Previous Article in Journal
Use of Electrochemical Impedance Spectroscopy, Capacity, and Electrochemical Noise Measurements to Study Aging of Lithium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing the Structural Stability and Electrochemical Performance of δ-MnO2 Cathodes via Fe3+ Doping for Aqueous Zinc-Ion Batteries

Ningbo Key Laboratory of Biomolecular Intelligent Design and Manufacturing, School of Biological and Chemical Engineering, NingboTech University, Ningbo 315100, China
*
Authors to whom correspondence should be addressed.
Solids 2025, 6(3), 45; https://doi.org/10.3390/solids6030045 (registering DOI)
Submission received: 3 July 2025 / Revised: 11 August 2025 / Accepted: 12 August 2025 / Published: 14 August 2025

Abstract

Due to its unique layered structure that facilitates ion intercalation and deintercalation, δ-MnO2 has emerged as a promising cathode material for aqueous zinc-ion batteries (ZIBs). However, its structural collapse and Mn dissolution during prolonged cycling significantly limit its practical application. In this study, we demonstrate that metal ion doping, particularly with Fe3+, can effectively stabilize the δ-MnO2 structure and enhance its electrochemical performance. Through a hydrothermal synthesis approach, δ-MnO2 materials with varying Fe3+ doping ratios are prepared and systematically investigated. Among them, the sample with a Mn:Fe molar ratio of 20:1 exhibits the best performance, maintaining the layered δ-MnO2 phase while significantly increasing Mn3+ content and promoting the formation of oxygen vacancies. At a current density of 0.5 A·g−1, the iron-doped sample exhibited an initial specific capacity of 116.24 mAh·g−1, with a capacity retention rate of 41.7% after 200 cycles. In contrast, the undoped δ-MnO2 showed an initial specific capacity of only 85.15 mAh·g−1, with a capacity retention rate of merely 19.9% after 200 cycles. The results suggest that Fe3+ doping not only suppresses Mn dissolution but also improves structural stability and Zn2+ transport kinetics. This work provides new insights into the development of durable Mn-based cathode materials for aqueous ZIBs.

1. Introduction

The global transition toward sustainable energy systems demands the development of safe, low-cost, and high-performance energy storage technologies. While lithium-ion batteries (LIBs) currently dominate the market due to their high energy density and mature manufacturing infrastructure, their large-scale deployment is increasingly hindered by the scarcity of lithium resources, rising material costs, and safety concerns associated with flammable organic electrolytes [1,2,3]. These challenges have caused intensive research into alternative technologies based on abundant, environmentally benign elements, among which aqueous zinc-ion batteries (ZIBs) have emerged as a highly promising candidate [4,5,6].
ZIBs offer several compelling advantages: zinc metal is naturally abundant, inexpensive, non-toxic, and features a high theoretical capacity of 820 mAh·g−1. Furthermore, ZIBs employ mild aqueous electrolytes, which not only improve ionic conductivity and manufacturing safety but also reduce environmental impact, making them suitable for large-scale, stationary energy storage applications [7,8,9]. In this system, metallic zinc is commonly used as the anode due to its low redox potential (−0.762 V vs. SHE), reversible stripping/plating behavior, and compatibility with aqueous media [10]. However, the commercial viability of ZIBs is largely constrained by the lack of cathode materials that simultaneously deliver high capacity, long-term structural integrity, and stable electrochemical cycling. Among the various explored cathodes—such as Prussian blue analogues, vanadium-based oxides, and organic materials—manganese dioxide (MnO2) stands out due to its low cost, environmental friendliness, multiple redox states, and rich polymorphic diversity (α, β, γ, δ, etc.) [11,12,13]. In particular, δ-MnO2, with its layered structure and large interlayer spacing (~7 Å), facilitates the intercalation and deintercalation of Zn2+ and H⁺ ions, and thus has attracted considerable interest as a ZIB cathode [14,15,16].
Nevertheless, δ-MnO2 faces critical drawbacks during long-term cycling. The Jahn–Teller distortion associated with Mn3+ leads to structural instability and Mn dissolution, causing the collapse of its layered framework and a rapid decline in the capacity retention rate [17]. To mitigate these issues, recent studies have explored cation doping strategies—particularly with transition metal or rare-earth metal ions. Transition metal ions can stabilize the MnO2 lattice and suppress Jahn–Teller distortion, while rare-earth metal ions, due to their larger ionic radii, expand the interlayer spacing upon incorporation, thereby facilitating ion transport during electrochemical processes. [18,19,20]. For example, Ding et al. investigated Cr-doped δ-MnO2 and found that the incorporation of Cr enhanced the interaction between the δ-MnO2 electrode material and Zn2+, thereby improving the material’s stability [21]. However, the effects of different dopants and doping levels on the crystal structure and electrochemical behavior of δ-MnO2 remain underexplored, especially in terms of maintaining phase integrity while enhancing cycling stability.
In this work, we address these challenges by developing a series of metal-ion-doped δ-MnO2 cathodes via a simple and scalable hydrothermal synthesis approach, with a particular focus on systematic Fe3+ doping. We investigate the influence of various Fe3+ doping ratios (Mn:Fe = 10:1, 20:1, 40:1) on the crystal structure, surface chemistry, and electrochemical performance of δ-MnO2. For broader comparison, we also examine Cu2+ and Ru3+ dopants in the matrix of δ-MnO2. Since Cu2+ exhibits better valence compatibility with Mn2+, its doping is expected to effectively modulate the structural stability of MnO2 materials. Indeed, Liu et al. have systematically investigated Cu2+ doping in MnO2-based systems [22]. In contrast, studies on Ru3+ incorporation into MnO2 lattices remain scarce, though Zheng et al. reported that Ru doping in TiO2 significantly alters the kinetic processes and structural stability [23]. Therefore, we selected these two dopants for comparative studies to elucidate their distinct effects. Structural characterizations (XRD, SEM, and XPS) confirm that moderate Fe3+ incorporation (20:1) preserves the δ-MnO2 phase while increasing Mn3+ content and oxygen vacancies, both of which contribute to enhanced ion diffusion and cycling durability.
Notably, the Fe3+-doped δ-MnO2 (b-FeMO) sample exhibits a specific capacity of 116.2 mAh·g−1 and a capacity retention rate of 41.7% after 200 cycles at 0.5 A·g−1, significantly outperforming undoped δ-MnO2 (which retains only 19.9% under the same conditions). Post-cycling analyses confirm that Fe3+ doping mitigates Mn dissolution, preserves the layered morphology, and maintains a favorable redox environment. This work not only highlights the crucial role of Fe3+ in reinforcing structural stability and prolonging cycling life but also offers comparative insights into the roles of different dopants. More broadly, it provides practical design guidance for engineering robust Mn-based cathodes for aqueous zinc-ion batteries.

2. Materials and Methods

2.1. Materials Synthesis

Preparation of δ-MnO2 (MO): The δ-MnO2 was synthesized via a hydrothermal route. Briefly, 0.25 g of KMnO4 and 0.066 g of Mn(NO3)2·4H2O were dissolved separately in deionized water and stirred for 30 min. The Mn(NO3)2 solution was then slowly added to the KMnO4 solution under continuous stirring. The mixture was transferred to a 100 mL Teflon-lined stainless-steel autoclave and heated at 160 °C for 10 h. After natural cooling, the product was washed alternately with deionized water and ethanol, followed by drying at 60 °C overnight. The product was labeled as MO.
Preparation of Fe-doped δ-MnO2 (FeMO): The synthesis followed the same hydrothermal procedure as MO, with additional Fe(NO3)3·9H2O added to the KMnO4 solution at different molar ratios of Mn:Fe (10:1, 20:1, and 40:1). The resulting products were labeled as a-FeMO, b-FeMO, and c-FeMO, respectively.
Preparation of Cu- and Ru-doped δ-MnO2: Similarly, Cu(NO3)2 (0.017 g) and RuCl3 (0.019 g) were introduced into the synthesis system, following the same hydrothermal process. The products were labeled as CuMO and RuMO, respectively.

2.2. Characterization

X-ray diffraction (XRD, Rigaku Ultima IV, Tokyo, Japan) was used to determine the crystalline structure. X-ray photoelectron spectroscopy (XPS, Thermo Scientific K-Alpha, Waltham, MA, USA) was conducted to analyze surface chemical states. Scanning electron microscopy (SEM, ZEISS Sigma 300, Oberkochen, Germany) along with energy-dispersive X-ray spectroscopy (EDS) and elemental mapping was used for morphology and composition analysis.

2.3. Electrochemical Measurements

The cathode slurry was prepared by mixing active material, acetylene black, and PVDF binder at a 7:2:1 weight ratio. The PVDF was dissolved in NMP and stirred at 45 °C. The active material and acetylene black were added and stirred for 30 min. The slurry was coated on iron foil and dried at 75 °C for 2 h under vacuum. CR2032 coin cells were assembled using zinc foil as the anode, glass fiber as the separator, and 2 M ZnSO4 aqueous solution as the electrolyte. Cathodes were prepared with mass loadings of 0.39~0.78 mg/cm2 for the active material.
Galvanostatic charge/discharge (GCD) and cycling tests were performed on a battery testing system (Neware, Shenzhen, China) within 0.7–1.8 V. Cyclic voltammetry (CV) was conducted using an electrochemical workstation (CHI760E, Shanghai, China) at a scan rate of 0.5 mV s−1.

3. Results

3.1. Structural and Morphological Characterization

Figure 1a shows the X-ray diffraction (XRD) patterns of the pristine δ-MnO2 (MO) and Fe3+-doped δ-MnO2 (FeMO) samples with varying doping ratios. All major diffraction peaks of MO correspond well to the reported δ-MnO2 phase (PDF#80-1098) [22], with apparent reflections at 12.5°, 25°, and 37°, corresponding to the (001), (002), and (111) crystal planes, respectively. The b-FeMO sample, with a Mn:Fe molar ratio of 20:1, retained these characteristic δ-MnO2 peaks without the emergence of impurity phases, indicating that Fe3+ doping at this level does not disrupt the host lattice structure. Furthermore, the sharpness and intensity of the diffraction peaks in b-FeMO suggest good crystallinity. In contrast, the a-FeMO sample (10:1) exhibited an increased intensity of the (002) plane, implying preferential growth along specific crystal orientations induced by higher Fe3+ incorporation. Conversely, c-FeMO (40:1) shows weakened and broadened diffraction peaks, indicative of partial structural disruption or amorphization due to insufficient doping at a low level of Fe3+ concentration. For the Cu2+- and Ru3+-doped samples, the XRD patterns (Figure 1b) reveal that CuMO (Cu2+-doped) retained the δ-MnO2 crystal structure, whereas RuMO (Ru3+-doped) shows distinct new peaks corresponding to an altered Mn/O stoichiometry, likely forming a different manganese oxide phase (PDF#15-0604), suggesting that Ru3+ doping significantly modifies the crystal lattice.
Figure 2 presents the SEM images of MO and b-FeMO, along with their energy-dispersive X-ray spectroscopy (EDS) elemental mapping. The MO sample consists of uniform flower-like microspheres assembled from interconnected nanosheets, in agreement with those reported in the literature findings [24]. Such a morphological structure can provide a high surface area beneficial for electrochemical activity. Similarly, the b-FeMO sample maintains the flower-like morphology, indicating that Fe3+ doping at 20:1 does not alter the microscale morphology of the δ-MnO2 but helps stabilize the structure. The EDS elemental mapping of b-FeMO demonstrates a homogeneous distribution of Mn, O, and Fe throughout the structure, confirming the successful and uniform incorporation of Fe3+ into the δ-MnO2 framework. In comparison, CuMO retained the nanosheet-assembled microsphere morphology (Figure S1), whereas RuMO displayed a different morphology, primarily composed of rod-like particles with residual flower-like structures (Figure S2). This morphological shift in RuMO is consistent with the structural changes observed in its XRD pattern, supporting the conclusion that Ru3+ doping affects both the crystal structure and the morphology [25].
Figure 3 shows the high-resolution X-ray photoelectron spectroscopy (XPS) spectra of the Mn 2p and O 1s orbitals for all samples, with their corresponding full spectra shown in Figure S3. As was expected, the majority of the dopants went into the MO matrix with the following dopant: the Mn ratio close to 1:20 as revealed by the XPS elemental analysis (Table S1), indicating the successful doping approach. The Mn 2p spectra exhibit the characteristic Mn 2p3/2 and Mn 2p1/2 peaks at approximately 641.4 eV and 653.1 eV, respectively, with a spin-energy separation of ~11.7 eV, typical of MnO2 [11]. The deconvolution of the Mn 2p3/2 peak indicates a higher Mn3+ to Mn4+ ratio in b-FeMO compared to MO, with b-FeMO showing a Mn3+ content of 74.27%, significantly higher than MO (66.07%) and the other doped ones. This increase in Mn3+ content correlates with enhanced oxygen vacancy concentration with Fe3+ doping, which can facilitate improved electronic conductivity and ion diffusion kinetics. The O 1s spectra further support this observation: b-FeMO shows a stronger Mn–O–H peak, reflecting the formation of more oxygen-related defect sites due to Fe3+ doping. These features are beneficial for Zn2+ intercalation/deintercalation and are aligned with the improved electrochemical performance observed in b-FeMO [26].
Taken together, the structural (XRD), morphological (SEM/EDS), and chemical state (XPS) analyses consistently confirm the successful doping of Fe3+ into the δ-MnO2 lattice without disrupting its crystal structure. The 20:1 Mn:Fe doping ratio (b-FeMO) was found to be optimal, preserving the structural integrity while introducing a higher Mn3+ content and enhanced oxygen vacancies. Although Cu or Ru could be also doped into the MO framework (CuMO or RuMO), the main structure or morphology was not persevered. These modifications in b-FeMO collectively contribute to the improved electrochemical performance and cycling stability of b-FeMO as a cathode material for aqueous zinc-ion batteries, validating the conclusions of this study.

3.2. Electrochemical Performance

Figure 4a–d presents the cyclic voltammetry (CV) curves of MO, b-FeMO, CuMO, and RuMO electrodes at a scan rate of 0.5 mV·s−1. All electrodes exhibit two prominent redox peaks, corresponding to the reversible intercalation/deintercalation of H⁺ and Zn2+ into/from the MnO2 layers [20]. For the pristine δ-MnO2 (MO), the peaks are broad and less defined, suggesting relatively sluggish electrochemical kinetics and poor reversibility. In contrast, the b-FeMO sample shows sharper and more symmetric redox peaks, indicating increased redox kinetics and improved electrochemical reversibility after Fe3+ doping. The enhanced CV response can be attributed to the increased Mn3+ content (confirmed by XPS), which facilitates pseudocapacitive charge storage via oxygen vacancy-mediated charge transfer pathways.
The improved kinetics in b-FeMO are also reflected in the smaller peak potential separation (ΔE) [27], which decreased from ~190 mV in MO to ~135 mV in b-FeMO, highlighting lower electrochemical polarization. This trend is consistent with literature reports showing that Fe3+ doping in Mn-based oxides enhances structural stability and electrochemical kinetics by tuning electronic configurations and suppressing Jahn–Teller distortion [28].
GCD measurements (Figure 5a) further support the above conclusions. Notably, the b-FeMO electrode displayed much higher discharge capacities than other electrodes with different Fe loadings (a-FeMO and c-FeMO) (Figure S4), confirming the optimal doping content in the b-FeMO sample. As shown in Figure 5a, MO delivered an initial discharge capacity of 85.15 mAh·g−1 at a current density of 0.5 A·g−1, which degraded rapidly to 16.95 mAh·g−1 after 200 cycles, corresponding to only a 19.9% retention rate. In comparison, b-FeMO exhibited an initial capacity of 116.24 mAh·g−1 with 48.47 mAh·g−1 remaining after 200 cycles—retaining 41.7% of its capacity, thus outperforming other Cu and Ru doped ones. This clearly demonstrates the superior cycling stability induced by Fe3+ doping.
Comparatively, similar MnO2-based cathodes reported in the literature often have a lower retention rate or require complex synthesis procedures. For instance, Kamenskii et al. reported MnO2 with 52% retention rate after 100 cycles at 0.3 A·g−1 [29], while our b-FeMO achieved 61.4% at a higher rate of 0.5 A·g−1, indicating better structural robustness.
The rate capability of the b-FeMO electrode (Figure 5b) was further evaluated at varying current densities ranging from 0.1 to 2.0 A·g−1, with their GCD curves shown in Figure S5. b-FeMO exhibited average discharge capacities of 113.2, 98.7, 83.4, 71.2, and 59.5 mAh·g−1 at current densities of 0.1, 0.2, 0.5, 1, and 2.0 A·g−1, respectively. Notably, when the current density returned to 0.1 A·g−1, 92.3% of the original capacity was recovered, suggesting excellent structural stability and rate adaptability. This rate performance significantly exceeds that of MO and other doped ones, which showed severe capacity degradation under similar conditions.
Table 1 presents the comparative results of electrochemical performance among different materials. As evidenced by both the capacity retention rate and average Coulombic efficiency, b-FeMO demonstrates superior performance compared to the other three materials, exhibiting relatively slower performance degradation after 200 cycles.
In comparison, unmodified δ-MnO2 cathodes typically suffer from fast capacity decay and limited rate response. For example, Feng et al. [30] and Ni et al. [31] reported the MnO2 cathodes with a capacity drop of more than 50% when current density was increased from 0.1 to 1.0 A·g−1. In contrast, our b-FeMO retains over 63% capacity at 2.0 A·g−1, highlighting the kinetic superiority of Fe3+-doped δ-MnO2.

3.3. Post-Cycling SEM and XPS Analysis:

To investigate the structural evolution and chemical composition of δ-MnO2 electrodes during extended cycling, post-mortem SEM and XPS characterizations were conducted for both pristine MO and Fe3+-doped b-FeMO electrodes before and after 200 cycles.
As shown in Figure 6, the surface morphology of the MO electrode underwent significant degradation after 200 charge–discharge cycles at 0.5 A·g−1. The pristine MO electrode originally exhibited spherical flower-like microspheres composed of thin nanosheets (Figure 6a,b), which are beneficial for Zn2+ diffusion and electrolyte penetration. However, after cycling (Figure 6c,d), these structures became fragmented, and extensive cracking and collapse of the microspheres was observed. This mechanical degradation likely results from Jahn–Teller distortions and Mn dissolution, both of which are known to cause lattice instability in δ-MnO2 during repeated Zn2+ insertion/extraction [32].
In contrast, the b-FeMO electrode (Figure 6e–h) retained mostly its morphological integrity after 200 cycles. The nanosheet-based microspheres remained well-defined with minimal cracking, suggesting that Fe3+ doping effectively stabilizes the δ-MnO2 layered structure. The improved mechanical robustness is likely due to the strengthened Mn–O bonding environment and reduction in lattice strain, which inhibit collapse and particle pulverization.
To further probe the chemical changes, high-resolution XPS spectra of Mn 2p and O 1s were collected before and after cycling for both MO and b-FeMO electrodes. For the MO electrode, post-cycling Mn 2p spectra (Figure 6i) revealed a reduced proportion of Mn3+, while the O 1s spectra (Figure 6j) showed a decrease in Mn–O–H species and an increase in lattice oxygen (Mn–O). This indicates a loss of oxygen vacancies, which are essential for facilitating Zn2+ diffusion and pseudocapacitive behavior. The reduction in Mn3+ and oxygen vacancies correlates with lower ionic conductivity and sluggish charge transfer, explaining the sharp capacity fading (from 85.15 to 16.95 mAh·g−1) over 200 cycles.
In contrast, the b-FeMO electrode (Figure 6k,l) maintained a higher Mn3+ content and a relatively stable Mn–O–H signal in the O 1s region even after 200 cycles. The retention rate of Mn3+ not only reflects reduced Mn dissolution but also implies a greater density of oxygen vacancies, which enhance Zn2+ adsorption and facilitate ion diffusion. Furthermore, the appearance of Zn signals in the post-cycled EDS maps (Figures S6 and S7) confirms the effective and reversible Zn2+ intercalation in b-FeMO. In contrast, MO exhibited stronger Mn dissolution signals and reduced Zn content, supporting the notion that Fe3+ doping mitigates Mn loss and structural degradation.
Quantitative elemental analysis (Table S2) further confirmed that Mn loss in MO was significantly more severe than in b-FeMO. The enhanced retention rate of Mn in b-FeMO implies stronger framework stability, attributed to Fe3+ occupying interstitial or substitutional sites, reinforcing the layered structure and suppressing dissolution pathways.

4. Discussion

The combined CV, GCD, and rate performance analysis firmly supports that Fe3+ doping significantly improves Zn2+ storage kinetics, rate capability, and structural stability of δ-MnO2. Compared to undoped and other doped samples (CuMO, RuMO), the optimized b-FeMO consistently outperforms in all aspects, with advantages that are further confirmed through comparison with recent literature. These findings validate the potential of Fe3+-doped δ-MnO2 as a robust and high-performing cathode material for aqueous ZIBs [20].
The improved rate capability can be explained by the enhanced ion diffusion and electron transport stemming from the increased Mn3+ content and resulting oxygen vacancies. Additionally, the introduction of Fe3+ suppresses Mn dissolution, preserving the layered structure and avoiding pore clogging—a major issue in long-term Zn2+ intercalation/deintercalation.
Collectively, the post-cycling SEM and XPS results provide compelling evidence that Fe3+ doping significantly enhances both the structural and chemical stability of δ-MnO2 during prolonged Zn2+ insertion/extraction. The b-FeMO electrode maintains its hierarchical structure, retains oxygen vacancies, suppresses Mn dissolution, and supports efficient Zn2+ reversibility—factors that are critical for achieving superior cycling performance (41.7% retention rate vs. 19.9% for MO after 200 cycles). These results reinforce the conclusion that Fe3+-doping is an effective and scalable strategy for improving the durability of manganese-based cathodes for aqueous zinc-ion batteries.

5. Conclusions

This study demonstrates that controlled Fe3+ doping significantly enhances the structural stability and electrochemical performance of δ-MnO2 cathodes for aqueous ZIBs. Among the tested dopants, Fe3+ at a Mn:Fe ratio of 20:1 achieved the best balance, yielding a high specific capacity of 116.24 mAh·g−1 and a retention rate of 41.7% after 200 cycles. The Fe3+ doping strategy effectively mitigates Mn dissolution, promotes oxygen vacancy formation, and stabilizes the layered structure. Comparative studies with Cu2+ and Ru3+ indicate that Cu2+ offers moderate improvements, while Ru3+ alters the crystal structure detrimentally. These findings provide a viable strategy for improving Mn-based cathodes for future zinc-ion battery development.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/solids6030045/s1: Figure S1: (a) and (b) SEM images of the CuMO; (c),(d), and (e) EDS mapping of CuMO; Figure S2: (a) and (b) SEM images of the RuMO; (c–f) EDS mapping of RuMO; Figure S3: XPS full spectra of (a) MO; (b) b-FeMO; (c) Cu-MO; and (d) RuMO; Figure S4: Cycling and rate performance of FeMO with different Fe loadings; Figure S5: GCD curves of the b-FeMO at different current rates; Figure S6: SEM and EDS mapping of MO after 200 cycles; Figure S7: SEM and EDS mapping of b-FeMO after 200 cycles; Table S1: XPS elemental analysis of various MO samples; and Table S2: XPS analysis of MO and b-FeMO electrodes before and after cycling.

Author Contributions

Conceptualization, P.W. and Z.C.; methodology, P.W.; validation, P.W., H.Y. and Z.C.; formal analysis, P.W.; investigation, P.W., H.Y., C.Z. and Y.W.; resources, Y.W. and Z.C.; data curation, P.W., C.Z. and Y.W.; writing—original draft preparation, P.W.; writing—review and editing, Z.C.; supervision, Z.C.; project administration, Z.C.; and funding acquisition, Z.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Natural Science Fund from Ningbo Municipal Bureau of Science and Technology (No. 2023J040).

Data Availability Statement

Data will be made available on request.

Acknowledgments

The authors extend their gratitude Rongqi Xia from Scientific Compass (www.shiyanjia.com) for providing invaluable assistance with the XPS analysis.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Cai, Y.; Chua, R.; Srinivasan, M. Anode Materials for Rechargeable Aqueous Al-Ion Batteries: Progress and Prospects. Chemnanomat 2022, 8, 1–17. [Google Scholar] [CrossRef]
  2. Guo, A.; Wang, Z.; Chen, L.; Liu, W.; Zhang, K.; Cao, L.; Liang, B.; Luo, D. A Comprehensive Review of the Mechanism and Modification Strategies of V2O5 Cathodes for Aqueous Zinc-Ion Batteries. ACS Nano 2024, 18, 27261–27286. [Google Scholar] [CrossRef]
  3. Zhang, Y.; Qin, J.; Batmunkh, M.; Li, W.; Fu, H.; Wang, L.; Al-Mamun, M.; Qi, D.; Liu, P.; Zhang, S.; et al. Scalable Spray Drying Production of Amorphous V2O5-EGO 2D Hetero-structured Xerogels for High-Rate and High-Capacity Aqueous Zinc Ion Batteries. Small 2022, 18, 2105761. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, L.; Liao, Y.; Ye, M.; Cai, W.; Xiao, M.; Hu, C.; Zhong, B.; Wan, F.; Guo, X. Regeneration of Spent Lithium Manganate Batteries into Al-Doped MnO2 Cathodes toward Aqueous Zn Batteries. ACS Appl. Mater. Interfaces 2023, 15, 59475–59481. [Google Scholar] [CrossRef]
  5. Qin, L.; Li, S.; Li, L.; Fang, G.; Cheng, H.; Zhu, Q.; Gao, H.; Chen, S. N/Br co-doped C coating Zn2VO4 as excellent electrochemical performance cathode material for aqueous zinc ion batteries. Mater. Lett. 2022, 315, 131949. [Google Scholar] [CrossRef]
  6. Wang, R.; Yao, M.; Huang, S.; Tian, J.; Niu, Z. An anti-freezing and anti-drying multifunctional gel electrolyte for flexible aqueous zinc-ion batteries. Sci. China Mater. 2022, 65, 2189–2196. [Google Scholar] [CrossRef]
  7. Ma, D.; Li, F.; Ouyang, K.; Chen, Q.; Zhao, J.; Chen, M.; Yang, M.; Wang, Y.; Chen, J.; Mi, H.; et al. An electrochemically driven hybrid interphase enabling stable versatile zinc metal electrodes for aqueous zinc batteries. Nat. Commun. 2025, 16, 4817. [Google Scholar] [CrossRef]
  8. Kim, A.; Park, Y.; Choi, J.; Yu, S.-H.; Nam, K.W. A Comprehensive Review of Cathode Materials for Advanced Aqueous Zinc-Ion Batteries. ACS Appl. Energy Mater. 2025, 8, 6806–6828. [Google Scholar] [CrossRef]
  9. Zhong, W.; Tan, C.; Li, L.; Zhang, S.; Wang, X.; Cheng, H.; Lu, Y. Regulation of aqueous electrolyte interface via electrolyte strategies for uniform zinc deposition. Nano Res. 2024, 17, 8678–8693. [Google Scholar] [CrossRef]
  10. Lin, C.; Zhang, H.; Zhang, X.; Liu, Y.; Zhang, Y. Kinetics-Driven MnO2 Nanoflowers Supported by Interconnected Porous Hollow Carbon Spheres for Zinc-Ion Batteries. ACS Appl. Mater. Interfaces 2023, 15, 14388–14398. [Google Scholar] [CrossRef]
  11. Minakshi, M.; Aughterson, R.; Sharma, P.; Sunda, A.P.; Ariga, K.; Shrestha, L.K. Micelle-Assisted Electrodeposition of γ-MnO2 on Lead Anodes: Structural and Electrochemical Insights. ChemNanoMat 2025, 1–11. [Google Scholar] [CrossRef]
  12. Zhang, N.; Cheng, F.; Liu, J.; Wang, L.; Long, X.; Liu, X.; Li, F.; Chen, J. Rechargeable aqueous zinc-manganese dioxide batteries with high energy and power densities. Nat. Commun. 2017, 8, 1–9. [Google Scholar] [CrossRef]
  13. Liu, X.J.; Liu, X.; Bi, Y.F.; Ke, J. Construction of Lithium-Rich Manganese-Based Cathodes Based on Activated Manganese Dioxide for Enhanced Energy Storage Performances. Energy Fuels 2022, 36, 13238–13245. [Google Scholar] [CrossRef]
  14. Zhong, S.; Xin, Y.; Mo, L.; He, B.; Zhang, F.; Zhao, C.; Hu, L.; Tian, H. Intercalation and Interface Engineering of Layered MnO2 Cathodes toward High-Performance Aqueous Zinc-Ion Batteries. J. Phys. Chem. C 2025, 129, 6684–6696. [Google Scholar] [CrossRef]
  15. Khamsanga, S.; Pornprasertsuk, R.; Yonezawa, T.; Mohamad, A.A.; Kheawhom, S. δ-MnO2 nanoflower/graphite cathode for rechargeable aqueous zinc ion batteries. Sci. Rep. 2019, 9, 8441. [Google Scholar] [CrossRef]
  16. Zhang, B.; Dong, P.; Yuan, S.; Zhang, Y.; Zhang, Y.; Wang, Y. Manganese-Based Oxide Cathode Materials for Aqueous Zinc-Ion Batteries: Materials, Mechanism, Challenges, and Strategies. Chem Bio Eng. 2024, 1, 113–132. [Google Scholar] [CrossRef]
  17. Zhai, X.; Yu, Y.; Hu, Y. Engineering Aqueous Zn-MnO2 Microbatteries Using a Synergistic Reaction Mechanism. ACS Appl. Energy Mater. 2023, 6, 6171–6182. [Google Scholar] [CrossRef]
  18. Li, W.; Qin, L.; Liu, Z.; Li, L.; Li, W.; Fang, G. Potassium-Ion-Doped Manganese Oxides and Kaolinite Electrolyte Additives for Aqueous Zinc-Ion Batteries. ACS Appl. Nano Mater. 2024, 7, 9720–9729. [Google Scholar] [CrossRef]
  19. Kumari, P.; Kundu, R. Zinc-Ion Batteries: Promise and Challenges for Exploring the Post-Lithium Battery Materials. ACS Appl. Energy Mater. 2024, 7, 9634–9669. [Google Scholar] [CrossRef]
  20. Jiang, S.; Tian, S.; Zhang, S.; Fang, L.; Wang, Z.; Nie, P.; Han, W.; Xue, X.; Zhao, C.; Lu, M.; et al. Iron-Doped Nanorods of MnO2 for Applications in Zinc-Ion Batteries. ACS Appl. Nano Mater. 2024, 7, 27648–27655. [Google Scholar] [CrossRef]
  21. Ding, X.; Wen, Y.; Qing, C.; Wei, Y.; Wang, P.; Liu, J.; Peng, Z.; Song, Y.; Chen, H.; Rong, Q. Cr-induced enhancement of structural stability in δ-MnO2 for aqueous Zn-ion batteries. J. Alloys Compd. 2024, 986, 174041. [Google Scholar] [CrossRef]
  22. Liu, Z.; Liu, Y.; Zhang, Y.; Liu, X.; Yan, D.; Huang, J.; Peng, S. Selection of Cu2+ intercalation from electronegativity perspective: Improving cycle stability and rate performance of δ-MnO2 cathode material for aqueous zinc-ion batteries. Sci. China Mater. 2022, 66, 531–540. [Google Scholar] [CrossRef]
  23. Zheng, J.; Xia, R.; Yaqoob, N.; Kaghazchi, P.; Elshof, J.E.T.; Huijben, M. Simultaneous Enhancement of Lithium Transfer Kinetics and Structural Stability in Dual-Phase TiO2 Elec trodes by Ruthenium Doping. ACS Appl. Mater. 2024, 16, 8616–8626. [Google Scholar] [CrossRef]
  24. Hashemzadeh, F.; Kashani Motlagh, M.M.; Maghsoudipour, A. A Comparative Study of Hydrothermal and Sol-Gel Methods in the Synthesis of MnO2 Nanostructures. J. Sol-Gel Sci. Technol. 2009, 51, 169–174. [Google Scholar] [CrossRef]
  25. Wang, J.; Sun, X.; Zhao, H.; Xu, L.; Xia, J.; Luo, M.; Yang, Y.; Du, Y. Superior-Performance Aqueous Zinc Ion Battery Based on Structural Transformation of MnO2 by Rare Earth Doping. J. Phys. Chem. C 2019, 123, 22735–22741. [Google Scholar] [CrossRef]
  26. 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]
  27. Wang, R.; Chen, W.; Zhang, C.; Zhao, R.; Wang, X. Electrochemically Active Mn2+ Enabling High-Performance Aqueous Zinc Ion Batteries. Energy Fuels 2024, 38, 13436–13443. [Google Scholar] [CrossRef]
  28. Fang, Y.; Xie, X.; Zhang, B.; Chai, Y.; Lu, B.; Liu, M.; Zhou, J.; Liang, S. Regulating Zinc Deposition Behaviors by the Conditioner of PAN Separator for Zinc-Ion Batteries. Adv. Funct. Mater. 2021, 32, 2109671. [Google Scholar] [CrossRef]
  29. Kamenskii, M.A.; Popov, A.Y.; Eliseeva, S.N.; Kondratiev, V.V. The Effect of the Synthesis Method of the Layered Manganese Dioxide on the Properties of Cathode Materials for Aqueous Zinc-Ion Batteries. Russ. J. Electrochem. 2023, 59, 1092–1101. [Google Scholar] [CrossRef]
  30. Feng, Y.P.; Zhang, Y.F.; Zhang, Y.C. In situ growth of the δ-manganese dioxide on carbon cloth by different concen-trations of reactants for eco-friendly battery applications. J. Solid State Electrochem. 2023, 27, 2691–2700. [Google Scholar] [CrossRef]
  31. Ni, Z.; Liang, X.; Zhao, L.; Zhao, H.; Ge, B.; Li, W. Tin doping manganese dioxide cathode materials with the improved stability for aqueous zinc-ion batteries. Mater. Chem. Phys. 2022, 287, 126238. [Google Scholar] [CrossRef]
  32. Xu, J.; Hu, X.; Alam, A.; Muhammad, G.; Lv, Y.; Wang, M.; Zhu, C.; Xiong, W. Al-doped α-MnO2 coated by lignin for high-performance rechargeable aqueous zinc-ion batteries. RSC Adv. 2021, 11, 35280–35286. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD patterns of (a) pristine MO and Fe-doped MOs; (b) Cu and Ru doped MOs.
Figure 1. XRD patterns of (a) pristine MO and Fe-doped MOs; (b) Cu and Ru doped MOs.
Solids 06 00045 g001
Figure 2. SEM images of (a,b) pristine MO; (c,d) Fe-doped MO, and (e) EDS mapping of Fe-doped b-FeMO: (f) O; (g) Mn; and (h) Fe.
Figure 2. SEM images of (a,b) pristine MO; (c,d) Fe-doped MO, and (e) EDS mapping of Fe-doped b-FeMO: (f) O; (g) Mn; and (h) Fe.
Solids 06 00045 g002
Figure 3. High-resolution of XPS of (a) Mn 2p for pristine MO, b-FeMO, CuMO, and RuMO; (b) O 1s for pristine MO, b-FeMO, CuMO, and RuMO. Note: the solid green lines are the baselines during fitting.
Figure 3. High-resolution of XPS of (a) Mn 2p for pristine MO, b-FeMO, CuMO, and RuMO; (b) O 1s for pristine MO, b-FeMO, CuMO, and RuMO. Note: the solid green lines are the baselines during fitting.
Solids 06 00045 g003
Figure 4. CV curves of (a) MO, (b) b-FeMO, (c) CuMO, and (d) RuMO.
Figure 4. CV curves of (a) MO, (b) b-FeMO, (c) CuMO, and (d) RuMO.
Solids 06 00045 g004
Figure 5. Cycling and rate performance of (a) MO, b-FeMO s; (b) MO, b-FeMOs.
Figure 5. Cycling and rate performance of (a) MO, b-FeMO s; (b) MO, b-FeMOs.
Solids 06 00045 g005
Figure 6. Post-cycling SEM images of (a,b) pristine MO electrode before cycling; (c,d) pristine MO electrode after 200 cycles; (e,f) b-FeMO electrode before cycling; (g,h) b-FeMO electrode after 200 cycles; and XPS curves of (i,j) Mn 2p and O 1s for MO before and after cycling; and (k,l) Mn 2p and O 1s for b-FeMO before and after cycling.
Figure 6. Post-cycling SEM images of (a,b) pristine MO electrode before cycling; (c,d) pristine MO electrode after 200 cycles; (e,f) b-FeMO electrode before cycling; (g,h) b-FeMO electrode after 200 cycles; and XPS curves of (i,j) Mn 2p and O 1s for MO before and after cycling; and (k,l) Mn 2p and O 1s for b-FeMO before and after cycling.
Solids 06 00045 g006
Table 1. Comparative electrochemical performance of materials.
Table 1. Comparative electrochemical performance of materials.
MaterialMOb-FeMOCuMORuMO
Initial discharge specific capacity
(mAh/g)
85.15116.24146.53210.57
Discharge specific capacity after 200 cycles (mAh/g)16.9548.4751.6744.58
Capacity retention rate (%)19.941.7035.2321.17
Average Coulombic efficiency (%)99.7210099.5999.42
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

Wang, P.; Yu, H.; Zou, C.; Wu, Y.; Chen, Z. Enhancing the Structural Stability and Electrochemical Performance of δ-MnO2 Cathodes via Fe3+ Doping for Aqueous Zinc-Ion Batteries. Solids 2025, 6, 45. https://doi.org/10.3390/solids6030045

AMA Style

Wang P, Yu H, Zou C, Wu Y, Chen Z. Enhancing the Structural Stability and Electrochemical Performance of δ-MnO2 Cathodes via Fe3+ Doping for Aqueous Zinc-Ion Batteries. Solids. 2025; 6(3):45. https://doi.org/10.3390/solids6030045

Chicago/Turabian Style

Wang, Pengfei, Haiyang Yu, Chengyan Zou, Yuxue Wu, and Zhengfei Chen. 2025. "Enhancing the Structural Stability and Electrochemical Performance of δ-MnO2 Cathodes via Fe3+ Doping for Aqueous Zinc-Ion Batteries" Solids 6, no. 3: 45. https://doi.org/10.3390/solids6030045

APA Style

Wang, P., Yu, H., Zou, C., Wu, Y., & Chen, Z. (2025). Enhancing the Structural Stability and Electrochemical Performance of δ-MnO2 Cathodes via Fe3+ Doping for Aqueous Zinc-Ion Batteries. Solids, 6(3), 45. https://doi.org/10.3390/solids6030045

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop