develop quantitative models to distinguish between intercalation, adsorption, and pore-filling contributions.
This dual-focused approach will drive the design of next-generation carbon materials, combining the high-rate capability of graphitic structures with the exceptional capacity of optimized hard carbons.
It is essential to know that the Al deposition which takes place at the existence of Al2Cl7− ions can only occur in acidic conditions.107 Hence, the storage performance is highly dependent on the molar ratio (r) of AlCl3 to [EMIm]Cl. The anion speciation diagram of AlCl3/[EMIm]Cl is shown in Fig. 12. First, r should be above 1.0 so that an acidic condition is maintained. Besides, 2.0 is the highest value for this electrolyte to remain completely in a liquid phase. Note that, the ratio value is strongly related to the theoretical energy density of electrolyte; at ratio values of 2.0 and 1.3, the theoretical energy densities are 65–70 Wh kg−1 and 33 Wh kg−1, respectively.108 AlCl3/[EMIm]Cl (2.1
:
1) utilizing an excess of AlCl3 was reported.109 The battery possesses a high energy density of 59.1 Wh kg−1, cathodic capacity of 103 mAh g−1, and an average discharge voltage of 1.71 V. Nevertheless, an excess amount of AlCl3 crystallized in the glass fiber separator. Moreover, the higher the ratio, the more acidic the electrolyte becomes, making it highly corrosive to conventional battery components.110 A molar ratio of 1.3 provides an optimal balance of AlCl4− and Al2Cl7−, enabling efficient and reversible aluminum deposition while minimizing corrosion. Higher ratios increase acidity and Al2Cl7− concentration, which can compromise electrode stability and long-term cycling. Consequently, a ratio of 1.3 is most commonly employed in research, with 1.5 used less frequently.
![]() | ||
| Fig. 12 The anion speciation diagram of AlCl3/[EMIm]Cl at room temperature, illustrating the electrolyte compositions at 0 < r < 2. Adapted from Palacin, M. R. et al., J. Phys.: Energy.107 Licensed under CC BY 4.0. ©2024 IOP Publishing. | ||
The practical application of AlCl3/[EMIm]Cl is limited by its high cost, hygroscopicity, leakage risk, and strong corrosivity. To address these issues, alternative electrolytes have been investigated, primarily to verify stable Al deposition. One approach involves AlCl3-based ionic liquids prepared with cheaper Lewis basic ligands. For example, AlCl3/1-butyl-3-methylimidazolium chloride ([BMIm]Cl) enables AGDIB operation at subzero temperatures, with no phase transitions observed between 0 and −35.5 °C during heating or cooling.111 However, compared to AlCl3/[EMIm]Cl, this electrolyte exhibits a lower diffusion coefficient of Al2Cl7− due to higher viscosity, and the resulting thicker Al deposits promote dendrite formation.112,113 Another strategy is to use mixtures of organic cations to disrupt ion cluster formation and enhance ion mobility. In AlCl3−[EMIm]Cl/[BMIm]Cl (2
:
1), the battery showed improved low-temperature performance, with specific capacity retention increased by 24% at −20 °C and 9% at −40 °C relative to AlCl3/[EMIm]Cl.114
In addition, triethylamine hydrochloride (Et3NHCl)-based electrolytes derived from otherwise intractable industrial waste have also attracted increasing attention. These electrolytes are considered highly valuable due to their extremely low cost—potentially thousands of times cheaper than [EMIm]Cl—while still supporting reversible Al electrochemistry.20 An AlCl3/Et3NHCl (1.7
:
1) electrolyte shows a high ionic conductivity of 13.5 mS cm−1.115 Most importantly, both AlCl4− and Al2Cl7− anions have been detected in AlCl3/Et3NHCl.12 Mukundan et al.116 stated that the ratio of the concentrations of AlCl4− to Al2Cl7− ions (denoted as n(AlCl4−)/n(Al2Cl7−)) equals 1 in both AlCl3/Et3NHCl (1.5
:
1) and AlCl3/[EMIm]Cl (1.5
:
1) electrolytes. They further observed that, during long-term cycling, the potential of the Al anode decreases and the morphology of the graphite matrix evolves, leading to an impressive 52% increase in specific capacity for cell with AlCl3/Et3NHCl compared to AlCl3/[EMIm]Cl electrolytes. Moreover, a reduced self-discharge capacity loss of 3% (instead of 10%) within 24 h was achieved using AlCl3/Et3NHCl electrolyte and glassy carbon current collector.26 As shown in Fig. 13a, the CVs of cell with AlCl3/Et3NHCl and AlCl3/[EMIm]Cl electrolytes show similar peaks, but the redox signals of the one with AlCl3/Et3NHCl are shifted to lower potentials, indicating that the AlCl4− intercalation into the graphite interlayer in AlCl3/Et3NHCl electrolyte is stronger than in AlCl3/[EMIm]Cl electrolyte.117 Xu et al.20 found that AlCl3/Et3NHCl electrolyte showed a higher decomposition voltage, implying a wider potential window for AGDIBs. The cell could run under an upper cut-off voltage of 2.62 V with an CE of 96% over 6000 cycles, while the CE drops significantly to 89% when using AlCl3/[EMIm]Cl electrolyte. A similar cost-effective electrolyte, AlCl3/trimethylamine hydrochloride (TMAHCl), was also studied in AGDIBs. At a ratio above 1.6
:
1, the ionic conductivity of AlCl3/TMAHCl is close to AlCl3/[EMIm]Cl and AlCl3/Et3NHCl at the same ratio (Fig. 15a). It endowed the cell to operate across temperatures ranging from −10 °C to 60 °C, and a full battery with AlCl3/TMAHCl (1.7
:
1) electrolyte was estimated to deliver a higher cell-level specific capacity than AlCl3/Et3NHCl and AlCl3/[EMIm]Cl.98 However, the ionic conductivity and electrochemical performance of this electrolyte at ratios below 1.6 have not been provided, making it difficult to compare with other electrolytes at their typical ratios. Additionally, this electrolyte has not been further studied in any other reports. Consequently, AlCl3/[EMIm]Cl remains the benchmark electrolyte due to its well-balanced properties; AlCl3/Et3NHCl is gaining popularity as a lower-cost alternative. Nevertheless, researchers often continue using familiar systems (AlCl3/[EMIm]Cl), highlighting that practical adoption depends not only on raw performance but also on the challenges of modifying long-established electrochemical protocols.
![]() | ||
Fig. 13 (a) CV curves of Al//expanded graphite cells with AlCl3/[EMIm]Cl and AlCl3/Et3NHCl electrolytes117 and Al//natural graphite cells employing AlCl3/urea and AlCl3/AcA electrolytes125 (molar ratio 1.5 : 1) at a scan rate of 1 mV s−1. (b) Charge/discharge profiles of Al//natural graphite cells with AlCl3/[EMIm]Cl, AlCl3/urea and AlCl3/AcA electrolytes (molar ratio 1.5 : 1) at a current density of 50 mA g−1.125 Adapted with permission from Elsevier.117 Copyright 2019. Adapted with permission from Wiley.125 Copyright 2021. | ||
:
1 molar ratio was reported by Angell et al.93 They found that several graphite oxidation peaks and the corresponding reduction peaks in CV curves were easily correlated with the galvanostatic charge/discharge curve, indicating good reversibility. Previously, it was believed that, in an AlCl3/urea electrolyte, Al plating/stripping proceeds through two pathways, namely reactions of electroactive anionic (Al2Cl7−) and cationic (AlCl2·(urea)2+) species,118 where the cationic one dominates. However, later, operando Raman spectroscopy of the Al plating/stripping process revealed a large decrease in the Al2Cl7− peak (313 cm−1) and a corresponding increase in the AlCl4− peaks (350 and 445 cm−1).119 These results strongly suggest that deposition at the Al anode primarily proceeds via Al2Cl7− and the AlCl4− intercalation at the cathode in AlCl3/urea follows the same mechanism as in AlCl3/[EMIm]Cl.
However, AlCl3/urea has higher viscosity (88.9 mPa s−1 at r = 1.5) and low conductivity (1.17 mS cm−1 at r = 1.5),119 which limits its performance at room temperature. Increasing the AlCl3-to-urea molar ratio or operating at elevated temperatures can improve electrochemical behavior,120 but such conditions are generally impractical for standard battery applications. The battery capacity with an AlCl3/urea electrolyte was reported to be much lower than the one with AlCl3/Et3NHCl electrolyte.115 Applying AlCl3/urea derivatives (e.g. N-methyl urea (Me-Ur) and N-ethyl urea (Et-Ur)) was found to greatly reduce the viscosity (by a factor of three) and increase ionic conductivity (by 40%), leading to improved coulombic efficiencies.119 Ternary electrolytes in proper molar ratios show improved performance. The battery performance with ternary AlCl3/Urea/[EMIm]Cl (13.5
:
9
:
X, X = 0, 0.6, 0.8, 1) electrolyte is improved with increasing [EMIm]Cl content.121 In situ XRD confirmed AlCl4− intercalation/deintercalation in graphite, but the broad shoulder of the peak of the fully discharged graphite indicates residual disorder and strain. Furthermore, the impact of the electrolyte composition on the Al electrodeposition mechanism has been analyzed.122 Decreasing the [EMIm]Cl concentration in ternary AlCl3/[EMIm]Cl/urea (60
:
40 − X
:
X) electrolytes causes uniformly electrodeposited Al crystals to grow larger. When X ≥ 30, Al nanoplatelets (width of 10–15 μm) oriented in the (111) direction spontaneously formed. A small amount of urea in AlCl3/Urea/[EMIm]Cl (1.3
:
X
:
1 − X, X = 0.25) greatly improved Al platting/stripping down to −40 °C.123 AlCl3/urea/caprolactam(CPL) was applied in AGDIBs, achieving a high initial specific capacity of 151 mAh g−1 at 5000 mA g−1 and retained 132 mAh g−1 after 3000 cycles.124 Cells with an AlCl3/acetamide (AcA) electrolyte were found to exhibit very similar peaks to the AlCl3/Urea electrolyte in CV plots (Fig. 13a).125 The redox signals of AlCl3/AcA shifted to lower potentials compared to AlCl3/Urea, corresponding to lower potential plateaus during charging/discharging (Fig. 13b). However, cells with AlCl3/AcA present a much lower specific capacity in comparison to AlCl3/urea in AGDIBs. Incorporating an F-substituted AcA additive into the AlCl3/AcA electrolyte can induce the formation of an F-rich interphase layer, which stabilizes Al plating/stripping and accelerates its kinetics.126 On the contrary, a comparative study revealed that a cell with AlCl3/AcA possessed a higher specific capacity than the one with commercial AlCl3/[EMIm]Cl,127 highlighting the importance of standardized testing conditions for objective electrolyte evaluation.
Fig. 14a and b compare the oxidation potential of various aluminate chloride-containing electrolytes and the specific capacity in Al//graphite full cells, respectively. The decomposition voltage of these electrolytes (including AlCl3/[EMIm]Cl) is generally below 2.5 V, except AlCl3/AcA with a dFAcA additive (2.78 V), AlCl3/[EMIm]Cl with an LiBOB additive (2.74 V), AlCl3/[BMP]Cl (2.83 V), and AlCl3/[BMIm]Cl (2.5–2.6 V depends on the molar ratio and temperature). The additives render electrolytes more stable at higher voltage, which benefits battery performance. Though AlCl3/AcA + dFAcA was only tried in an Al//sulfur cell, the existence of Al2Cl7− and AlCl4− anions should allow its application in AGDIBs. AlCl3/[BMP]Cl shows the highest voltage, unfortunately, the full cell specific capacity was found to be 0 because of the inability of Al electrodeposition. AlCl3/urea/CPL delivers the highest specific capacity, which may be due to its higher concentrations of Al2Cl7− and AlCl4− anions. However, the electrochemical performance depends not only on the electrolyte but also on various cell parameters. Therefore, differences in performance between studies may not only result from the electrolyte but also from the use of different cell components and configurations.
![]() | ||
Fig. 14 (a) The oxidation onset potentials of various AlCl3/[X]Cl and other alternative electrolytes.114,126 (b) Specific capacities of Al/graphite cells with various AlCl3/[X]Cl and other alternative electrolytes. AlCl3/[X]Cl (X = EMIm to DiPIm : EMIm 1 : 2) at 0.01 A g−1,114 AlCl3/urea/CPL and AlCl3/CPL at 5 A g−1,124 AlCl3/TMAHCl at 2 A g−1,98 AlCl3/NaCl/KCl at 1 A g−1,128 AlCl3/Et3NHCl at 5 A g−1,117 AlCl3/[EMIm]Cl at 1 A g−1,14 and AlCl3/urea at 0.1 A g−1.93 Adapted with permission from American Chemical Society.114 Copyright 2023. Adapted with permission from John Wiley and Sons.126 Copyright 2023. | ||
Since the corrosive nature of electrolytes is primarily caused by Al2Cl7− ions, some researchers have attempted to explore Cl-free electrolytes. Particularly, aluminum trifluoromethanesulfonate (Al(OTF)3)−based electrolytes, which may be capable of fast Al3+ transport and reduced corrosivity, have been investigated. Unfortunately, it was found that these non-corrosive electrolytes lead to unexpected challenges related to the reversibility of the redox reactions.129 No Al plating/stripping occurs on the Al electrode, which is believed probably due to the low corrosivity and the resulting inability to consume the Al2O3 passivation layer.130,131 Wang et al.132 found pitting corrosion of the Al foil in an Al(OTF)3−based electrolyte after it had been previously immersed in an AlCl3/[BMIm]Cl (1.1
:
1) electrolyte; however, no evidence was provided regarding the Al plating/stripping process. Overall, Cl-free electrolytes offer a promising route toward addressing the corrosivity of electrolytes, although significant challenges remain regarding the reversibility of Al plating/stripping.
Recently, a low-cost, non-corrosive, and air-stable, chloride-free Al-based hydrated eutectic electrolyte was synthesized composed of aluminum perchlorate nonahydrate (Al(ClO4)3·9H2O) and methyl urea (MU) ligand.133 The stability of Al plating/stripping was evaluated in symmetrical cells. The cell ran over 150 h with a steady voltage hysteresis, and the formation of Al deposits was observed on the surface of the Al foil after cycling. In addition, a symmetrical cell with two pieces of titanium (Ti) metallic foil was assembled to further evaluate Al plating. Al uniformly distributed on the cycled Ti electrode, confirming the Al deposition. Although this electrolyte was only utilized in Al//V2O5 full cells, its potential application in AGDIBs is worth further exploration.
From an experimental perspective, in most cases, a comparison to AlCl3/[EMIm]Cl under the same battery conditions has not been conducted, hence making it difficult to identify the advantages of alternative electrolytes. According to a literature survey on the pretreatment of electrolytes for AGDIBs, to date, only 1% dried and purified electrolytes before cell assembly, 20% only dried but not purified, and surprisingly 80% without mentioning this process. Obviously, the conditions of electrolytes, specifically moisture and purity, have significant impacts on cell performance. Therefore, careful attention to the electrolyte preparation is essential. Highly acidic electrolytes can corrode cell components, while the concentration of acidic anions may limit capacity, making it crucial to achieve a balanced formulation for optimal performance. The development of low-cost, high-performance electrolytes—particularly those that are non- or less corrosive—remains a key research priority.
![]() | ||
Fig. 15 (a) Ionic conductivity of various AlCl3/X electrolytes. X = [EMIm]Cl (10–20 mS cm−1), TMAHCl (average ∼7.5 mS cm−1), Et3NHCl (∼9.6 mS cm−1), [BMIm]Cl (∼9.4 mS cm−1), and urea (∼1.3 mS cm−1) were measured at 25 °C.98 [EMIm]Cl-PAM (1.66 × 10–3 S cm−1, measured at 20 °C) is gel polymer electrolyte made from AlCl3 complexed acrylamide and AlCl3/[EMIm]Cl (1.5 : 1).134 [EMIm]Cl-PEO-SiO2 (13 mS cm−1) is gel polymer electrolyte made from PEO, AlCl3/[EMIm]Cl (1.5 : 1) and an additive-fumed SiO2.2 [EMIm]Cl-PA (5.89 to 7.24 mS cm−1, measured at room temperature) and Et3NHCl-PA (3.86 mS cm−1, measured at 25 °C) is PA-based gel polymer electrolyte made with AlCl3/[EMIm]Cl (2 : 1 to 1.3 : 1)138 and AlCl3/Et3NHCl (1.7 : 1),139 respectively. AlCl3/Et3NHCl (1.5 : 1) add CTAC (12.06 mS cm−1)33 and F127 (9.31 mS cm−1)32 are named Et3NHCl + CTAC and Et3NHCl + F127, respectively. (b) CV curves of Al//natural graphite cells with AlCl3/[EMIm]Cl and AlCl3/[EMIm]Cl-PEO-SiO2 electrolytes showing anions intercalation at a scan rate of 1 mV s−1. Adapted from Leung, O. M. et al., Adv. Energy Mater.2 Licensed under CC BY 3.0. ©2024 John Wiley and Sons. (c) CV curves of Al//natural graphite cells with AlCl3/[EMIm]Cl-PA138 (not mentioning the scan rate) and AlCl3/Et3NHCl-PA139 (without air not mentioning the scan rate, and after being exposed to the air for 10 min at a scan rate of 1 mV s−1) presenting Al plating/stripping. Adapted with permission from Elsevier.138 Copyright 2021. Adapted with permission from American Chemical Society.139 Copyright 2021. | ||
Schoetz and coworkers137 found that Al can be deposited from a Lewis neutral ion gel (based on polyethylene oxide (PEO) and AlCl3/[EMIm]Cl) without any dendrite growth within a wide potential stability window of 5 V. Building on this, a solid polymer electrolyte was fabricated using a solvent-free method with PEO, AlCl3/[EMIm]Cl (molar ratio = 1.5), and an additive-fumed silicon dioxide (SiO2).2 This electrolyte (AlCl3/[EMIm]Cl-PEO-SiO2) maintained a high ionic conductivity of 13 mS cm−1, which is higher than AlCl3/[EMIm]Cl (Fig. 15a). When the charging cut-off potential is increased to 2.8 V, the specific capacity increases by approximately 48%, reaching 194 mAh g−1 after three cycles representing a 58% improvement compared to the ionic liquid. Meanwhile, due to the decreased ion mobility, the CV curves of the cells with AlCl3/[EMIm]Cl-PEO-SiO2 do not exhibit distinct redox peaks, unlike those of AlCl3/[EMIm]Cl (Fig. 15b).
Polyamide (PA), another efficient polymer, was used as the polymer matrix for AlCl3/[EMIm]Cl and AlCl3/Et3NHCl ILs. During the fabrication of the gel polymer electrolyte, the strong hydrogen bonds in the PA chains are initially broken down in ILs and then reorganize into a configuration where the molecules are connected, increasing the polymer's flexibility. AlCl3/[EMIm]Cl-PA with an ionic conductivity of 6.61 mS cm−1 was prepared through a solvent-free procedure enabling the Al//graphite cells to cycle well in a wide temperature range from −30 °C to 45 °C.138 Simply adding PA into the acidic AlCl3/Et3NHCl through stirring and heating, a AlCl3/Et3NHCl-PA was obtained after cooling.139 However, the ionic conductivity of AlCl3/Et3NHCl-PA is slightly lower than AlCl3/[EMIm]Cl-PA at the same IL molar ratio (Fig. 15a). After being exposed to air, reversible Al plating/stripping was still observed in both AlCl3/Et3NHCl- and AlCl3/[EMIm]Cl-PA gel polymer electrolytes (Fig. 15c). The Al//graphite cell demonstrated excellent interfacial stability, operating for over 3800 hours. Remarkably, even when partially severed, the remaining section was still capable of lighting a red LED lamp—despite being folded—without any electrolyte leakage. In situ XRD analysis revealed the presence of a stage 3 GIC at the fully charged state. The melt flow index of the PA-based AlCl3/Et3NHCl gel polymer electrolyte increased nearly linearly with AlCl3 concentration, indicating reduced molecular cohesion. This behavior likely arises from the preferential formation of bidentate coordination between Al atoms and the O and N atoms in the polymer, which suppresses interchain hydrogen bonding and lowers cohesion.140 As a result, PA-based gel polymer electrolytes significantly improve cell stability by enhancing moisture resistance and reducing the risk of leakage and corrosion.
In terms of ionic conductivity, polymer matrices based on PEO and PA outperform those using acrylamide. While gel polymer electrolytes address key safety concerns, their performance is still limited by a narrow operational window at higher current densities. For practical implementation in high-power systems, substantial improvements in electrochemical performance are necessary. Key challenges that remain include limited rate capability, a lack of long-term cycling data, and poor interfacial compatibility under high-rate conditions.
Table 3 summarizes the various electrolytes reported for Al||graphite batteries over the years, including ionic liquids, deep eutectic electrolytes (DEEs), and gel polymer electrolytes. Among them, AlCl3/[EMIm]Cl remains the most widely employed electrolyte due to its well-balanced electrochemical performance since 2015, and has been the benchmark system for AGDIBs. However, its practical application is hampered by critical drawbacks, such as high cost, corrosiveness, hygroscopicity, and high viscosity, which restrict material compatibility and limit rate performance, particularly under low-temperature conditions.
| Electrolyte | Ionic conductivity | Test voltage range | Cell performance (mAh g−1/mA g−1, capacity retention after cycles) | Year Ref. |
|---|---|---|---|---|
| “—” means not mentioned. | ||||
| ILs and DEEs | ||||
AlCl3/urea/[EMIm]Cl 1.3 : 0.25 : 0.75 |
— | 0.2 V to 2.4 V | 110/60 | 2024 123 |
AlCl3/acetamide 1.4 : 1 |
— | 0.3 V to 2.4 V | 101/50; 80/500, 92% retention after 1000 cycles | 2024 127 |
AlCl3/Et3NHCl 1.5 : 1 |
10 mS cm−1 (ref. 98) | 0.5 V to 2.5 V | 94.2/1000, 157% retention after 1000 cycles | 2023 116 |
AlCl3/[EMIm]Cl–[BMIm]Cl 2 : 1 |
13.8 mS cm−1 | 0.5 V to 2.45 V | 125/10 | 2023 114 |
AlCl3/urea 1.5 : 1 |
1.3 mS cm−1 (ref. 98) | 0.3 V to 2.4 V | 50/2000, 80% retention after 8000 cycles | 2021 125 |
AlCl3/[EMIm]Cl 1.3 : 1 |
12 mS cm−1 (ref. 98) | 0 V to 2.35 V | 120/100 | 2021 109 |
AlCl3/caprolactam (1.85 : 1)–9.5% urea |
— | 0.5 V to 2.6 V | 161/1000; 132/5000, 87% retention after 3000 cycles | 2020 124 |
AlCl3-TMAHCl 1.7 : 1 |
7.5 mS cm−1 | 0.5 V to 2.4 V | 134/2000, 167% retention after 3000 cycles | 2020 98 |
AlCl3/NaCl/KCl 61 : 26 : 13 |
— | 0.5 V to 2.35 V | 125/100, 90/4000, 100% retention after 300 cycles | 2018 128 |
| Gel polymer electrolytes | ||||
| AlCl3/[EMIm]Cl–polyethylene oxide–SiO2 | 13 mS cm−1 | 0.5 V to 2.8 V | 194/66 | 2024 2 |
AlCl3/Et3NHCl (2.0 : 1)–polyamide |
0.15 mS cm−1 | 0.8 V to 2.35 V | 31/20 | 2023 140 |
AlCl3/Et3NHCl (1.7 : 1)–polyamide |
3.86 mS cm−1 | 1.0 V to 2.45 V | 90/100; 94.6/200, % retention after 2000 cycles | 2021 139 |
AlCl3/[EMIm]Cl (1.3 : 1)–polyamide |
6.61 mS cm−1 | 1.0 V to 2.45 V | 94/500, 98.9% retention after 4200 cycles | 2021 138 |
AlCl3/Et3NHCl (1.7 : 1)–polyacrylamide |
5.77 mS cm−1 | 0.3 V to 2.3 V | 91/1000, 97.8% retention after 800 cycles | 2019 136 |
To address these limitations, a variety of alternative electrolytes have been investigated recent years, such as AlCl3/urea, AlCl3/acetamide, and AlCl3/Et3NHCl, as well as mixed ionic liquid systems. These systems generally offer reduced corrosiveness, lower cost, and improved safety, with some (e.g., AlCl3/urea) achieving ultralong cycling stability (>8000 cycles with 80% retention). In parallel, gel polymer electrolytes have emerged as promising solid-like electrolytes that mitigate leakage and corrosion. Although their ionic conductivities are generally lower than those of liquid systems, several formulations already deliver remarkable durability (>98% retention after 4000 cycles), highlighting their potential for safer and more robust AGDIBs.
Overall, as a benchmark electrolyte, AlCl3/[EMIm]Cl offers balanced performance but suffers from high cost, strong corrosiveness, hygroscopicity, and high viscosity. Consequently, the development of cost-effective and less corrosive alternatives has become a key research priority.
• Alternative chloroaluminate electrolytes (e.g. AlCl3/Et3NHCl, AlCl3/urea and its derivatives): show reduced toxicity and improved affordability, yet still face challenges in long-term stability and efficiency.
• Gel polymer electrolytes: enhance safety and structural integrity while achieving promising cycling stability, though the ionic conductivity remains relatively low.
Thus, developing cost-effective, less corrosive, and durable electrolytes that combine high ionic conductivity with long-term stability and mechanical robustness is essential for the practical advancement of AGDIB technology.
![]() | ||
| Fig. 16 (a) The oxidative stability, in terms of the voltage versus Al/Al3+, of Ni–S/Ni,143 MXene and other various current collector materials144 in AlCl3/[EMIm]Cl electrolyte with molar ratios of 1.3 (left) and 2.0 (right). Adapted from Wang, S. et al., Adv. Sci.108 Licensed under CC BY 4.0. ©2018 Wiley. (b) Number of publications reporting various binders used in AGDIBs based on literature survey (‘/’ indicates not mentioned). (c) Cell configurations applied in AGDIBs from the literature survey. | ||
Research focusing on current collectors is scarce, with less than ten publications identified. Given that Ni can be easily corroded in AlCl3/[EMIm]Cl electrolyte, it is important to point out that it might not affect the cell performance since no electrochemical activity from Ni was observed in blank electrode tests.4,82 A low-cost Ni–S/Ni current collector was made by electroplating amorphous Ni3S2 onto Ni foil using a scalable electrodeposition method.143 Amorphous Ni3S2 has excellent flexibility, high oxidation potential (2.7 V) and conductivity (2.3 × 106 S m−1) which are comparable to Ta current collectors (2.7 V, 8 × 106 S m−1, respectively). A full cell with a Ni–S/Ni current collector shows similar long-term performance to the Ta current collector. Considering the high price of Ta, a Ni–S/Ni current collector can serve as an alternative. Wang et al.108 fabricated a flexible current collector with an electrical conductivity of 4–4.5 × 106 S m−1 by magnetron sputtering titanium nitride (TiN) on polyimide substrates. Stable cathodic operation is observed at voltages of up to 2.5 V versus Al3+/Al. The nitrogen atoms are believed to shield the underlying metal ions, protecting them from oxidation, which results in better stability than the pure metals. At a current density of 10 A g−1, the power density is estimated to be 4500 W kg−1. It is a highly promising alternative; however, the relatively complex synthesis has hindered its application beyond the initial study. Guan et al.144 utilized MXene (Ti3C2Tx film) as a current collector. It indicates that the OH− and COO− groups incorporated onto the surface of Ti3C2Tx film could stop AlCl4− ions from fully reaching the surface of the current collector, thus preventing the cathode materials from falling off the current collector because of its volume expansion. However, MXene materials have drawbacks, such as complex synthesis processes, sensitivity to air and moisture, and potential issues with electrolyte compatibility. Most importantly, MXenes exhibit high surface reactivity and can interact with metal components, such as aluminum, potentially leading to metal dissolution in the battery.
In summary, the stability of current collector is strongly influenced by the corrosive nature of electrolytes. Molybdenum is the most widely used due to its high conductivity and relative stability, though it can still undergo oxidation and contribute to capacity loss. Alternatives such as Ta, Ni–S/Ni, TiN-coated substrates, and MXene films offer promising conductivity and electrochemical stability, but each has trade-offs related to cost, synthesis complexity, or compatibility with electrolytes.
According to available publications, only one research on separators was reported. A separator based on polyacrylonitrile (PAN) was fabricated by a simple electrospinning technique.145 In the AlCl3/[EMIm]Cl (1.5
:
1) electrolyte, a highly uniform Al deposition was observed. When tested at a low current density of 25 mA g−1, the battery employing a PAN separator achieved a specific capacity of 73.5 mAh g−1, slightly outperforming the 68.4 mAh g−1 obtained with a glass fiber separator. Furthermore, the battery with the PAN separator demonstrated significantly enhanced performance at 50 °C compared to 25 °C under a higher current density of 200 mA g−1. In this study, six commercially available separators were evaluated: (i) polypropylene (PP) monolayer, (ii) glass fibers, (iii) cellulose/PAN blended fibers, (iv) polyethylene (PE)/PP coated with poly (vinyl alcohol) (PVA), (v) open-structure PP micro- and nanofibers, and (vi) polyimide microfibers. These were immersed in the AlCl3/[EMIm]Cl electrolyte for seven days to assess the chemical stability. Among them, only the glass fiber separator showed good compatibility, exhibiting minimal mass loss (1.2%) and no structural degradation. However, due to the limited data available on separator performance in AGDIBs, further systematic studies are needed. In summary, glass fiber is the most widely used and chemically compatible separator, though its low mechanical strength and high thickness limit practical applications. PAN-based separators, particularly those prepared via electrospinning, have shown promising performance, including uniform Al deposition and improved high-temperature stability, indicating strong potential for future development.
PTFE was reported to be insoluble in AlCl3/[BMIM]Cl (mole ratio 1.1
:
1) electrolyte,149 but its insolubility has not been confirmed in AlCl3/[EMIm]Cl electrolyte. It is challenging to maintain consistent mass and thickness of the positive electrodes while using PTFE as a binder, which typically results in high mass loading. Additionally, a high graphite loading can lead to lower capacity, as the increased electrode volume hinders the smooth diffusion of AlCl4− ions into the host material.150 Nevertheless, the using PVDF or PTFE binders typically requires N-methylpyrrolidone (NMP), a solvent that is toxic and moisture-sensitive.151 In contrast, water-soluble CMC and sodium alginate avoid NMP. These binders are low-cost, environmentally friendly, and demonstrate good thermal stability and long-term cycling stability. Rich in carboxyl and hydroxyl groups, they are expected to interact with functional groups on the surface of graphitic materials, forming an ideal gel phase network in the electrode. At the same time, polar/ionic interactions may occur between these functional groups and anions from electrolyte. Graphite electrodes using sodium alginate as the binder exhibited higher cycling stability compared to those based on PVDF at a current density of 1000 mA g−1.152 However, the solubility or potential side reactions of CMC and sodium alginate in AlCl3/[EMIm]Cl electrolyte remain insufficiently studied and warrant further investigation. Overall, the choice of binder should be more critically examined in the future.
000 cycles72), use polypropylene liners that have been shown to corrode in Al-based electrolytes.145 The long-term stability of this casing remains uncertain. Swagelok-type cells are ideal for lab-scale evaluations and closely mimic the environments of coin and pouch cells; however, their PTFE construction may suffer from corrosion, as fluorinated polymers are not always electrochemically inert.153 Indeed, we found that initial cycling tests (100 cycles) revealed that Swagelok-type cells were less stable than coin cells. PEEK-based cells, though rarely used, offer promising features such as excellent corrosion resistance, high thermal stability, and mechanical durability, making them attractive for long-term cycling applications. Nevertheless, the high cost and manufacturing complexity of PEEK materials remain significant barriers to widespread adoption. In conclusion, while various casing materials offer specific benefits, none provide a fully optimal solution, highlighting the need for continued material innovation and systematic evaluation to ensure long-term stability and compatibility with corrosive electrolytes in AGDIB systems.
| This journal is © The Royal Society of Chemistry 2025 |