Next Article in Journal
The Strains Enterococcus faecalis as Contaminants of Raw Goat Milk and Their Treatment with Postbiotic Active Substances Produced by Autochthonous Lactococci
Previous Article in Journal
Investigation on the Flow and Solidification Characteristic of Steel During Continuous Casting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Graphene-Stabilized δ-MnO2 Cathode for High-Capacity Aqueous Aluminum-Ion Batteries

1
Electrochemistry and Electroplating Group, Technische Universität Ilmenau, 98693 Ilmenau, Germany
2
Department of Materials Engineering, Tarbiat Modares University, Tehran 14115111, Iran
*
Author to whom correspondence should be addressed.
Processes 2025, 13(11), 3551; https://doi.org/10.3390/pr13113551
Submission received: 25 September 2025 / Revised: 24 October 2025 / Accepted: 30 October 2025 / Published: 4 November 2025

Abstract

Aluminum-ion batteries (AIBs) are emerging as promising alternatives to lithium-ion batteries due to their cost-effectiveness and resource abundance. However, their practical application is hindered by low capacity, poor cycle life, and limited rate capability. In this study, δ-MnO2 and δ-MnO2/Graphene composite cathodes are developed and tested in aqueous aluminum-ion batteries (AAIBs) using a mixture of 0.5 M Al2(SO4)3 and 0.4 M MnSO4 as the electrolyte. The electrochemical properties are evaluated alongside treated aluminum (TAl) and Zn–Al alloy anodes. Among the configurations tested, the δ-MnO2/Graphene|0.5 M Al2(SO4)3, 0.4 M MnSO4|Zn–Al system showed the best performance, achieving a high discharge voltage of 1.63 V, a specific capacity of 746 mAh g−1, and excellent cycling stability over 352 cycles. The stabilizing effect of graphene, due to increased oxygen vacancies and the formation of Mn–O–C bonds, enhances electron and ion transport, thereby improving cathode integrity and the overall performance of the AAIB. Additionally, the Zn–Al alloy anode extends the battery’s cycle life compared to the TAl anode. This work demonstrates the potential for low-cost, high-performance AAIBs, paving the way for more sustainable and scalable energy storage solutions.

1. Introduction

AIBs have garnered significant interest as viable substitutes for LIBs due to the abundance of aluminum, low cost, and high theoretical capacity (2980 mAh g−1) [1,2]. By leveraging the trivalent redox chemistry of aluminum, AIBs offer a capacity that far surpasses lithium-ion counterparts, establishing them as a prominent candidate for future energy storage innovations. AAIBs further enhance this potential by providing improved safety, environmental compatibility, and economic feasibility, making them particularly suitable for extensive applications. Unlike systems that rely on expensive ionic liquids, which primarily provide monovalent anions such as AlCl4 and Al2Cl7, AAIBs harness the trivalent Al3+ ions, offering the potential for higher capacity and significantly improved overall cost-efficiency [3].
Despite these advantages, the practical realization of AAIBs faces several critical challenges. The presence of a passivating layer of aluminum oxide on the anode is a major problem because it prevents effective electrochemical reactions. Moreover, the hydrogen evolution side reaction, which happens before the aluminum deposition potential (−1.66 V vs. SHE) at the aluminum anode [4], reduces the overall efficiency and stability of the battery system. Another key limitation is the scarcity of high-performance cathode materials. A variety of approaches have been explored to tackle the issues associated with AAIBs. On the anode side, surface modification of aluminum [5,6,7,8,9] and the incorporation of aluminum alloys [1,3,8,10,11,12] have shown potential in improving anode compatibility with aqueous electrolytes, thereby mitigating corrosion and passivation issues. On the cathode side, different oxide materials such as vanadium oxides [13,14,15,16] (e.g., V2O5, FeVO4, VOPO4), bismuth oxide [17] (e.g., Bi2O3), and tungsten oxide [18] (e.g., WO3) have been reported in AAIBs. Among them, manganese oxide-based cathodes [5,19,20] have demonstrated auspicious performance due to their mixed charge storage mechanisms, combining Al3+ intercalation and capacitive surface reactions; their structural tunability through doping or nano structuring; and the formation of stable surface layers during cycling, which can enhance electrochemical reversibility [21]. Although manganese-based oxides have shown promising performance in AAIBs [1,5,6,8,11,12,22,23], their poor cycling stability in energy systems remains a challenge due to the dissolution of manganese species into the aqueous electrolyte [24,25,26,27], necessitating effective stabilization strategies. Stabilizing MnO2 by inhibiting the Jahn-Teller effect [28], creating oxygen vacancy (VO) [22,29,30], and using dopants [11] can further enhance its electrochemical performance. Another practical approach involves integrating cathode materials with carbonaceous materials, such as graphene, to overcome their inherently low conductivity and structural instability during cycling [31,32,33]. The graphene not only enhances electron transport but also prevents structural degradation and increases the cathode’s surface area, all of which contribute to the improved performance in manganese oxide-based energy systems [34,35,36,37,38,39,40,41]. In our previous work [42], we demonstrated that coupling MnO2 with graphene effectively enhanced electrochemical performance in LIBs. Although the benefits are apparent, the MnO2/graphene cathode has not yet been studied in AAIBs, suggesting that it is a promising avenue for further AAIB development.
Closing this gap, this study proposes a system comprising a stabilized manganese oxide/graphene composite cathode, a zinc-aluminum alloy anode, and a cost-effective fluorine-free aqueous electrolyte consisting of 0.5 M Al2(SO4)3 and 0.4 M MnSO4. To the best of our knowledge, this is the first AAIB to use an MnO2-graphene cathode. The configuration aims to enhance cycling stability and energy density, thereby supporting the development of high-performance AAIBs for practical, real-world energy storage applications.

2. Materials and Methods

2.1. Synthesis of Cathode Materials

The δ-MnO2 and δ-MnO2/Graphene were synthesized using the method described in our previous work [42]. A 4 mL aqueous solution containing 0.300 g Mn(NO3)2·4H2O (≥99.9%) was mixed with an 8 mL aqueous solution of 0.190 g NaOH (≥99.0%) and 0.530 g of H2O2 (30 wt.%) under stirring for 20 min. Subsequently, the resulting precipitate was mixed with 30 mL of 2 M NaOH (all chemicals sourced from Merck, Germany). The mixture was heated for 16 h at 150 °C in a Teflon-lined autoclave [43]. After cooling, the product was collected by centrifugation, thoroughly washed with deionized water until a neutral pH was achieved, and subsequently dried under ambient conditions.
For δ-MnO2/Graphene synthesis, 20.0 mg graphene (purity >98%, United Nanotech Innovations Pvt Ltd., India) was first ultrasonically dispersed in water, and 0.300 g Mn(NO3)2·4H2O was added. The graphene–Mn-containing solution was then added to 8 mL of an aqueous solution containing 0.190 g NaOH and 1.080 g H2O2. After stirring for 20 min, the resulting mixture was again combined with 30 mL of 2 M NaOH. The final mixture was transferred to the autoclave and subjected to hydrothermal treatment at 150 °C for 2 h. The product was collected, washed, and dried following the same procedure.

2.2. Battery Assembly

Aluminum foils (≥99.3%, 16 μm thick) were surface-treated with an ionic liquid (IL) in an argon-filled glovebox. The foils were submerged in AlCl3:[EMIm]Cl (1.5:1) for ten days. Before assembling the batteries, the treated foils were wiped clean with tissue paper inside the glovebox. The aqueous electrolyte was prepared by dissolving Al2(SO4)3·xH2O (x = 14–18, ≥97%, Thermo Fisher Scientific, USA) and MnSO4.4H2O (99%, Thermo Fisher Scientific, USA) in deionized water. δ-MnO2 and δ-MnO2/Graphene cathodes were fabricated by mixing the active material, carbon black (KS6L, TIMCAL, Switzerland), and polyvinylidene fluoride (PVDF, Solef-5130, Solvay, Belgium) in a weight ratio of 7:1.5:1.5 with N-Methyl-2-Pyrrolidone (NMP). After applying the prepared slurry to a 0.2 mm thick piece of 316L stainless-steel foil, it was dried for 16 h at 80 °C in an electric oven. The battery components, consisting of the prepared cathode, TAl or zinc foil (≥ 99.95 percent, 0.24 mm thick—used as a substrate to form Zn-Al anode in situ in the battery), and a glass microfiber filter (Whatman 1822-025, Cytiva, UK) as the separator, were assembled in Swagelok cells with stainless-steel current collector rods (12 mm diameter) under ambient air condition. To prevent corrosion, the rod side surfaces were completely insulated, exposing only the cross-section for electrode contact.

2.3. Electrochemical Measurements

A VMP3 multichannel potentiostat (BioLogic, France) was used for all electrochemical measurements. Galvanostatic charge/discharge and cycling experiments were conducted at 180 mA g−1 with charge capacity regulation instead of a voltage limit to prevent probable side reactions and due to the lack of an explicit voltage cutoff. Each measurement sequence began with a discharge step. For anode voltage investigation, a three-electrode system with a Swagelock cell was employed, comprising a Hg/Hg2SO4 reference electrode, the fabricated working electrodes, and either aluminum or zinc foil as the counter electrode.
Full cells and a three-electrode setup were used for cyclic voltammetry (CV) tests. A platinum counter electrode, a glassy carbon working electrode with a 3 mm diameter (BASi, UK), and a reference electrode of Hg/Hg2SO4 were employed in the three-electrode setup. CV scans were conducted for full cells between 0.1–1.5 V for TAL anode batteries and 0.4–1.9 V for Zn–Al anode batteries at various scan rates. For the three-electrode configuration, the CV scans were recorded in the range of −0.9 to 0.5 V vs. Hg/Hg2SO4 for TAL and −1.0 to 0.5 V vs. Hg/Hg2SO4 for the Zn–Al system. Unless otherwise specified, all electrochemical measurements were conducted using an electrolyte composed of 0.5 M Al2(SO4)3 and 0.4 M MnSO4.
Open circuit voltage (OCV) conditions were used to perform electrochemical impedance spectroscopy (EIS) measurements, at a state of charge, using a 5 mV voltage amplitude, six data points per decade, and two averages per frequency, in the frequency range of 200 kHz to 0.1 Hz. EC-Lab software (version 11.50) was used to fit equivalent circuits.

2.4. Material Characterization

Atomic absorption spectroscopy (AAS) using a ContrAA 800F spectrometer (Analytik Jena, Germany) was employed to measure Mn and Na contents in the electrolyte after discharge. Measurements were performed at wavelengths of 403.0755 nm (Mn) and 330.237 nm (Na), with atomization using acetylene (55 L/h) and compressed air (75 L/h). For analysis, the battery components were disassembled, and the electrolyte was washed with deionized water to a total volume of 50 mL. To examine the surface topography and roughness of the cathode electrode (δ-MnO2) before and after use in the aluminum battery, an atomic force microscopy (AFM) device (Bruker Co.,USA) was utilized.
A WITec alpha300 apyron Raman microscope (WITec, Germany) equipped with a 532 nm argon-ion laser was used to perform the Raman analysis. A Philips PW3373/00 Cu LFF DK 433781 diffractometer (Philips, Netherlands) operating at 30 mA and 40 kV over a 2θ range of 10–80° was used to perform X-ray diffraction (XRD) analysis of synthesized powders. The XRD grazing incident mode was applied to analyze the electrode surface with an incident angle (θ) fixed at 3, step size 0.02 degrees, and step time 1 s per step using a Siemens D5000 diffractometer (Siemens, Germany). Field emission scanning electron microscopy (FE-SEM) was performed using an FEI ESEM Qanta 200 instrument (FEI, USA). In addition, energy-dispersive X-ray spectroscopy (EDS, Thermo Fisher Scientific, USA) and high-resolution SEM (Hitachi S-4800, Japan) were employed. A field emission transmission electron microscope (TEM, Philips Tecnai-F20, Netherlands) was used to take the TEM image. The surface chemical composition was examined using X-ray photoelectron spectroscopy (XPS, XR50M, SPECS Surface Nano Analysis GmbH, Germany), which used a monochromatic AlKα source running at 300 W and an ultra-high vacuum chamber with a base pressure of 2 × 10−8 mbar. CASAXPS software (version 2.3.22PR1.0) was utilized to perform curve fitting.

3. Results and Discussion

3.1. Synthesized Cathode Materials Characterization

To investigate the effects of compositing δ-MnO2 with graphene, various characterization techniques, XRD, Raman, SEM, TEM, and XPS, were used and will be discussed in detail. The XRD patterns of the synthesized manganese oxide and the graphene-composite material verify the crystalline phase of birnessite (Figure 1a), which is consistent with PDF No. 43-1456. A distinctive diffraction peak at 12.4°, corresponding to the (001) plane with a d-spacing of 7.14 Å along the c-axis, is observed in both samples. Furthermore, a subtle peak at 26.3°, as shown in the inset of Figure 1a, corresponds to the (002) plane of graphene within the δ-MnO2/Graphene composite. The SEM images (Figure 1c) reveal the layered morphology of δ-MnO2 in both samples. The δ-MnO2 nanosheets in the δ-MnO2/Graphene composite are evenly spaced throughout the graphene sheets, suggesting that δ-MnO2 has successfully anchored onto the graphene matrix. The δ-MnO2/Graphene morphology is shown in the TEM images of Figure 1d, which validates the distribution of nanosheet δ-MnO2 over the graphene matrix.
Birnessite crystallite size, as determined by the Debye-Scherrer equation, was 20 nm for δ-MnO2/Graphene and 36 nm for δ-MnO2. The birnessite crystallites in δ-MnO2/Graphene are therefore considerably smaller than those in δ-MnO2.
The Raman spectra of graphene, δ-MnO2, and the δ-MnO2/Graphene composite material are shown in Figure 1b. In both the δ-MnO2 and δ-MnO2/Graphene samples, a stretching vibration of Mn–O is observed at 633 cm−1. This vibration was ascribed to vibrations in the basal plane of MnO6 octahedra within the MnO2 crystalline structure [44]. The D-band, which is linked to structural disorder, and the G-band, which is related to the C–C stretching mode in sp2-hybridized carbon, are the two distinct peaks detected in the graphene’s Raman spectrum at 1337 cm−1 and 1565 cm−1, respectively. For determining the defect density in carbon materials, one commonly used metric is the intensity ratio of the D to G band (ID/IG). When the ID/IG increases from 0.75 in pure graphene to 0.94 in the manganese oxide-graphene composite, it signifies a notable rise in defect density, which may help the interfacial electron transfer between MnO2 and graphene [45], consistent with the improved electrochemical properties of δ-MnO2/Graphene. Furthermore, the D and G bands observed shift to 1343 cm−1 and 1576 cm−1, respectively, suggesting that graphene and MnO2 are strongly interacting and transferring charges [46]. Compared to graphene, the G-band in the composite is weaker and broader, indicating an increase in defects in graphene, possibly due to edge effects or reduction during the hydrothermal synthesis process [47]. This analysis further confirms the successful composite formation and the interaction of graphene sheets with MnO2 without the formation of by-products.
Figure 2 presents the XPS analysis comparing δ-MnO2 and δ-MnO2/Graphene samples. In Figure 2a, the survey XPS spectrum for the δ-MnO2 reveals the presence of Mn, O, C, and Na peaks, confirming the absence of extraneous elements and the purity of the sample. Similarly, the survey spectrum for the δ-MnO2/Graphene composite displays peaks corresponding to Mn 2p, O 1s, C 1s, and Na 1s, with a higher carbon content attributed to the addition of graphene.
Figure 2b–d, which demonstrates high-resolution spectra of individual elements, shows that Mn is present in 2+, 3+, and 4+ oxidation states in the Mn 2p spectrum for both δ-MnO2 and the δ-MnO2/Graphene composite (Figure 2b). The satellite peak (~645.5 eV), which is frequently linked to manganese oxides [48], was isolated from the Mn 2p3/2 peaks in order to quantify the relative amounts of Mn2+, Mn3+, and Mn4+ that are claimed to exist for birnessite in the literature [49,50]. Both δ-MnO2 and δ-MnO2/Graphene have almost the same relative percentage of manganese oxidation states. However, the XPS O 1s spectra (Figure 2c) reveal notable structural changes. The peaks at 530.1, 531.2, and 533.4 eV, corresponding to lattice oxygen (Mn–O–Mn), VO [38,51], and C–O or H–O–H bonds [52], respectively. In the δ-MnO2/Graphene composite, the Mn–O–Mn peak intensity decreases, while the C–O peak intensity and VO increase. The lower intensity of the Mn–O vibration in the Raman spectra for δ-MnO2/Graphene could also be an additional confirmation of an increase in VO [53]. Mn 2p3/2 Mn 2p1/2 exhibits a slight but steady shift in binding energy (~0.1 eV) to lower values, which supports this structural change. These changes have been ascribed to interfacial electronic interactions in composite systems and defect generation, specifically VO, in earlier research [54,55]. It has been observed that VO improves the mobility of charge carriers by altering electronic characteristics, which in turn speeds up the kinetics of ion insertion (EIS-verified; the following sections will address this). Moreover, the presence of VO at the electrolyte-electrode interface influences the surface’s thermodynamic characteristics, which facilitates phase transformations [29,53]. Thus, a contributing factor to the composite’s superior cycling stability is its higher VO content (8.8% to 17.2%). The following sections will illustrate this enhanced performance.
The C–C/C=C, C–O–Mn, C=O, O–C=O, and π-π* transitions are responsible for the peaks at 284.6, 286.1, 288, 289.6, and 291.2 eV in the C1s spectrum for the δ-MnO2/Graphene composite [56,57]. Adventitious carbon is attributed to the peak for δ-MnO2 at 284.6 eV, which is most likely the result of air exposure (Figure 2d). The presence of Mn–O–C bonds in δ-MnO2/graphene indicates the formation of strong interfacial interactions between graphene and δ-MnO2, which can help mitigate the Jahn–Teller effect [57]. This effect, associated with Mn3+, originates from the uneven occupancy of eg orbitals, leading to structural distortions that lower the system energy [28]. During battery cycling, manganese dioxide is prone to partial dissolution into the electrolyte, as confirmed by AAS measurement (discussed in the section that follows). The formation of Mn–O–C bonds could therefore be an efficient way of stabilizing the MnO2 cathode.
It is reported that the interfacial bonding in metal oxide–graphene composites can result in band gap narrowing. Ti–O–C bond formation, for instance, has been demonstrated to modify the band structure and to considerably narrow the band gap in a TiO2-graphene system [58]. Building on our earlier findings [42], the band gap decreased from 1.7 eV in pristine MnO2 to 1.4 eV in MnO2/Graphene. The observed band gap reduction is probably caused by similar effects induced by the formation of Mn–O–C bonds between MnO2 and graphene. Prior research has also documented that in MnO2-graphene composites, this type of interfacial bonding can directly affect the band structure and reduce the energy needed for electronic transitions [59]. On the other hand, the MnO2 nanostructures are securely anchored to the conductive graphene sheets by Mn–O–C bonds that form at the MnO2/graphene interface [60]. therefore, while reducing mechanical strain during repeated cycling, these strong interfacial bonds enable effective electron transport between MnO2 and graphene. So, the formation of Mn–O–C bonds, as another stabilizing effect of graphene, could enhance the structural stability of the composite and improve its cycling performance.

3.2. Study of the Battery System Utilizing TAl Anode

To investigate the effect of surface modification of Al anode on battery performance, batteries were assembled using both untreated Al and TAl anodes. As shown in Figure S1, the surface treatment of the aluminum anode enhances electrochemical performance in the TAl|δ-MnO2 system, increasing the capacity from 255 mAh g−1 to 890 mAh g−1 and reducing polarization. The anode potential also shifted from −1.0 V to −1.1 V vs. Hg/Hg2SO4 reference electrode, while the OCV of the fresh cell (measured as the initial potential between working and counter electrode) increased by 0.15 V from 1.2 V to 1.35 V. These improvements stem from changes in anode surface chemistry that facilitate charge transfer. This supports earlier findings that an artificial solid electrolyte interphase (ASEI) layer can improve interfacial charge transfer and inhibit additional oxide formation by corroding the native Al2O3 film [61]. Although the cathode remained unchanged, the higher capacity with the TAl anode suggests that the untreated Al surface previously limited MnO2 utilization due to poor interfacial kinetics. Thus, the anode plays a critical role in determining the practical capacity of the cell, even when the cathode is nominally the limiting component.
AAS analysis of the electrolyte after the first discharge of the δ-MnO2 cathode in 0.5 M Al2(SO4)3 electrolyte revealed the dissolution of manganese (1.16 mg/L) and sodium (4.48 mg/L), which can degrade battery performance. To mitigate this, MnSO4 was added to the electrolyte to stabilize MnO2 and preserve the cathode material. Figure S2 compares the cycling performance of cells using δ-MnO2 and δ-MnO2/Graphene cathodes with TAl anode under different MnSO4 concentrations. For the δ-MnO2 cathode, the addition of 0.2 M MnSO4 resulted in an increase in capacity compared to the MnSO4-free electrolyte, while 0.4 M MnSO4 led to even higher capacity over extended cycling. A similar trend was observed for the δ-MnO2/Graphene cathode, where the 0.4 M MnSO4 electrolyte consistently outperformed the electrolyte without MnSO4.
Furthermore, charge–discharge profiles (Figure S1) demonstrate that the presence of 0.4 M MnSO4 reduces polarization and increases the discharge capacity, supporting previous findings [6] that Mn2+ ions contribute to Mn-oxide-based cathode stabilization in AAIBs. In addition, Mn2+ may also participate in redox processes, providing an auxiliary charge storage mechanism beyond Al3+ intercalation, as reported in the literature [25].
Based on these findings, the electrolyte composition of 0.5 M Al2(SO4)3 and 0.4 M MnSO4 was selected as the optimized composition for subsequent electrochemical evaluations.
The XRD results for the δ-MnO2 cathode (Figure 3a) indicate that the distinct birnessite (001) diffraction peak at 12.4° disappears completely in the discharged state and does not reappear upon charging. This suggests that δ-MnO2 undergoes a structural transformation during the first discharge. It has been reported that during this process, a new amorphous manganese oxide compound containing aluminum (AlₓMn(1−x)O2) forms [6]. Similarly, XRD analysis of the δ-MnO2/Graphene cathode after the second discharge (Figure 3a) shows the absence of the birnessite peak, reinforcing the conclusion that structural change of MnO2 occurs in both types of cathodes during the first cycle.
Charge-discharge curves (Figures S1 and S3) and CV analysis (Figure S4) for aluminum batteries using δ-MnO2 and δ-MnO2/Graphene cathodes further support these findings. The significant differences between the first and second cycles indicate that the aluminum is irreversibly incorporated into the cathode, thus changing the active substance. This implies again that the manganese dioxide in both cathodes undergoes an irreversible structural change during the first cycle.
To further investigate morphological changes, SEM analysis was performed on δ-MnO2 and δ-MnO2/Graphene electrodes before battery assembly and after the second discharge (Figure 3b). The fresh electrodes exhibit a well-defined layered morphology of manganese dioxide. However, after the second discharge, the layered morphology for the δ-MnO2 cathode appears less coherent, and a noticeable surface change suggests structural rearrangements that could impact electrode performance. AFM analysis (Figure 3c) reveals an increase in surface roughness (Ra) after the second discharge for the δ-MnO2 cathode, likely due to the formation of the amorphous AlₓMn(1−x)O2 phase and other cycling-induced structural changes. These results align with previous studies that associate increased roughness with structural modifications and void formation caused by repeated charge-discharge cycles [62].
Despite these transformations, SEM images in Figure 3b confirm that the overall layered morphology remains preserved after the second discharge, as also observed in the AFM data (Figure 3c). Notably, Figure 3b shows that in δ-MnO2/Graphene electrodes, graphene plays a crucial role in maintaining cathode integrity. Although the newly formed cathode material became XRD-amorphous, graphene helped stabilize the manganese oxide layers, likely due to Mn–O–C interactions at the δ-MnO2/graphene interface. This stabilizing effect, as previously discussed in the XPS section, contributes to improved cathode integrity during cycling. Consequently, the incorporation of graphene enhanced stability, electrochemical performance, and energy storage capacity.
EDS analysis provided additional insight into compositional changes in the cathode. Figure 4a verifies that the δ-MnO2/Graphene electrode contains Mn, Na, and O prior to use, consistent with the Na-birnessite elements. The EDS mapping results further evidence a uniform elemental distribution. Signals from Fe, Ni, Cr, Mo, and Si correspond to the stainless-steel substrate. After the first charge, EDS analysis of the δ-MnO2/Graphene cathode (Figure 4b) reveals the appearance of an aluminum peak, indicating that during the initial discharge, aluminum from the electrolyte is integrated into the cathode material. Notably, aluminum remains in the cathode after charging, supporting the irreversible formation of AlₓMn(1−x)O2. The absence of a sodium peak suggests a structural change in the birnessite structure, consistent with the XRD findings. The uniform distribution of the elements Al, O, C, and Mn is further revealed by the EDS map.
Figure 5a displays the second cycle charge-discharge curves for δ-MnO2 and δ-MnO2/Graphene electrodes at a current density of 180 mA g−1. The battery incorporating the δ-MnO2/Graphene cathode exhibits lower voltage polarization and an elevated discharge voltage compared to the δ-MnO2 cathode. This enhancement is ascribed to improved reaction kinetics made possible by graphene [63], as confirmed by the EIS analysis (discussed later). The discharge curves reveal two distinct voltage plateaus for the δ-MnO2 electrode at approximately 1.1 V and 0.5 V. In contrast, the δ-MnO2/Graphene electrode exhibits higher discharge voltage plateaus at 1.3 V and 0.7 V, indicating improved energy efficiency and reduced internal resistance. With a specific capacity of 1017 mAh g−1, the δ-MnO2/Graphene electrode outperforms the pure δ-MnO2 electrode by 171 mAh g−1 (890 mAh g−1). This enhancement highlights the role of graphene in forming a conductive network that mitigates polarization and improves electrochemical performance [64].
The cycling stability and coulombic efficiency of δ-MnO2 and δ-MnO2/Graphene electrodes are presented in Figure 5b,c. Following 45 cycles at 180 mA g−1, δ-MnO2/Graphene’s discharge capacity dropped from 934 mAh g−1 to 739 mAh g−1. At the same time, δ-MnO2 underwent a significant degradation, going from 705 mAh g−1 to 259 mAh g−1. The greater coulombic efficiency of δ-MnO2/Graphene compared to δ-MnO2 is shown in Figure 5c. The rate capability test at various charge/discharge rates (Figure 5d) confirms superior rate performance for δ-MnO2/Graphene compared to δ-MnO2. After cycling at 3.6 A g−1, the δ-MnO2/Graphene composite exhibits a 64% capacity retention, which is significantly higher than that of pristine δ-MnO2, which retains only 47%.
The Nyquist plots in Figure 5e, along with the fitted EIS parameters in Table 1, highlight the distinct electrochemical impedance characteristics of δ-MnO2 and δ-MnO2/Graphene in full cells. According to the corresponding equivalent circuit, R1 stands for the electrolyte resistance and R2 for the SEI layer resistance. Q1 and Q2 are constant phase elements (CPEs) that account for non-ideal capacitive behavior, incorporating the effects of electrode surface roughness. R3 refers to the charge transfer resistance (Rct) [65].
The lower ohmic resistance (R1 = 1.78   Ω ) of δ-MnO2/Graphene compared to δ-MnO2 (R1 = 2.6 Ω ) indicates that graphene enhances electronic conductivity within the electrode, reduces resistive losses at the electrode-electrolyte interface, and improves overall electron transport pathways. Furthermore, the charge transfer resistance for δ-MnO2/Graphene (R3 = 1110   Ω ) is considerably lower than that observed for δ-MnO2 (R3 = 1333 Ω ), demonstrating improved electrochemical reaction kinetics [41]. Another notable feature, presented in Table 1, is the significantly higher Q2 value of the δ-MnO2/Graphene electrode. Given the similar CPE exponent (n ≈ 0.45) in both cases, this can be interpreted as an improvement in pseudocapacitive behavior. This improvement may stem from the increased surface area and the greater number of electroactive MnO2 sites uniformly dispersed on the graphene framework. The presence of graphene effectively minimizes MnO2 agglomeration, promoting better exposure of active material and facilitating more efficient ion and electron transport. These effects lead to higher charge storage capacity and improved electrochemical kinetics, which contribute to the superior rate performance and overall energy storage capability of the δ-MnO2/Graphene composite. This trend is further supported by Figure S5, where the δ-MnO2/Graphene electrode exhibits a noticeably larger enclosed area in the CV curves compared to δ-MnO2 alone.
The stabilizing effect of graphene, due to the increase in VO and Mn–O–C bonds formation as discussed in the XPS section, further enhances the electrochemical performance of the δ-MnO2/Graphene composite cathode. In the case of AAIBs, Gu et al. [22] demonstrated that introducing VO into γ-MnO2 significantly accelerated ion intercalation by expanding the MnO6 octahedral framework, reducing diffusion barriers, and increasing reaction kinetics. This modification led to a lowered band gap of MnO2 and a higher discharge capacity. Similarly, Han et al. [53] examined β-MnO2 in Zn-ion batteries and, based on density functional theory (DFT), showed that oxygen vacancies increase ion-transport channels and enhance electronic conductivity, thereby yielding higher specific capacity and improved cycling performance. In our case, the reduced band gap and enhanced electrochemical performance of the composite are therefore likely due to the dual effect of VO and Mn–O–C bond formation.

3.3. Study of the Battery System Utilizing Zn–Al Anode

The results in the previous section indicated that the TAl anode in aqueous electrolyte exhibits limited cycling stability, enduring only 45 charge/discharge cycles. This instability is mainly due to the progressive degradation of the aluminum surface, where the initially beneficial ASEI gradually loses effectiveness due to passivation. In aqueous electrolytes, these ASEIs tend to degrade during cycling, resulting in unstable byproducts [66]. Despite the improvement in initial cycle performance, it has been reported that the ASEI does not support the electrochemical reduction of Al3+ to metallic aluminum, but instead the development of corrosion and hydrogen gas contributes to secondary reactions [61]. As shown in Figure S7a, the TAl anodes extracted from the batteries after cycling exhibit aluminum dissolution over time, leading to an irreversible loss of active material due to self-corrosion. In the SEM image shown in Figure S7b, the EDS results reveal the formation of a non-conductive surface layer containing aluminum, oxygen, and sulfur on the aluminum surface. Additionally, the SEM image in Figure S7c clearly displays pits and pores on the Al surface, indicating aluminum dissolution due to self-corrosion and the potential occurrence of gas evolution. These findings are in line with previous literature reports: under highly acidic (pH < 4) or alkaline (pH > 8.6) conditions, aluminum corrodes to Al3+ or AlO2 with a subsequent development of hydrogen gas until either the metal or water is exhausted, whereas under medium pH conditions, an oxide film may be formed on the surface [67]. Thus, the low cycling stability may be attributed to anode passivation, anode dissolution, or depletion of the electrolyte.
To mitigate these issues, alloying aluminum with other metals has been explored as a strategy to suppress self-corrosion, prevent hydrogen evolution reactions, and inhibit the formation of inactive oxide layers [68]. Here, zinc has been chosen as a substrate for the formation of the Zn-Al anode material by co-deposition of Zn2+ and Al3+ on the Zn. The Zn–Al|δ-MnO2 system exhibited an initial OCV of 1.65 V, with the anode potential at −1.4 V vs. Hg/Hg2SO4. In contrast, the TAl|δ-MnO2 system showed an initial OCV of 1.35 V. XRD analysis (Figure 6a) was conducted on both bare zinc and the anode after the third charge in a battery with δ-MnO2/Graphene cathode. The results indicate that all detected peaks correspond to metallic zinc (PDF-40-0831), with no evidence of metallic aluminum or aluminum oxide formation. The significant variation in peak intensity ratios, particularly I(002)/I(100), provides further evidence for the occurrence of aluminum alloying with zinc [8]. These two peaks also exhibit a subtle shift of approximately 0.02° toward lower angles, while the positions of the other peaks remain unchanged.
SEM and EDS analyses were used to examine the surface morphology and elemental distribution of the anode (Figure 6b–d). Compared to the initially smooth morphology of the zinc substrate (Figure 6a), the roughened surface observed after the third charge in the SEM image (Figure 6c) indicates the formation of a coating layer on the zinc after battery operation. EDS analysis (Figure 6d) confirms a uniform distribution of aluminum, with an atomic concentration of 2% in the alloy. The co-existence of zinc and aluminum in the formed layer further validates the formation of the Zn-Al alloy.
Figure 6e–g present the SEM image and EDS analysis of the δ-MnO2/Graphene cathode after the third charge in a battery with a Zn-Al alloy anode. Zinc signals are absent from the EDS results, indicating that Zn2+ ions are not involved in the cathodic discharge process, hence offering no contribution to the battery’s capacity. However, Zn2+ is involved in the stripping/plating process at the anode and plays a role in forming the alloy with aluminum, as also reported in previous studies [8,69]. The appearance of aluminum peaks in the EDS indicates the incorporation of aluminum into the cathode structure, resembling behavior observed in batteries with an aluminum anode. The uniform distribution of Al, O, and Mn elements and the absence of sodium peaks strongly imply that birnessite has undergone a structural change and the formation of a new phase, most likely AlₓMn(1−x)O2.
In a battery consisting of a Zn-Al anode, Figure 7a compares the charge/discharge profiles of δ-MnO2 and δ-MnO2/Graphene cathodes. The δ-MnO2/Graphene cathode demonstrates a higher discharge voltage and lower polarization compared to δ-MnO2. Furthermore, the capacity of the second cycle for δ-MnO2/Graphene is 584 mAh g−1, which is 248 mAh g−1 more than that of δ-MnO2.
The coulombic efficiency and cycling performance of δ-MnO2 and δ-MnO2/Graphene electrodes are shown in Figure 7c,b. As shown in Figure 7b, δ-MnO2/Graphene retains a capacity of 746 mAh g−1 after 352 cycles at 180 mAh g−1, while δ-MnO2 maintains a capacity of 642 mAh g−1 after 107 cycles. Both δ-MnO2 and δ-MnO2/Graphene electrodes exhibit a gradual increase in capacity over the initial cycles (Figure 7a,b), indicating an activation process. This behavior is probably due to the progressive formation of an AlₓMnO2 phase, as reported by Pan et al. [23], which enhances Al3+ storage capability. Additionally, Ejigu et al. [1] attributed similar capacity increases to gradual surface activation via electrodeposition of Mn species from Mn2+ additive. Based on charge/discharge profile evolution over cycling, they proposed enhanced intercalation/extraction kinetics of H+ and Al3+ ions, as well as improved Al plating/stripping behavior at the anode. Additionally, Figure 7c demonstrates that, similar to the capacity, the coulombic efficiency of both electrodes improves over cycles, approaching nearly 100% in later stages. These findings align with our observations. Supporting this, the Nyquist plot also shows the emergence of a Warburg diffusion tail after 100 cycles, further indicating enhanced ion transport and activation of the cathode material over time, as will be discussed later in this section.
The rate capability test at various charge/discharge rates (Figure 7d) highlights the superior rate capability of δ-MnO2/Graphene over δ-MnO2, with higher capacity retention even after cycling under high-rate conditions up to 1.8 A g−1. At this current, δ-MnO2/Graphene delivers 444 mAh g−1, whereas δ-MnO2 achieves 334 mAh g−1. Furthermore, after prolonged cycling at 1.8 A g−1, δ-MnO2/Graphene retains 95% of its initial capacity, compared to 90% for δ-MnO2.
As illustrated in Figure 8a, the discharge profiles reveal that the battery with a δ-MnO2/Graphene cathode and a Zn-Al anode displays a discharge plateau around 1.63 V, while the same cathode paired with a TAl anode exhibited a plateau at 1.3 V. During charging, Zn2+ (produced by oxidation of Zn at the anode during discharge) and Al3+ ions co-deposit onto the zinc substrate, forming a Zn-Al alloy that causes a higher discharge voltage plateau and reduced polarization in the Zn-Al system. This alloy formation helps suppress the formation of an insulating Al2O3 layer and mitigates aluminum self-corrosion, both of which would otherwise contribute to increased overpotential, particularly due to the competing hydrogen evolution reaction [3].
Comparing the Zn-Al system to the battery with a TAl anode, the latter achieves a higher capacity (Figure 8a). In the Zn-Al system, zinc lowers the total amount of aluminum that is electrochemically active. While zinc forms a large part of the alloy, it does not contribute to the battery’s capacity. The lower aluminum proportion in the alloy results in fewer active aluminum atoms available for charge storage compared to the pure aluminum system, where 100% of the anode material consists of aluminum, leading to a higher total capacity. Due to the solubility limit of aluminum sulfate in water (a maximum of 0.5 M; exceeding this concentration results in precipitates), the initial concentration of Al3+ in the electrolyte is 1 M. However, the effective concentration of available Al3+ during cycling is lower, as a portion is consumed in alloy formation with Zn at the anode. Consequently, the remaining amount of Al3+ in the electrolyte is reduced, which may also contribute to the decrease in capacity. Moreover, certain side reactions, such as hydrogen evolution or self-corrosion of aluminum, may contribute to the total measured capacity, especially in early cycles.
Three-electrode CVs were conducted (Figures S6 and S8), confirming that the electrolyte remains stable within the operating voltage window of the TAl and Zn–Al batteries. However, the CV of the electrolyte from −1.0 to +0.6 V vs. Hg/Hg2SO4 in Figure S8 shows an anodic redox process emerging above ~+0.5 V that is attributable to the oxidation of Mn2+ to MnOx/MnO2 on the cathode during charge [70]. This feature is absent when the sweep is limited to +0.5 V. Using the potential–cell-voltage mapping in Figures S6 and S8, the onset corresponds to ~1.5 V in the TAl cell and ~1.9 V in the Zn–Al cell; thus, in the presence of Mn2+ (from MnSO4), anodic MnO2 electrodeposition is expected only when the positive electrode is driven beyond these voltages. Consistent with this, in Figure 8b, the TAl|δ-MnO2 cell reaches ~1.6 V during charging, where MnO2 deposition on the δ-MnO2 electrode may occur, whereas the Zn–Al|δ-MnO2 cell remains below ~1.9 V, making such deposition unlikely.
Although MnO2 deposition is unlikely to occur in our Zn–Al system, Mn2+ may still contribute to charge storage alongside Al3+. Upon discharge, both cations may form the phase AlxMn1−xO2, as proposed by He et al. [6], who observed a conversion/dissolution–deposition process involving Mn2+ and AlₓMnO2. Notably, their system was charged beyond 1.5 V in an Al-anode configuration, which could allow MnO2 deposition. Additionally, in our system, no evidence of new MnO2-related morphology was observed on the cathode, and the layered morphology of the birnessite was retained after cycling. Therefore, although we do not claim MnO2 deposition as an active mechanism in our system, we also cannot rule it out, as additional experiments would be required for confirmation.
Furthermore, CV data (Figure 7e,f) show two prominent redox peaks at ~1.6 V and 1.9 V for δ-MnO2 and δ-MnO2/Graphene coupled with a Zn-Al anode, closely matching voltages reported by Yan et al. [8] and Meng et al. [69]. for reversible Al3+ intercalation/deintercalation into AlₓMnO2. Notably, the first CV cycle differs significantly from subsequent ones, confirming an initial irreversible conversion reaction. The close overlapping of the second and third cycles indicates a stabilized and reversible redox process. These features are consistently observed in both δ-MnO2 and δ-MnO2/Graphene electrodes. Taken together, our results are consistent with literature reports, suggesting that the dominant charge storage mechanism in our system involves an initial irreversible MnO2 conversion, followed by reversible Al3+ intercalation/deintercalation into amorphous AlₓMn(1−x)O2. However, it should be noted that the detailed charge storage mechanism is not completely elucidated and requires additional merits.
As shown earlier in Figure 5 and Figure 7, the Zn–Al anode-based battery demonstrated superior cycling stability and higher coulombic efficiency compared to the TAl anode system, likely due to the suppression of side reactions such as hydrogen evolution. The introduction of additional metals to interact with aluminum is reported to alter its electrochemical behavior, potentially enabling underpotential deposition and promoting reversible metal deposition and dissolution at the anode, as suggested in previous studies [8,71].
In AAIBs, the co-deposition of Al and Zn involves the simultaneous reduction of Al3+ and Zn2+ ions onto a substrate, forming a Zn–Al alloy layer that enhances electrochemical performance. The deposition process begins with preferential Zn deposition due to its higher reduction potential relative to Al. The deposited Zn then facilitates Al nucleation by providing a conductive and chemically compatible substrate, lowering the energy barrier for Al deposition. Alloying between Zn and Al modifies the electrochemical environment at the electrode surface, mitigating passivation effects from aluminum oxide/hydroxide layers and suppressing the competing hydrogen evolution reaction [1,66]. Figure 9 shows a schematic representation of the suggested electrochemical reactions taking place in the δ-MnO2/Graphene|Zn–Al cell.
Figure 8. (a) Charge/discharge profiles of the second cycle at 180 mA g−1 for full cell batteries δ-MnO2/Graphene cathodes using Zn-Al and TAl anode. (b) Potential of the working electrode versus time during charge/discharge for δ-MnO2 cathodes with TAl and Zn–Al anodes. MnO2 deposition on the anode may occur in the δ-MnO2|TAl cell due to charging overpotentials reaching ~1.6 V, while it does not occur in the δ-MnO2|Zn–Al system, where voltages remain below 1.9 V throughout cycling. (c) Comparing the electrochemical properties (capacity and voltage) of this study with those of recently reported AAIBs [1,6,8,9,10,11,12,22,23,72,73].
Figure 8. (a) Charge/discharge profiles of the second cycle at 180 mA g−1 for full cell batteries δ-MnO2/Graphene cathodes using Zn-Al and TAl anode. (b) Potential of the working electrode versus time during charge/discharge for δ-MnO2 cathodes with TAl and Zn–Al anodes. MnO2 deposition on the anode may occur in the δ-MnO2|TAl cell due to charging overpotentials reaching ~1.6 V, while it does not occur in the δ-MnO2|Zn–Al system, where voltages remain below 1.9 V throughout cycling. (c) Comparing the electrochemical properties (capacity and voltage) of this study with those of recently reported AAIBs [1,6,8,9,10,11,12,22,23,72,73].
Processes 13 03551 g008
Figure 9. Schematic representation of the electrochemical reactions occurring in the Zn–Al|0.5 M Al2(SO4)3, 0.4 M MnSO4|δ-MnO2/Graphene system after the first cycle.
Figure 9. Schematic representation of the electrochemical reactions occurring in the Zn–Al|0.5 M Al2(SO4)3, 0.4 M MnSO4|δ-MnO2/Graphene system after the first cycle.
Processes 13 03551 g009
Compared to reported AAIBs, the obtained specific capacity (746 mAh g−1) ranks among the highest values achieved to date. Table 2 and Figure 8c summarize the electrochemical properties of AAIBs batteries with oxide-based cathodes from recent literature, benchmarking them against the results of this study.
The Nyquist plots in Figure 10a,b, along with the fitting parameter values in Table 3, reveal the electrochemical performance differences between δ-MnO2 and its graphene composite (δ-MnO2/Graphene) in a Zn-Al anode battery, demonstrating the enhanced electrochemical properties of the δ-MnO2/Graphene composite. The same equivalent circuit used for the TAl system (Figure 5e) was applied to fit the Nyquist plots of the Zn-Al system after three cycles and after 100 cycles, with the difference after 100 cycles being the appearance of the Warburg element (W), which represents the impedance associated with solid-state ion diffusion into the bulk material. Similar to the impedance results observed in the battery with a TAl anode, the battery with a δ-MnO2/Graphene cathode exhibits lower R values (Table 3). The δ-MnO2/Graphene composite electrode demonstrates lower R values compared to δ-MnO2 in a battery with a Zn-Al anode, even after 100 cycles, which reflects the composite stability over cycling. Additionally, the decrease in R3 for both materials over 100 cycles, compared to R3 after three cycles, suggests enhanced charge transfer, likely due to electrode surface activation during cycling.
The equation below can be used to calculate the bulk ionic diffusion coefficient (Dion) [74,75]:
D i o n =   R 2 T 2 2 n 4 A 2 C i o n 2 σ 2
where R stands for the gas constant (8.314 J mol−1K−1), T for temperature (K), n for number of electrons transferred, A for electrode area (cm2), and Cion for ion concentration (mol cm−3), and σ (Ω s−1/2) refers to the Warburg coefficient extracted from the slope of the Z′ vs. ω−1/2 plot (Figure 10c). Based on this equation, the lower σ observed for δ-MnO2/Graphene indicates a higher Dion, suggesting faster ion diffusion and better electrochemical performance. This can be attributed to the smaller MnO2 nanoparticles (as determined by Scherrer analysis), which are anchored on a conductive graphene network, thereby reducing diffusion distances.
Comparison of the resistance values obtained from the Nyquist plots for batteries with Zn-Al and TAl anodes (Table 3 and Table 1, respectively) shows that the TAl system exhibits lower ohmic resistance (R1) and charge transfer resistance (R3). However, the Zn-Al alloy likely mitigates side reactions, such as hydrogen evolution and aluminum oxide formation, which are more pronounced in pure aluminum anodes. As a result, despite its higher charge transfer resistance, the Zn-Al system undergoes less voltage drop during discharge.
In contrast, while the TAl anode may initially offer more favorable charge transfer properties, it is more prone to degradation over time due to parasitic reactions. These effects reduce voltage stability and overall efficiency. Thus, although the Zn-Al system may have slower initial kinetics, it provides better energy retention, leading to a higher and more stable discharge plateau.

4. Conclusions

The incorporation of graphene into δ-MnO2 significantly enhanced the electrochemical performance of AAIBs. XPS analysis confirmed the presence of oxygen vacancies (VO) and Mn–O–C bonding in the δ-MnO2/Graphene composite, both of which improve electronic conductivity and stabilize the MnO2 during cycling. Notably, Mn–O–C bonding is known to mitigate Jahn–Teller distortions in Mn-based oxides, thereby preserving the integrity of the cathode material under repeated charge–discharge conditions, even though the δ-MnO2 in the composite became XRD-amorphous during cycling.
In addition, defect-induced structural and electronic modifications, evidenced by the increased ID/IG ratio in the Raman spectra, further enhanced interfacial charge transport between MnO2 and the graphene matrix. The EIS data also showed that the δ-MnO2/Graphene electrode has reduced ohmic and charge transfer resistances. Furthermore, a lower Warburg coefficient (σ) is correlated with a higher ionic diffusion coefficient (Dion), which indicates an improvement in ionic transport. Shorter diffusion pathways, due to smaller particle crystallite size, reduced particle agglomeration, and the open structure provided by graphene, further contributed to this improvement.
On the anode side, although TAl-based full cells initially exhibited higher capacities and lower charge transfer resistance, the introduction of Zn as a substrate for in situ Zn and Al co-deposition effectively extended cycle life and increased the discharge voltage to 1.63 V, compared to 1.3 V for the TAl system. The δ-MnO2/Graphene|0.5 M Al2(SO4)3, 0.4 M MnSO4|Zn–Al system delivered an outstanding capacity of 746 mAh g−1 over 352 cycles, while δ-MnO2 alone retained a value of 642 mAh g−1 for 107 cycles.
To the best of our knowledge, this study is the first to demonstrate a MnO2–graphene cathode in AAIBs, highlighting the potential of graphene-based cathode composites for developing high-performance, cost-effective AAIBs, offering a promising pathway toward next-generation sustainable energy storage technologies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pr13113551/s1.

Author Contributions

A.A.: Operated the work, Conceptualization, Software, Investigation, Writing—Original Draft. R.S.-M.: Supervision, Writing—Review and Editing. M.S.: Supervision, Writing—Review and Editing. C.B.: Data Analysis, Writing—Review and Editing. M.K.: Data Analysis, Writing—Review and Editing. A.B.: Supervision, Writing—Review and Editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All data supporting the findings of this study are provided in the article and Supplementary Information. Additional details are available from the corresponding author upon request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ejigu, A.; Le Fevre, L.W.; Elgendy, A.; Spencer, B.F.; Bawn, C.; Dryfe, R.A.W. Optimization of Electrolytes for High-Performance Aqueous Aluminum-Ion Batteries. ACS Appl. Mater. Interfaces 2022, 14, 25232–25245. [Google Scholar] [CrossRef]
  2. Tu, J.; Song, W.L.; Lei, H.; Yu, Z.; Chen, L.L.; Wang, M.; Jiao, S. Nonaqueous Rechargeable Aluminum Batteries: Progresses, Challenges, and Perspectives. Chem. Rev. 2021, 121, 4903–4961. [Google Scholar] [CrossRef]
  3. Kim, J.; Kim, M.; Selvamani, T.; Tak, Y.; Lee, G. Multi-Ionic Capacity of Zn-Al/V6O13 Systems Enable Fast-Charging and Ultra-Stable Aqueous Aluminium-Ion Batteries. ChemElectroChem 2022, 9, e202200964. [Google Scholar] [CrossRef]
  4. Wang, B.; Tang, Y.; Deng, T.; Zhu, J.; Sun, B.; Su, Y.; Ti, R.; Yang, J.; Wu, W.; Cheng, N.; et al. Recent progress in aqueous aluminum-ion batteries. Nanotechnology 2024, 35, 362004. [Google Scholar] [CrossRef]
  5. Zhao, Q.; Zachman, M.J.; Al Sadat, W.I.; Zheng, J.; Kourkoutis, L.F.; Archer, L. Solid electrolyte interphases for high-energy aqueous aluminum electrochemical cells. Sci. Adv. 2018, 4, eaau8131. [Google Scholar] [CrossRef] [PubMed]
  6. He, S.; Wang, J.; Zhang, X.; Chen, J.; Wang, Z.; Yang, T.; Liu, Z.; Liang, Y.; Wang, B.; Liu, S.; et al. A High-Energy Aqueous Aluminum-Manganese Battery. Adv. Funct. Mater. 2019, 29, 1905228. [Google Scholar] [CrossRef]
  7. Lv, H.; Yang, S.; Li, C.; Han, C.; Tang, Y.; Li, X.; Wang, W.; Li, H.; Zhi, C. Suppressing passivation layer of Al anode in aqueous electrolytes by complexation of H2PO4− to Al3⁺ and an electrochromic Al ion battery. Energy Storage Mater. 2021, 36, 350–357. [Google Scholar] [CrossRef]
  8. Yan, C.; Lv, C.; Wang, L.; Cui, W.; Zhang, L.; Dinh, K.N.; Tan, H.; Wu, C.; Wu, T.; Ren, Y.; et al. Architecting a Stable High-Energy Aqueous Al-Ion Battery. J. Am. Chem. Soc. 2020, 142, 15295–15304. [Google Scholar] [CrossRef]
  9. Yan, C.; Lv, C.; Jia, B.E.; Zhong, L.; Cao, X.; Guo, X.; Liu, H.; Xu, W.; Liu, D.; Yang, L.; et al. Reversible Al Metal Anodes Enabled by Amorphization for Aqueous Aluminum Batteries. J. Am. Chem. Soc. 2022, 144, 11444–11455. [Google Scholar] [CrossRef]
  10. Ran, Q.; Zeng, S.P.; Zhu, M.H.; Wan, W.B.; Meng, H.; Shi, H.; Wen, Z.; Lang, X.Y.; Jiang, Q. Uniformly MXene-Grafted Eutectic Aluminum-Cerium Alloys as Flexible and Reversible Anode Materials for Rechargeable Aluminum-Ion Battery. Adv. Funct. Mater. 2023, 33, 2211271. [Google Scholar] [CrossRef]
  11. Chen, S.; Kong, Y.; Tang, C.; Gadelhak, N.A.; Nanjundan, A.K.; Du, A.; Yu, C.; Huang, X. Doping Regulation Stabilizing δ-MnO2 Cathode for High-Performance Aqueous Aluminium-ion Batteries. Small 2024, 20, 2312229. [Google Scholar] [CrossRef] [PubMed]
  12. Ran, Q.; Shi, H.; Meng, H.; Zeng, S.P.; Wan, W.B.; Zhang, W.; Wen, Z.; Lang, X.Y.; Jiang, Q. Aluminum-copper alloy anode materials for high-energy aqueous aluminum batteries. Nat. Commun. 2022, 13, 576. [Google Scholar] [CrossRef] [PubMed]
  13. González, J.R.; Nacimiento, F.; Cabello, M.; Alcántara, R.; Lavela, P.; Tirado, J.L. Reversible intercalation of aluminium into vanadium pentoxide xerogel for aqueous rechargeable batteries. RSC Adv. 2016, 6, 62157–62164. [Google Scholar] [CrossRef]
  14. Zhao, Q.; Liu, L.; Yin, J.; Zheng, J.; Zhang, D.; Chen, J.U.; Archer, L.A.; Zhao, Q.; Yin, J.; Frederick, L.A.A.R.; et al. Proton Intercalation/De-Intercalation Dynamics in Vanadium Oxides for Aqueous Aluminum Electrochemical Cells. Angew. Chem. Int. Ed. 2020, 59, 3048–3052. [Google Scholar] [CrossRef]
  15. Kumar, S.; Satish, R.; Verma, V.; Ren, H.; Kidkhunthod, P.; Manalastas, W.; Srinivasan, M. Investigating FeVO4 as a cathode material for aqueous aluminum-ion battery. J. Power Sources 2019, 426, 151–161. [Google Scholar] [CrossRef]
  16. Pang, Q.; Yang, S.; Yu, X.; He, W.; Zhang, S.; Tian, Y.; Xing, M.; Fu, Y.; Luo, X. Realizing reversible storage of trivalent aluminum ions using VOPO4·2H2O nanosheets as cathode material in aqueous aluminum metal batteries. J. Alloys Compd. 2021, 885, 161008. [Google Scholar] [CrossRef]
  17. Nandi, S.; Das, S.K. An electrochemical study on bismuth oxide (Bi2O3) as an electrode material for rechargeable aqueous aluminum-ion battery. Solid. State Ion. 2020, 347, 115228. [Google Scholar] [CrossRef]
  18. Lahan, H.; Das, S.K. Reversible Al3+ ion insertion into tungsten trioxide (WO3) for aqueous aluminum-ion batteries. Dalton Trans. 2019, 48, 6337–6340. [Google Scholar] [CrossRef]
  19. Joseph, J.; Nerkar, J.; Tang, C.; Du, A.; O’Mullane, A.P.; Ostrikov, K.K. Reversible Intercalation of Multivalent Al3+ Ions into Potassium-Rich Cryptomelane Nanowires for Aqueous Rechargeable Al-Ion Batteries. ChemSusChem 2019, 12, 3753–3760. [Google Scholar] [CrossRef]
  20. Wu, C.; Gu, S.; Zhang, Q.; Bai, Y.; Li, M.; Yuan, Y.; Wang, H.; Liu, X.; Yuan, Y.; Zhu, N.; et al. Electrochemically activated spinel manganese oxide for rechargeable aqueous aluminum battery. Nat. Commun. 2019, 10, 73. [Google Scholar] [CrossRef]
  21. Melzack, N.; Wills, R.G.A. A Review of Energy Storage Mechanisms in Aqueous Aluminium Technology. Front. Chem. Eng. 2022, 4, 778265. [Google Scholar] [CrossRef]
  22. Gu, H.; Yang, X.; Chen, S.; Zhang, W.; Yang, H.Y.; Li, Z. Oxygen Vacancies Boosted Proton Intercalation Kinetics for Aqueous Aluminum-Manganese Batteries. Nano Lett. 2023, 23, 11842–11849. [Google Scholar] [CrossRef] [PubMed]
  23. Pan, W.; Mao, J.; Wang, Y.; Zhao, X.; Leong, K.W.; Luo, S.; Chen, Y.; Leung, D.Y.C. High-Performance MnO2/Al Battery with In Situ Electrochemically Reformed AlxMnO2 Nanosphere Cathode. Small Methods 2021, 5, 2100491. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, G.; Zhu, J.; Lin, L.; Liu, Y.; Li, S.; Li, Q.; Liu, X.X.; Sun, X. A polydopamine coating enabling the stable cycling of MnO2 cathode materials in aqueous zinc batteries. Chem. Sci. 2024, 15, 3545–3551. [Google Scholar] [CrossRef] [PubMed]
  25. Yang, X.; Jia, Z.; Wu, W.; Shi, H.Y.; Lin, Z.; Li, C.; Liu, X.X.; Sun, X. The back-deposition of dissolved Mn2+ to MnO2 cathodes for stable cycling in aqueous zinc batteries. Chem. Commun. 2022, 58, 4845–4848. [Google Scholar] [CrossRef]
  26. Wu, Q.; Li, S.; Han, Y.; Yang, C.; Gao, J. Constructing accommodational space in MnO2 cathode for Mn2+ transport and electrodeposition for aqueous zinc-ion batteries. Ionics 2022, 28, 4295–4301. [Google Scholar] [CrossRef]
  27. Ren, Y.; Li, H.; Rao, Y.; Zhou, H.; Guo, S. Aqueous MnO2/Mn2+ electrochemistry in batteries: Progress, challenges, and perspectives. Energy Environ. Sci. 2024, 17, 425–441. [Google Scholar] [CrossRef]
  28. Xu, Z.; Wang, J.; Zhang, W.; Shi, Z.; Feng, Y.; Liu, C.; Fu, H.; Yong, Z.; Li, Q. Hydrogen-bond chemistry inhibits Jahn–Teller distortion caused by Mn 3d orbitals for long-lifespan aqueous Zn//MnO2 batteries. J. Mater. Chem. A Mater. 2024, 12, 25491–25503. [Google Scholar] [CrossRef]
  29. Zou, Y.; Zhang, W.; Chen, N.; Chen, S.; Xu, W.; Cai, R.; Brown, C.L.; Yang, D.; Yao, X. Generating oxygen vacancies in MnO hexagonal sheets for ultralong life lithium storage with high capacity. ACS Nano 2019, 13, 2062–2071. [Google Scholar] [CrossRef]
  30. Yang, X.; Sun, Q.; Chai, L.; Chen, S.; Zhang, W.; Yang, H.Y.; Li, Z. α-MnO2 Cathode with Oxygen Vacancies Accelerated Affinity Electrolyte for Dual-Ion Co-Encapsulated Aqueous Aluminum Ion Batteries. Small 2024, 20, 2400335. [Google Scholar] [CrossRef]
  31. Wang, C.; Zeng, Y.; Xiao, X.; Wu, S.; Zhong, G.; Xu, K.; Wei, Z.; Su, W.; Lu, X. γ-MnO2 nanorods/graphene composite as efficient cathode for advanced rechargeable aqueous zinc-ion battery. J. Energy Chem. 2020, 43, 182–187. [Google Scholar] [CrossRef]
  32. Li, J.; Kang, L.; Luo, K.; Huang, Y.; Zhong, S.; Yan, D. Encapsulating Cu2Se into 3D porous carbon as high-voltage electrode materials for aluminum-ion batteries. Ceram. Int. 2023, 49, 2613–2618. [Google Scholar] [CrossRef]
  33. Zhao, Q.; Zhang, H.; Wang, X.; Xu, T.; Zhang, M.; Wang, Y.; Zhu, L.; Tong, S.; Zhou, X.; Li, J.; et al. Highly reversible and rapid charge transfer Zn-MnO2 battery by MnO2 nanosheet arrays anchored nanocellulose-based carbon aerogel. Adv. Compos. Hybrid. Mater. 2024, 7, 90. [Google Scholar] [CrossRef]
  34. Wang, X.; Yin, M.; Xue, H.; Su, Y.; Tian, S. Simple microwave synthesis and improved electrochemical performance of Nb-doped MnO2/reduced graphene oxide composite as anode material for lithium-ion batteries. Ionics 2018, 24, 2583–2590. [Google Scholar] [CrossRef]
  35. Saadi-motaallegh, S.; Javanbakht, M.; Omidvar, H.; Habibzadeh, S.; Gao, B. Nanostructure Design of MnO2/Partially Oxidized Graphene Nanocomposite Cathode via One-Step Pulse Potential Strategy: Toward Ultralong-Life Zinc-Ion Batteries. ACS Appl. Eng. Mater. 2023, 1, 2941–2953. [Google Scholar] [CrossRef]
  36. Rosaiah, P.; Divya, P.; Sambasivam, S.; Tighezza, A.M.; Kalaivani, V.; Muthukrishnaraj, A.; Ayyar, M.; Niyitanga, T.; Kim, H. Carbon based manganese oxide (MnO2, MnO2/MWCNT and MnO2/rGO) composite electrodes for high-stability Li-ion batteries. Carbon. Lett. 2024, 34, 215–225. [Google Scholar] [CrossRef]
  37. Zhao, Y.; Hao, H.; Song, T.; Wang, X.; Li, C.; Li, W. MnO2-graphene based composites for supercapacitors: Synthesis, performance and prospects. J. Alloys Compd. 2022, 914, 165343. [Google Scholar] [CrossRef]
  38. Ding, S.; Zhang, M.; Qin, R.; Fang, J.; Ren, H.; Yi, H.; Liu, L.; Zhao, W.; Li, Y.; Yao, L.; et al. Oxygen-Deficient β-MnO2@Graphene Oxide Cathode for High-Rate and Long-Life Aqueous Zinc Ion Batteries. Nanomicro Lett. 2021, 13, 173. [Google Scholar] [CrossRef]
  39. Shi, J.; Yu, C.; Zou, Z.; Zheng, S.; Zhang, X.; Wang, B. Reasonable Designing of Free-standing MnO2/Graphene Composite Membrane for Lithium-ion Storage. Chem. Res. Chin. Univ. 2024, 40, 508–512. [Google Scholar] [CrossRef]
  40. Liu, X.; Liang, B.; Hong, X.; Long, J. Electrochemical Performance of MnO2/Graphene Flower-like Microspheres Prepared by Thermally-Exfoliated Graphite. Front. Chem. 2022, 10, 870541. [Google Scholar] [CrossRef]
  41. Zhong, X.; Yang, C.; Zhao, Y.; Qiu, J.; Zang, L. Rational Design of Layered MnO2@Graphene with Hierarchical Structure for Flexible Quasisolid-State Aqueous Zinc-Ion Battery via Laser Activation. Adv. Mater. Technol. 2023, 8, 2201430. [Google Scholar] [CrossRef]
  42. Abdi, A.; Sarraf-Mamoory, R.; Stich, M.; Baumer, C.; Ullmann, F.; Krischok, S.; Bund, A. Optimization of in situ hydrothermal synthesis of birnessite MnO2/graphene composite: Thermodynamic insights and enhanced electrochemical performance for Li-ion battery anodes. J. Mater. Sci. Mater. Electron. 2025, 36, 1750. [Google Scholar] [CrossRef]
  43. Liu, Z.H.; Ooi, K.; Kanoh, H.; Tang, W.P.; Tomida, T. Swelling and delamination behaviors of birnessite-type manganese oxide by intercalation of tetraalkylammonium ions. Langmuir 2000, 16, 4154–4164. [Google Scholar] [CrossRef]
  44. Qiu, S.; Li, R.; Huang, Z.; Huang, Z.; Tsui, C.P.; He, C.; Han, X.; Yang, Y. Scalable sonochemical synthesis of petal-like MnO2/graphene hierarchical composites for high-performance supercapacitors. Compos. B Eng. 2019, 161, 37–43. [Google Scholar] [CrossRef]
  45. Zahoor, A.; Faizan, R.; Elsaid, K.; Hashmi, S.; Butt, F.A.; Ghouri, Z.K. Synthesis and experimental investigation of δ-MnO2/N-rGO nanocomposite for Li-O2 batteries applications. Chem. Eng. J. Adv. 2021, 7, 100115. [Google Scholar] [CrossRef]
  46. Ossonon, B.D.; Tavares, A.C. Innovative approach for the synthesis of graphene/MnO2 nanocomposites and their electrochemical behavior. Electrochem. Sci. Adv. 2022, 2, 2100029. [Google Scholar] [CrossRef]
  47. Wang, Y.; Guan, H.; Du, S.; Wang, Y. A facile hydrothermal synthesis of MnO2 nanorod–reduced graphene oxide nanocomposites possessing excellent microwave absorption properties. RSC Adv. 2015, 5, 88979–88988. [Google Scholar] [CrossRef]
  48. Biesinger, M.C.; Payne, B.P.; Grosvenor, A.P.; Lau, L.W.M.; Gerson, A.R.; Smart, R.S.C. Resolving surface chemical states in XPS analysis of first row transition metals, oxides and hydroxides: Cr, Mn, Fe, Co and Ni. Appl. Surf. Sci. 2011, 257, 2717–2730. [Google Scholar] [CrossRef]
  49. Wang, J.; Zhang, D.; Nie, F.; Zhang, R.; Fang, X.; Wang, Y. The role of MnO2 crystal morphological scale and crystal structure in selective catalytic degradation of azo dye. Environ. Sci. Pollut. Res. 2023, 30, 15377–15391. [Google Scholar] [CrossRef]
  50. Li, X.; Zhou, Q.; Yang, Z.; Zhou, X.; Qiu, D.; Qiu, H.; Huang, X.; Yu, Y. Unraveling the Role of Nitrogen-Doped Carbon Nanowires Incorporated with MnO2 Nanosheets as High Performance Cathode for Zinc-Ion Batteries. Energy Environ. Mater. 2023, 6, e12378. [Google Scholar] [CrossRef]
  51. Zhang, Y.; Han, X.; Huang, Z.H.; Lei, L.; Duan, X.; Li, H.; Ma, T. Shielding Mn3+ Disproportionation with Graphitic Carbon-Interlayered Manganese Oxide Cathodes for Enhanced Aqueous Energy Storage System. Small 2024, 20, 2401849. [Google Scholar] [CrossRef]
  52. Huang, L.; Luo, X.; Chen, C.; Jiang, Q. A high specific capacity aqueous zinc-manganese battery with a ε-MnO2 cathode. Ionics 2021, 27, 3933–3941. [Google Scholar] [CrossRef]
  53. Han, M.; Huang, J.; Liang, S.; Shan, L.; Xie, X.; Yi, Z.; Wang, Y.; Guo, S.; Zhou, J. Oxygen Defects in β-MnO2 Enabling High-Performance Rechargeable Aqueous Zinc/Manganese Dioxide Battery. IScience 2020, 23, 100797. [Google Scholar] [CrossRef]
  54. Zhong, W.; Zhao, R.; Zhu, Y.; Xu, Y.; Chen, W.; Peng, C. Vacancy Engineering on MnSe Cathode Enables High-Rate and Stable Zinc-Ion Storage. Adv. Funct. Mater. 2024, 35, 2419720. [Google Scholar] [CrossRef]
  55. Zhu, K.; Wang, C.; Camargo, P.H.C.; Wang, J. Investigating the effect of MnO2 band gap in hybrid MnO2-Au materials over the SPR-mediated activities under visible light. J. Mater. Chem. A 2019, 7, 925–931. [Google Scholar] [CrossRef]
  56. Biesinger, M.C. Accessing the robustness of adventitious carbon for charge referencing (correction) purposes in XPS analysis: Insights from a multi-user facility data review. Appl. Surf. Sci. 2022, 597, 153681. [Google Scholar] [CrossRef]
  57. Deng, S.; Tie, Z.; Yue, F.; Cao, H.; Yao, M.; Niu, Z. Rational Design of ZnMn2O4 Quantum Dots in a Carbon Framework for Durable Aqueous Zinc-Ion Batteries. Angew. Chem. Int. Ed. 2022, 61, e202115877. [Google Scholar] [CrossRef] [PubMed]
  58. Awang, H.; Talalah, N.I. Synthesis of Reduced Graphene Oxide-Titanium (rGO-TiO2) Composite Using a Solvothermal and Hydrothermal Methods and Characterized via XRD and UV-Vis. Nat. Resour. 2019, 10, 17–28. [Google Scholar] [CrossRef]
  59. Susee, S.K.; Sandhya, S.; Kumar, M.S.; Chidhambararajan, B. Reduced graphene oxide–MnO2 nanocomposites by hydrothermal method for histamine sensor and photocatalytic activity. J. Mater. Sci. Mater. Electron. 2024, 35, 1948. [Google Scholar] [CrossRef]
  60. Jia, H.; Cai, Y.; Lin, J.; Liang, H.; Qi, J.; Cao, J.; Feng, J.; Fei, W.D. Heterostructural Graphene Quantum Dot/MnO2 Nanosheets toward High-Potential Window Electrodes for High-Performance Supercapacitors. Adv. Sci. 2018, 5, 1700887. [Google Scholar] [CrossRef]
  61. Zhang, Y.; Bian, Y.; Lv, Z.; Han, Y.; Lin, M.C. Aqueous Aluminum Cells: Mechanisms of Aluminum Anode Reactions and Role of the Artificial Solid Electrolyte Interphase. ACS Appl. Mater. Interfaces 2021, 13, 37091–37101. [Google Scholar] [CrossRef]
  62. Zeng, K.; Zhu, J. Surface morphology, elastic modulus and hardness of thin film cathodes for Li-ion rechargeable batteries. Mech. Mater. 2015, 91, 323–332. [Google Scholar] [CrossRef]
  63. Hamade, F.; Radich, E.; Davis, V.A. Microstructure and electrochemical properties of high performance graphene/manganese oxide hybrid electrodes. RSC Adv. 2021, 11, 31608–31620. [Google Scholar] [CrossRef] [PubMed]
  64. Nie, R.; Nie, Y.; Wu, J.; Yu, L.; Liu, L.; Xi, J. An aqueous alkaline zinc–sulfur flow battery. Chem. Commun. 2024, 60, 2946–2949. [Google Scholar] [CrossRef] [PubMed]
  65. Ette, P.M.; Bosubabu, D.; Ramesha, K. Graphene anchored mesoporous MnO2 nanostructures as stable and high-performance anode materials for Li-ion batteries. Electrochim. Acta 2022, 414, 140164. [Google Scholar] [CrossRef]
  66. Wang, X.; Xi, Z.; Zhao, Q. Progress on aqueous rechargeable aluminium metal batteries. Ind. Chem. Mater. 2025, 3, 7–30. [Google Scholar] [CrossRef]
  67. Zheng, X.; Han, C.; Lee, C.S.; Yao, W.; Zhi, C.; Tang, Y. Materials challenges for aluminum ion based aqueous energy storage devices: Progress and prospects. Prog. Mater. Sci. 2024, 143, 101253. [Google Scholar] [CrossRef]
  68. Li, L.; Jia, S.; Shi, Y.; Wang, C.; Qiu, H.; Ji, Y.; Cao, M.; Zhang, D. Recent progress in aluminum anodes for high-performance rechargeable aqueous Al-ion batteries. Inorg. Chem. Front. 2024, 11, 2246–2259. [Google Scholar] [CrossRef]
  69. Meng, H.; Ran, Q.; Zhu, M.H.; Zhao, Q.Z.; Han, G.F.; Wang, T.H.; Wen, Z.; Lang, X.Y.; Jiang, Q. Benzoquinone-Lubricated Intercalation in Manganese Oxide for High-Capacity and High-Rate Aqueous Aluminum-Ion Battery. Small 2024, 20, 2310722. [Google Scholar] [CrossRef]
  70. Siamionau, U.V.; Aniskevich, Y.M.; Ragoisha, G.A.; Streltsov, E.A. MnO2 electrodeposition at the positive electrode of zinc-ion aqueous battery containing Zn2+ and Mn2+ cations. J. Solid. State Electrochem. 2023, 27, 1911–1918. [Google Scholar] [CrossRef]
  71. Hu, E.; Jia, B.E.; Zhu, Q.; Xu, J.; Loh, X.J.; Chen, J.; Pan, H.; Yan, Q. Engineering High Voltage Aqueous Aluminum-Ion Batteries. Small 2024, 2309252. [Google Scholar] [CrossRef]
  72. Yang, X.; Gu, H.; Chai, L.; Chen, S.; Zhang, W.; Yang, H.Y.; Li, Z. Construction of V2O5@MXene Cathodes toward a High Specific Capacity Aqueous Aluminum-Ion Battery. Nano Lett. 2024, 24, 8542–8549. [Google Scholar] [CrossRef]
  73. Yuan, X.; Yuan, X.; Zhang, S.; Chen, P.; Wu, X.; Ye, J.; Liu, L.; Fu, L.; Wang, T.; Ozoemena, K.I.; et al. An Aqueous Rechargeable Al-Ion Battery Based on Cobalt Hexacyanoferrate and Al Metal. Adv. Energy Mater. 2024, 14, 2302712. [Google Scholar] [CrossRef]
  74. Mohammadi, E.; Taheri-Nassaj, E.; Shahalizade, T.; Shahroudi, H. Synthesis and electrochemical performance of SnSe/PANI composite as a long-cycle-life lithium-ion battery anode. J. Electroanal. Chem. 2025, 979, 118930. [Google Scholar] [CrossRef]
  75. Jin, D.; Dong, X.; Xin, S.; Yang, L.; Liu, J.; Pang, Q. Effect of MnSO4 concentration on the electrochemical performance of β-MnO2/3D graphene-carbon nanotube hybrids cathode for aqueous zinc-ion batteries. Ionics 2024, 30, 3329–3338. [Google Scholar] [CrossRef]
Figure 1. (a) Graphene, δ-MnO2, and δ-MnO2/Graphene XRD patterns. The tiny graphene peak in the composite material is highlighted in the inset. (b) Graphene, δ-MnO2, and δ-MnO2/Graphene Raman spectra. (c) SEM images of δ-MnO2 and δ-MnO2/Graphene composites at various magnifications reveal the characteristic nanosheet morphology of δ-MnO2. (d) TEM images of δ-MnO2/Graphene in different magnifications.
Figure 1. (a) Graphene, δ-MnO2, and δ-MnO2/Graphene XRD patterns. The tiny graphene peak in the composite material is highlighted in the inset. (b) Graphene, δ-MnO2, and δ-MnO2/Graphene Raman spectra. (c) SEM images of δ-MnO2 and δ-MnO2/Graphene composites at various magnifications reveal the characteristic nanosheet morphology of δ-MnO2. (d) TEM images of δ-MnO2/Graphene in different magnifications.
Processes 13 03551 g001
Figure 2. XPS analysis of synthesized δ-MnO2 and δ-MnO2/Graphene: (a) survey spectrum, high-resolution (b) Mn 2p, (c) O 1s, and (d) C 1s spectrum.
Figure 2. XPS analysis of synthesized δ-MnO2 and δ-MnO2/Graphene: (a) survey spectrum, high-resolution (b) Mn 2p, (c) O 1s, and (d) C 1s spectrum.
Processes 13 03551 g002
Figure 3. (a) δ-MnO2 and δ-MnO2/Graphene cathode XRD patterns at different states of charge and discharge. (b) SEM images of fresh and post-second discharge δ-MnO2 and δ-MnO2/Graphene cathodes. (c) AFM images of δ-MnO2 before and after the second discharge. All measurements were conducted in the full cell battery with a TAl anode.
Figure 3. (a) δ-MnO2 and δ-MnO2/Graphene cathode XRD patterns at different states of charge and discharge. (b) SEM images of fresh and post-second discharge δ-MnO2 and δ-MnO2/Graphene cathodes. (c) AFM images of δ-MnO2 before and after the second discharge. All measurements were conducted in the full cell battery with a TAl anode.
Processes 13 03551 g003
Figure 4. SEM images, EDS elemental peaks, and element distribution maps of the δ-MnO2/Graphene cathode: (a) in its fresh state and (b) after the first charge in a full cell battery with a TAl anode.
Figure 4. SEM images, EDS elemental peaks, and element distribution maps of the δ-MnO2/Graphene cathode: (a) in its fresh state and (b) after the first charge in a full cell battery with a TAl anode.
Processes 13 03551 g004
Figure 5. Electrochemical properties of the full cell batteries with δ-MnO2 and δ-MnO2/Graphene cathodes, and a TAl anode: (a) Charge/discharge profiles of the second cycle at 180 mA g−1. (b) Cycling test at 180 mA g−1. (c) Coulombic efficiency. (d) Rate capability. (e) Nyquist plots measured at frequencies between 200 kHz and 0.1 Hz after the third charge. In (b,d,e), the measured or fitted data are represented by symbols. (lines are drawn only to guide the eye).
Figure 5. Electrochemical properties of the full cell batteries with δ-MnO2 and δ-MnO2/Graphene cathodes, and a TAl anode: (a) Charge/discharge profiles of the second cycle at 180 mA g−1. (b) Cycling test at 180 mA g−1. (c) Coulombic efficiency. (d) Rate capability. (e) Nyquist plots measured at frequencies between 200 kHz and 0.1 Hz after the third charge. In (b,d,e), the measured or fitted data are represented by symbols. (lines are drawn only to guide the eye).
Processes 13 03551 g005
Figure 6. Analysis of the Zn substrate and δ-MnO2/Graphene cathode charged and discharged at 180 mA g−1: (a) XRD patterns of the zinc substrate in its fresh state and after the third charge. (b) SEM image of fresh zinc. (c) SEM image of zinc after the third charge. (d) EDS mapping confirms the formation of Zn-Al alloy following the third charge by displaying the distribution of Al and Zn. (e) SEM image, (f) EDS elemental peaks, and (g) EDS mapping results of the δ-MnO2/Graphene cathode after the third charge.
Figure 6. Analysis of the Zn substrate and δ-MnO2/Graphene cathode charged and discharged at 180 mA g−1: (a) XRD patterns of the zinc substrate in its fresh state and after the third charge. (b) SEM image of fresh zinc. (c) SEM image of zinc after the third charge. (d) EDS mapping confirms the formation of Zn-Al alloy following the third charge by displaying the distribution of Al and Zn. (e) SEM image, (f) EDS elemental peaks, and (g) EDS mapping results of the δ-MnO2/Graphene cathode after the third charge.
Processes 13 03551 g006
Figure 7. Electrochemical properties of full cell batteries using δ-MnO2 and δ-MnO2/Graphene cathodes, and Zn-Al anode: (a) Charge/discharge profiles of the second cycle at 180 mA g−1. (b) Cycling test at 180 mA g−1. (c) Coulombic efficiency. (d) Rate capability. In (d) the measured data is represented by symbols. (lines are drawn only to guide the eye). CV analysis of batteries with (e) δ-MnO2 and (f) δ-MnO2/graphene cathodes, a Zn-Al anode, and a 0.5 M Al2(SO4)3, 0.4 M MnSO4 electrolyte, highlighting the differences between the first and subsequent cycles. CV measurements were performed within a voltage range of 0.4 V to 1.9 V at a scan rate of 10 mVs−1.
Figure 7. Electrochemical properties of full cell batteries using δ-MnO2 and δ-MnO2/Graphene cathodes, and Zn-Al anode: (a) Charge/discharge profiles of the second cycle at 180 mA g−1. (b) Cycling test at 180 mA g−1. (c) Coulombic efficiency. (d) Rate capability. In (d) the measured data is represented by symbols. (lines are drawn only to guide the eye). CV analysis of batteries with (e) δ-MnO2 and (f) δ-MnO2/graphene cathodes, a Zn-Al anode, and a 0.5 M Al2(SO4)3, 0.4 M MnSO4 electrolyte, highlighting the differences between the first and subsequent cycles. CV measurements were performed within a voltage range of 0.4 V to 1.9 V at a scan rate of 10 mVs−1.
Processes 13 03551 g007
Figure 10. Nyquist plots measured over a frequency range of 200 kHz to 0.1 Hz after the (a) 3rd and (b) 100th charge. (c) Warburg plots fitted from the low-frequency region of the EIS spectra are presented in (b). The measured and fitted impedance data are represented by symbols (lines are drawn only to guide the eye).
Figure 10. Nyquist plots measured over a frequency range of 200 kHz to 0.1 Hz after the (a) 3rd and (b) 100th charge. (c) Warburg plots fitted from the low-frequency region of the EIS spectra are presented in (b). The measured and fitted impedance data are represented by symbols (lines are drawn only to guide the eye).
Processes 13 03551 g010
Table 1. Parameters of the equivalent circuit obtained from EIS data shown in Figure 5e, analyzed with EC-Lab software.
Table 1. Parameters of the equivalent circuit obtained from EIS data shown in Figure 5e, analyzed with EC-Lab software.
CathodeR1
(Ω)
R2
(Ω)
R3
(Ω)
Q1
(μF.s1−n)
Q2
(μF.s1−n)
δ-MnO22.60 ± 0.0536.9 ± 2.61333 ± 30150 ± 514.5 ± 3.1
δ-MnO2/Graphene1.78 ± 0.045.90 ± 0.481110 ± 20530 ± 1280 ± 4
Table 2. The electrochemical characteristics of recently published AAIBs based on oxide cathode materials are contrasted with the findings of this study.
Table 2. The electrochemical characteristics of recently published AAIBs based on oxide cathode materials are contrasted with the findings of this study.
No.CathodeAnodeElectrolyteVoltage (V)Capacity (mAh g−1), Cycle NumberRef.
1AlxMnO2Zn-Al2 M Al(OTF)31.6460, 80[8]
2V2O5@MXeneTAl5 M Al(OTF)31626, 100[72]
3MnO2Zn-Al3 M Al[TFSI]31.75450, 400[1]
4AlxMnO2a-Al0.5 M Al2(SO4)31.8<60, 200[9]
5AlxMn(1−x)O2TAl2 M Al(OTF)3,
0.5 M MnSO4
1.35320, 65[6]
6AlxMnO2/CMXene/ EAl97Ce32 M Al(OTF)3,0.2 M Mn(OTF)21.4306, 500[10]
7MnAlxO2Al16.57 M AlCl3,
3.86 M MnSO4
1.9285, 500[23]
8V6O13Al3 M Al(OTF)30.7100, 1400[73]
9AlxMnO2EAl82Cu182 M Al(OTF)31.5400, 400[12]
10V-δ-MnO2Al alloy2 M Al(OTF)31.14518, 400[11]
11Ov-MnO2TAl5 M Al(OTF)31.5128, 200[22]
12AlxMn(1−x)O2/GrapheneZn-Al0.5 M Al2(SO4)3,
0.4 M MnSO4
1.63746, 352This work
Table 3. Parameters of the equivalent circuit obtained from EIS data shown in Figure 10a,b, analyzed with EC-Lab software.
Table 3. Parameters of the equivalent circuit obtained from EIS data shown in Figure 10a,b, analyzed with EC-Lab software.
CathodeR1
(Ω)
R2
(Ω)
R3
(Ω)
Q1
(μF.s1−n)
Q2
(μF.s1−n)
δ-MnO2
After 3rd cycle
4.12 ± 0.32 7.1 ± 0.43891 ± 10087.7 ± 5.142.1 ± 3.4
δ-MnO2/Graphene
After 3rd cycle
3.9 ± 0.214.4 ± 0.13055 ± 633.1 ± 0.469.5 ± 1.2
δ-MnO2
After 100th cycle
7.40 ± 0.0547.2 ± 0.6276 ± 326.7 ± 1.728.6 ± 0.4
δ-MnO2/Graphene
After 100th cycle
6.70 ± 0.0412.1 ± 0.1136 ± 17.0 ± 0.1141 ± 1
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Abdi, A.; Sarraf-Mamoory, R.; Stich, M.; Baumer, C.; Kurniawan, M.; Bund, A. Graphene-Stabilized δ-MnO2 Cathode for High-Capacity Aqueous Aluminum-Ion Batteries. Processes 2025, 13, 3551. https://doi.org/10.3390/pr13113551

AMA Style

Abdi A, Sarraf-Mamoory R, Stich M, Baumer C, Kurniawan M, Bund A. Graphene-Stabilized δ-MnO2 Cathode for High-Capacity Aqueous Aluminum-Ion Batteries. Processes. 2025; 13(11):3551. https://doi.org/10.3390/pr13113551

Chicago/Turabian Style

Abdi, Azadeh, Rasoul Sarraf-Mamoory, Michael Stich, Christoph Baumer, Mario Kurniawan, and Andreas Bund. 2025. "Graphene-Stabilized δ-MnO2 Cathode for High-Capacity Aqueous Aluminum-Ion Batteries" Processes 13, no. 11: 3551. https://doi.org/10.3390/pr13113551

APA Style

Abdi, A., Sarraf-Mamoory, R., Stich, M., Baumer, C., Kurniawan, M., & Bund, A. (2025). Graphene-Stabilized δ-MnO2 Cathode for High-Capacity Aqueous Aluminum-Ion Batteries. Processes, 13(11), 3551. https://doi.org/10.3390/pr13113551

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop