Next Article in Journal
Facile Synthesis of Non-Noble CuFeCo/C Catalysts with High Stability for ORR in PEMFC
Next Article in Special Issue
Alloying Design Strategies for High-Performance Zn Anodes in Aqueous Zinc-Ion Batteries
Previous Article in Journal
Microscopic Mechanisms and Pavement Performance of Waterborne Epoxy Resin-Modified Emulsified Asphalt
Previous Article in Special Issue
Preparation of Metallic Zr from ZrO2 via Carbothermal and Electrochemical Reduction in Molten Salts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Suppressing Jahn–Teller Distortion in Manganese Oxides for High-Performance Aqueous Zinc-Ion Batteries

1
Institute for Advanced Interdisciplinary Research (iAIR), School of Chemistry and Chemical Engineering, University of Jinan, Jinan 250022, China
2
School of Materials and Chemical Engineering, Xuzhou University of Technology, Xuzhou 221018, China
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(12), 2817; https://doi.org/10.3390/ma18122817
Submission received: 10 May 2025 / Revised: 4 June 2025 / Accepted: 10 June 2025 / Published: 16 June 2025

Abstract

Manganese oxides (MnOx) have been confirmed as the most promising candidates for aqueous zinc-ion batteries (AZIBs) due to their cost-effectiveness, high theoretical capacity, high voltage platforms, and environmental friendliness. However, in practical applications, AZIBs are hindered by the Jahn–Teller distortion (JTD) effect, primarily induced by Mn3+ (t2g3eg1) in octahedral coordination, which leads to severe structural deformation, rapid capacity fading, and poor cycling stability. This review systematically outlines the fundamental mechanisms of JTD in MnOx cathodes, including electronic structure changes, lattice distortions, and their side effects on Zn2+ storage performance. Furthermore, we critically discuss advanced strategies to suppress JTD, such as cation/anion doping, interlayer engineering, surface/interface modification, and electrolyte optimization, aimed at enhancing both structural stability and electrochemical performance. Finally, we propose future research directions, such as in situ characterization, machine learning-guided material design, and multifunctional interfacial engineering, to guide the design of high-performance MnOx hosts for next-generation AZIBs. This review may provide a promising guideline for overcoming JTD challenges and advancing MnOx-based energy storage systems.

1. Introduction

The growing demand for large-scale and sustainable energy storage systems has driven extensive research into safe, cost-effective, and environmentally friendly battery systems. Currently, lithium-ion batteries (LIBs) dominate the market due to their long cycle life, high energy density, and superior specific capacity, making them the preferred choice for large-scale energy storage and electric vehicles [1,2,3,4]. However, their widespread adoption faces significant challenges, including the scarcity of lithium resources and safety concerns associated with flammable organic electrolytes. Therefore, new energy storage systems are urgently needed to meet the demand for sustainable and eco-friendly energy systems.
Among these, aqueous zinc-ion batteries (AZIBs) have emerged as a promising candidate for LIBs due to their high theoretical capacity (820 mAh g−1 and 5855 mAh cm−3), high energy density, low cost, eco-friendliness, and ease of recycling [5,6,7,8]. Among various cathode materials, manganese oxides (MnOx) have emerged as one of the most promising candidates owing to their cost-effectiveness, high theoretical capacity, multiple oxidation states, environmental benignity, and abundant natural reserves [9,10,11,12]. However, in practical application, MnOx cathodes in AZIBs are significantly hampered by the Jahn–Teller distortion (JTD) effect, which originates from the electronic configuration of Mn3+ (t2g3eg1) in octahedral coordination, resulting in elongation or compression of Mn-O bonds, and thus, inducing asymmetric lattice distortions and destabilizing the host structure [13,14,15,16]. Although Mn (IV) can undergo multi-electron transfer processes to deliver higher capacity, during the discharge process, partial reduction of Mn (IV) to Mn (III) occurs alongside disproportionation reactions (2Mn3+ → Mn2+ + Mn4+) and exacerbate the capacity fading. When MnO2 serves as the cathode material, the dissolution of Mn2+ induces structural collapse in the crystal framework, resulting in diminished ion storage sites. Concurrently, Mn2+ dissolution reduces active material content, thereby compromising the cathode’s reversible capacity. Therefore, the JTD effect will trigger a severe structural collapse, compromised cyclability, and deteriorate Zn2+ diffusion kinetics [17,18,19]. Nowadays, to address the challenges, extensive efforts have been devoted to suppressing JTD through strategies such as heteroatom doping (e.g., cation, anion, or oxygen vacancies), interlayer engineering, surface modification, and electrolyte optimization [20,21,22,23,24,25]. These approaches aim to stabilize the MnOx framework, enhance intrinsic conductivity and charge/mass transport efficiency, and improve Zn2+ storage reversibility.
In this review, we systematically elucidated the mechanism of Jahn–Teller distortion in MnO2, influencing factors, and demonstrated adverse effects on the structure while analyzing how these structural changes affect electrochemical performance. Then, we introduced current common strategies for suppressing Jahn–Teller distortion and finally outlined future directions for mitigating this phenomenon.

2. Fundamentals of Jahn–Teller Effect in MnOx

Jahn–Teller distortion is an important phenomenon in transition metal oxides, particularly in MnO2-based materials, where structural deformation arises from the inherent electronic instability of specific transition metal ions and their tendency to minimize energy in octahedral coordination environments [26,27,28,29,30]. As a cornerstone of crystal field theory, this effect explains how Mn3+-containing octahedral complexes (t2g3eg1 configuration) undergo spontaneous symmetry breaking through tetragonal distortion to achieve system stabilization. The distortion originates from the degeneracy of eg orbitals (dz2 and dx2 − y2) in high-spin Mn3+ (d4) ions, where asymmetric electron occupation triggers lattice deformation to lift orbital degeneracy and lowers the overall energy state [1,31,32,33,34].
In MnO2 cathodes for aqueous zinc-ion batteries, this effect manifests through three critical aspects. Firstly, the electronic structure, the labile eg electron in Mn3+, induces unequal Mn-O bond lengths [35], such as elongation and/or compression. Secondly, structural consequences trigger irreversible phase transitions (layered → spinel) and MnO6 octahedra tilting [36]. Ultimately, electrochemical impact accelerates Mn dissolution via disproportionation and deteriorates Zn2+ diffusion [37]. The aforementioned aspects will be discussed in detail below.

2.1. Electronic and Structural Basis

In manganese dioxide (MnO2), the Jahn–Teller (JT) effect plays a pivotal role in determining the electronic and structural properties, which arises from the electronic instability of Mn3+ (3d4) ions in octahedral coordination, and thus significantly influencing their electrochemical behaviors [38]. Universally, the valence state of Mn is +4 in MnO2. However, during battery operation, the cathode gains electrons, and the valence of Mn is reduced from Mn4+ to Mn3+. For Mn3+ (3d4), under the influence of an octahedral crystal field with sixfold coordination, the 3d orbitals split into two energy levels [39,40]: t2g (triply degenerate, lower energy) and eg (doubly degenerate, higher energy). According to the Aufbau principle, the four 3d electrons preferentially occupy the lower-energy t2g orbitals before filling the higher-energy eg orbitals [41,42]. As illustrated in Figure 1 [43], three electrons occupy the t2g orbitals, while the remaining electron resides in the eg orbitals. The eg orbitals (comprising dz2 and dx2 − y2) host only one electron, creating an asymmetric electron distribution [44,45]. This imbalance induces directional shielding effects: the dz2 orbital provides stronger shielding along the axial (z-axis) direction. The dx2 − y2 orbital exhibits reduced electron shielding in the equatorial plane compared to the dz2 orbital, which would create the asymmetry of the electron and establishes unequal electrostatic repulsion between Mn3+ ions and the surrounding oxygen atoms [46,47]. Therefore, this imbalance drives the characteristic Jahn–Teller distortion, where the MnO6 octahedron undergoes either axial elongation or compression to achieve a lower-energy configuration. During the electrochemical reduction of MnO2, the conversion from Mn4+ to Mn3+ promotes this distortion process, then elongates the Mn-O bonds along the y-axis direction [48,49]. The resulting structural deformation generates substantial localized lattice strain that accumulates with cycling, eventually leading to the degradation as follows: a microstructural collapse that exposes fresh active surfaces to the electrolyte, initiation of the Mn3+ disproportionation reaction (2Mn3+ → Mn2+ + Mn4+ [50]) at the electrode–electrolyte interface, and irreversible phase transformation to the thermodynamically stable spinel-type MnO2. Finally, the capacity will fade due to both active material loss and increased charge transfer resistance, originating from the dissolved Mn2+ species migrating into the electrolyte solution and structural reorganization of the remaining Mn4+ [51].
This comprehensive degradation mechanism highlights the critical interplay between electronic structure (orbital asymmetry), atomic-scale structural changes (Jahn–Teller distortion), and macroscopic electrochemical performance in MnOx cathodes. The understanding of these coupled phenomena provides essential insights for developing stabilization strategies targeting each degradation pathway.

2.2. Impact of Jahn–Teller Effect on Electrochemical Implications in AZIBs

Based on this fundamental understanding, we will investigate the correlation between Jahn–Teller distortion effects and the electrochemical performance of MnO2 cathodes in AZIBs [13,14,52]. As one of the most promising host materials for AZIBs, MnO2’s intrinsic properties govern the performance by the following key aspects such as the specific capacity, determined by accessible redox-active sites and their utilization efficiency, cycling stability, dictated by structural integrity, and the rate capability, influenced by ion transport kinetics [17,53,54,55].
The charge-storage mechanism in MnO2 mainly relies on the reversible insertion/extraction of Zn2+/H+ ions [56]. However, during the reversible process, strong electrostatic interactions with the host lattice will be created during migration due to the substantial ionic radius and high charge density of Zn2+. These interactions aggravated the inherent Jahn–Teller distortion of Mn3+-containing octahedra, with each electrochemical cycle progressively accumulating lattice strain. This cyclic distortion ultimately leads to manganese dissolution via disproportionation and structural collapse through bond-length fatigue, both contributing to capacity fading [57,58,59,60]. The specific capacity, cycling performance, and dissolved Mn2+ comparison of the modified MnO2 cathode materials of aqueous Zn-ions batteries are summarized in Table 1.
Jahn–Teller effect impacts capacity through the following interrelated mechanisms [68]. First, manganese dissolution reduces the quantity of electroactive material [69]. Then, lattice distortion decreases the accessibility of remaining Zn2+ storage sites by modifying coordination environments [70,71]. Finally, the induced strain field disrupts ion migration pathways, increasing the activation energy for Zn2+ diffusion in severely distorted regions, which can be confirmed by recent density functional theory calculations [72].
In conclusion, the structural evolution studies reveal that the Jahn–Teller effect drives irreversible phase transitions in MnO2 polymorphs, which reduce both Zn2+ storage sites and diffusion coefficients by an order of magnitude, consistent with the phenomenon observed experimentally. The cumulative Jahn–Teller distortion during cycling initiates a cascade of degradation processes in MnO2 cathodes, including local lattice strain accumulation, manganese dissolution, structural collapse, and detrimental phase transitions. These interrelated phenomena collectively limit the cycle life and restrict the practical capacities of MnO2 in most reported systems. Currently, the mitigation strategies mainly focus on orbital engineering through transition metal doping and strain–relief structural designs, which will be discussed in subsequent sections.

3. Mitigation Strategies to Suppress the Jahn–Teller Effect

To overcome these limitations, strategic modification of MnO2 has become imperative, with the central objective of stabilizing the crystal lattice against Jahn–Teller-induced deformations.
Figure 2 shows the currently reported modification strategies to suppress the Jahn–Teller effect mentioned. It focuses on the following fundamental aspects of the distortion mechanisms: strengthening the main structure through cationic/anionic doping [73], electronic structure modulation to reduce eg orbital degeneracy, and interface stabilization via surface engineering [74]. These strategies simultaneously address the root causes of Jahn–Teller effects and improve zinc-ion diffusion kinetics and charge transfer efficiency. Notably, recent advances also combine the above approaches with electrolyte optimization to promote synergistic stabilization effects [75]. This multi-dimensional modification paradigm will demonstrate promising improvement strategies toward advanced MnO2-based materials.

3.1. Strategic Cationic Doping for Structural Stabilization

Nowadays, to suppress the Jahn–Teller distortions and enhance the structural stability of the MnO2 based materials, cation doping has been confirmed as one of the most effective strategies, which fundamentally operates through the strategic substitution of Mn4+ in MnO6 octahedra with foreign cations, thus simultaneously modulates bond strength through altered orbital hybridization, optimizes the local crystal field environment, and regulates valence electron distribution to effectively suppress Jahn–Teller distortions. For example, Zhao et al. [61] employed electrochemical oxidation to synthesize Al-doped δ-MnO2 (AlxMnO2, x ≈ 0.1), where controlled Al3+ substitution created beneficial cation vacancies (Figure 3b) that enabled three-dimensional Zn2+ diffusion with reduced energy barriers while providing additional redox-active sites. In addition, the stabilization mechanism is confirmed through the comprehensive characterizations. First, inductively coupled plasma (ICP) analysis demonstrated a significant reduction in Mn dissolution. Second, density functional theory (DFT) calculations revealed a dramatic increase in Mn dissolution energy, indicating enhanced lattice cohesion. Third, the variation in Mn-O bond length during Zn2+ insertion was markedly suppressed, with the decrease in bond length, ultimately delivering exceptional electrochemical performance (327.9 mAh g−1 at 1 A g−1 with 92.4% capacity retention over 1000 cycles) (Figure 3c). In addition to Al3+ doping discussed in the aforementioned work, other cation dopants such as Ni also exhibit inhibitory effects on Jahn–Teller distortion. Liang et al. [79] prepared Ni-doped α-MnO2 (Ni-MnO2) as shown in Figure 3d. Compared to Mn4+, Ni2+ has a smaller atomic radius; its occupation in the lattice causes the surrounding O atoms to be arranged more closely, thereby Ni2+ doping resulting in shorter and more uniform chemical bond lengths compared to Mn-O bonds at equivalent positions. The electron localization function (ELF) diagram shown in Figure 3e also shows that the electron localization around the Ni2+ is enhanced compared to the Mn ion at the same position, indicating that the Ni-O has a stronger interaction and the material’s stability is improved. The analysis of the results is shown in Figure 3f. Ni-MnO2 has extremely excellent cycle stability. This optimized bond length effectively suppressed Jahn–Teller distortion, significantly enhancing the material’s cycling performance. Collectively, these findings validate the effectiveness of cationic doping in mitigating Jahn–Teller distortions while simultaneously enhancing both the capacity and cycling stability in AZIBs.

3.2. Anionic Doping for Enhanced Structural Resilience

The incorporation of anionic species at the O sites of MnO6, attributed to the varying covalency of Mn-X interactions, where X represents an anion such as F, S2−, and PO43−, has also been confirmed as one of the most promising approaches to suppress the JTD and improve the electrochemical properties. By introducing these species, the distortion of MnO6 octahedra can be effectively suppressed. Thus, Ye et al. [63] explored the anion species of Se to mitigate Jahn–Teller distortions of MnO2 (Figure 4a). The mechanism of doping effect and inhibition of Jahn–Teller distortions are evidenced by comprehensive experimental and computational analyses. As shown in Figure 4e, the charge density differentials reveal a substantial weakening of Zn2+ host electrostatic coupling in Se-MnO2, directly correlating with reduced lattice strain and suppressed octahedral distortion. The stabilization effects are quantitatively demonstrated through comparative bond length analysis (Figure 4b), confirming effective suppression of both bond elongation and manganese dissolution, which are the hallmark manifestations of Jahn–Teller distortion. The stabilization mechanism of the doping effect to suppress the Jahn–Teller is primarily reflected through the following aspects: optimized ion adsorption energetics (Figure 4d) and inhibited H+ co-intercalation that typically exacerbates structural degradation. These synergistic improvements enhanced the electrochemical performance of Se-MnO2, which maintains 98.5% capacity retention over 1000 cycles and delivers 325 mAh g−1 reversible capacity, confirming that anionic doping is a potent strategy for developing stable MnO2 cathodes. Similar to the Se element doping, other anion dopants also substitute oxygen positions in the original [MnO6] octahedra. Zhao et al. [80] synthesized S-MnO2 (Figure 4f). The roles of these anion dopants in cathode materials are analogous: they regulate the electronic structure of MnO2 and weaken the electrostatic interactions between Zn2+ and the host material. As shown in Figure 4g, S-MnO2 has a lower diffusion energy barrier of Zn2+, and the charge density difference distribution in Figure 4h indicates the same result. In sharp contrast, the insertion of Zn in S-MnO2 possessed a smaller charge depletion area, indicating less electronic interaction between the inserted Zn2+ and S-MnO2. The capacity fading caused by Jahn–Teller distortion is mitigated (Figure 4i). The introduction of anion dopants primarily modulates the electronic structure while structurally exerting dual effects: weakening Zn2+-host electrostatic interactions and strengthening Mn-O bonds. These synergistic actions collectively suppress Jahn–Teller distortion and enhance cycling stability.

3.3. Interlayer Engineering for Structural Stabilization

The strong electrostatic interactions between Zn2+ ions and the host lattice interlayers serve as a primary driver of Jahn–Teller distortion. Therefore, it is crucial to make efforts to mitigate these interactions to regulate the structural stability of the MnO2 host. Interlayer engineering has emerged as a multifaceted strategy to counteract Jahn–Teller distortions in MnO2 cathodes through simultaneous geometric confinement and electronic structure modulation. Among the various modulation strategies, the pre-intercalation of pillaring structures between the layers is an effective approach to simultaneously stabilize the interlayer framework and expand the interlayer spacing. Zhang et al. [76] demonstrated this concept by introducing Cu2+ pillars (CuMO) into the MnO2 structure, as shown in Figure 5a, achieving an expanded interlayer spacing of 0.7 nm. This structural modification yielded remarkable electrochemical performance, with CuMO delivering enhanced specific capacity and maintaining stable cycling over 700 cycles at 5 A g−1 in Figure 5b. The reduced Zn2+ insertion energy in CuMO (Figure 5c) directly confirms the weakened electrostatic interactions between the host structure and Zn2+ ions, facilitating easier ion intercalation. Additionally, the charge-discharge mechanism of CuMO (Figure 5d) reveals the mechanism of the Cu2+ pillars maintaining structural integrity during the electrochemical processes according to the stable lattice spacing observed before and after cycling. This dual functionality of Cu2+ pillars—both expanding the interlayer space and providing robust structural support—ensures the preservation of crystal structure integrity while simultaneously mitigating the detrimental effects of Jahn–Teller distortion. In addition to ion pillars, other species can also be confirmed as the ideal interlayer structures, such as Na+ [81], K+ [82], and PVA [83], et al. have also been demonstrated as effective interlayer engineering strategies for MnO2.
Beyond inorganic atoms, organic molecules also serve as effective interlayer pillars. PVP-pillared δ-MnO2 synthesized by Zhang et al. [84] in Figure 5e,g indicates PVP expanded the interlayer spacing and weakened the electrostatic interactions between Zn2+ and the host structure. Figure 5g shows a conspicuous terrace-shaped superlattice structure. In MnO2/MXene, the superlattice significantly reduced ion diffusion barriers, and the diagram of the PVP-MnO2 hybrid superlattice in Figure 5h shows the increased electron entropy triggers more t2g–eg transitions for better reactivity, while the selective proton Grotthuss intercalation behavior is thus stemmed from the hybrid superlattice structure and optimized charge distribution. This inhibition of Zn2+ insertion provided high specific capacity (Figure 5i) and improved the rate ability of the material (Figure 5j).

3.4. Hierarchical Surface–Interface Stabilization Strategies

In practical applications, the MnO2 cathodes demand an integrated stabilization approach coordinating atomic-scale electronic modulation with macroscopic structural engineering to address the multiscale challenge of Jahn–Teller distortion in AZIBs. At the atomic level, surface doping induces orbital hybridization, which will reduce the eg orbital splitting energy and prevent localized strain accumulation during cycling. Therefore, surface coating is one of the effective measures to simultaneously enhance the intrinsic conductivity and suppress the volume expansion/Mn dissolution to mitigate structural collapse caused by Jahn–Teller effects. Carbon-based materials have emerged as the most promising candidates for this application due to their exceptional electrical conductivity, remarkable structural versatility, and excellent cost-performance ratio.
Wu et al. [62] developed a graphene-coated MnO2 nanowire composite (MGS) through a controlled hydrothermal synthesis process (Figure 6a). The conformal graphene coating serves as both an efficient electron conductor and structural stabilizer. This core-shell architecture demonstrates excellent rate capability (89% capacity retention at 3 A g−1, Figure 6c), as well as outstanding cycling stability (92% capacity retention over 3000 cycles at 3 A g−1, Figure 6d).
The kinetic process was analyzed with galvanostatic intermittent titration technique (GITT), revealing a two-stage Zn2+ insertion mechanism and Figure 6b plays the two-step intercalation mechanism of MGS cathode: an initial rapid intercalation phase (DP I) with high diffusion coefficient at the MnO2/graphene interface, and a slower diffusion-limited phase (DP II) with Zn2+ penetrating into the 2 × 2 tunnels of α-MnO2. The kinetic transition at 1.30 V causes reactant accumulation, with DP II producing substantially greater lattice strain that typically triggers structural damage and Mn dissolution in bare MnO2. However, the graphene coating in MGS effectively suppresses these degradation pathways, as evidenced by a 78% reduction in Mn2+ dissolution compared to uncoated MnO2 nanowires.
Jiang et al. [85] prepared a β-MnO2 material coated with thin graphite films via a P-milling strategy, as shown in Figure 6e. The pores within β-MnO2 facilitated electrolyte infiltration, while the graphite-integrated structure significantly enhanced the material’s conductivity and suppressed Mn dissolution. Consequently, the β-MnO2@C cathode exhibited exceptional rate capability and cycling stability (Figure 6f). As illustrated in the discharge mechanism schematic (Figure 6g), during discharge, zinc ions precipitate within the pores between the graphite layers and β-MnO2. The graphite coating preserves the structural integrity of β-MnO2, maintains its ion storage sites, and ensures high capacity retention in β-MnO2@C (Figure 6h).
It suggested that the comprehensive protective mechanism, combining strain redistribution, electronic connectivity maintenance, and dissolution suppression, shows a generalizable modification approach, which can be successfully extended to other coating materials, including conductive polymers (e.g., PEDOT) and metal oxides (e.g., Al2O3). Therefore, the surface engineering strategies demonstrate the validity of stabilizing Jahn–Teller-active electrode materials in rechargeable battery systems.

3.5. Electrolyte Optimization in Suppressing Jahn–Teller Effect

In addition to the electron and structural factors, electrolyte engineering has also emerged as one of the effective strategies for mitigating Jahn–Teller distortions in MnO2 cathodes through interfacial chemistry control and bulk electrolyte modulation. Recent advances demonstrate that it can effectively stabilize the Mn3+ state through tailoring electrolyte formulations by modifying the solvation structure of Zn2+ ions and creating protective interfacial layers.
Acetate-based electrolytes, for instance, exhibit unique advantages through the formation of adsorption-induced passivation layer on MnO2 surfaces. For example, Zeng et al. [86] study the MnO2 charge/discharge mechanism (Figure 7a). The reaction pathway shifts from single-electron to two-electron transfer by replacing conventional sulfate anions with acetate groups, enabling direct Mn4+ → Mn2+ conversion and, thus, suppressing Jahn–Teller-active Mn3+ intermediates (Figure 7c). This change is mainly attributed to the modification effect of the acetates, as shown below. First, the H2O* adsorption energy is weakened (Figure 7b) while the Mn2+ discharge product is stabilized. Then, the transformed reaction kinetics yield exceptional electrochemical performance, including remarkable rate capability (Figure 7d) (70 mA cm−2 with minimal capacity loss) and ultra-stable cycling over 4000 cycles. The acetate electrolyte simultaneously addresses the challenges of eliminating Jahn–Teller distortion, suppressing manganese dissolution, and enhancing proton transfer kinetics, providing a new paradigm in aqueous zinc battery design.
Beyond suppressing Mn dissolution by adding Mn2+-containing salts, high-concentration electrolytes also effectively mitigate Mn dissolution. Huang et al. [87] developed a 46.5 M NH − Ac − NH3 − Zn(Ac)2 − Mn(Ac)2 buffered solution (HCDCE electrolyte) for aqueous zinc-ion batteries. Figure 7e illustrates the mechanism of HCDCE: the complexation between NH4Ac/NH3 and Zn2+ weakens Zn2+ solvation, reducing structural damage from electrostatic interactions during Zn2+ insertion/extraction. Additionally, the NH4+−NH3 buffer pair maintains stable electrolyte pH (Figure 7f,g), minimizing by-product formation and preserving electrochemical activity. This endows the material with exceptional cycling stability (4500 cycles with near-100% capacity retention, Figure 7h) and rate capability (Figure 7i).
Modification of electrolytes is achieved through three effective approaches: applying Le Chatelier’s principle by adding homologous salt ions to suppress Mn dissolution, utilizing high-concentration salt solutions to reduce Zn2+ solvation, and employing ionic buffer pairs to maintain pH stability. This strategy simultaneously addresses the challenges of eliminating Jahn–Teller distortion, suppressing manganese dissolution, and enhancing proton transfer kinetics, providing a new paradigm in aqueous zinc battery design.

4. Summary and Perspectives

In this review, we systematically studied the fundamental mechanisms and external triggers of Jahn–Teller distortions, analyzed the unfavorable effects on MnO2 cathodes, and provided targeted mitigation strategies. In AZIBs, while the MnO2-based host has achieved remarkable progress, the critical challenges for commercial implementation still need to be resolved. Then, we would suggest future prioritized research as the following directions:
First, it should deeply investigate the correlation between Jahn–Teller effects and charge storage mechanisms. Currently, the understanding remains incomplete due to the complex, condition-dependent reaction pathways observed in MnO2 cathodes. Therefore, the advanced in situ characterization platforms combining synchrotron X-ray diffraction, electrochemical mass spectrometry, and atomic force microscopy could provide unprecedented insights into real-time structural evolution and ion transport dynamics.
Second, we should focus on the design of comprehensive regulation strategies. Although Jahn–Teller distortion can arise from multiple factors, its fundamental origin lies in the intrinsic electronic structure of MnO2. Thus, cationic/anionic doping strategies that modulate electronic configurations play pivotal roles in suppressing Jahn–Teller distortion. Notably, the electrostatic interactions between Zn2+ and the host framework, coupled with structural degradation induced by water molecule intrusion, exacerbate Jahn–Teller effects. Consequently, integrating ion doping with interfacial engineering (e.g., pillar interlayer design) and electrolyte optimization to achieve multi-pronged suppression of Jahn–Teller distortion is critical for advancing MnO2-based electrodes. Firstly, synergistic doping-pillaring simultaneously tunes electronic structures and mitigates Zn2+-framework electrostatic forces. Secondly, doping-electrolyte coupling modifies charge/discharge mechanisms while optimizing electron delocalization.
Thirdly, research on by-products, intermediate products, and electrolyte additives should be strengthened. During the charge/discharge process, MnO2 cathode surfaces tend to generate by-products or intermediate products, which can affect battery performance and even alter the charge/discharge mechanisms. Currently, there is a lack of systematic studies on the types, structures, and influencing factors of these intermediate products. Electrolyte additives can effectively suppress surface-side reactions and stabilize material structures during charge/discharge cycles. While the effect of a single additive is limited, effective synergistic combinations of multiple additives may yield better results.
Fourth, Jahn–Teller distortion plays a critical role in battery systems utilizing MnO2-type octahedral frameworks as electrode materials, owing to structural deformation in transition metal complexes with octahedral coordination under specific electronic states to achieve energy minimization. This phenomenon manifests broadly across various battery systems, necessitating strategies to enhance the structural stability of manganese-based materials. Representative examples include suppression of Jahn–Teller distortion in LiMn2O4 spinel cathodes for Li-ion batteries, structural stabilization of layered NaMnO2 oxides for Na-ion batteries, and optimization of MnO2-based cathodes for Ca-ion batteries. Thus, mitigating Jahn–Teller distortion not only enhances the performance of zinc-ion battery cathodes but also serves as a universal strategy for advancing multivalent ion battery technologies.

Funding

This work was supported by the Taishan Scholars Project Special Funds (tsqnz20221145), National Natural Science Foundation of China (52302257, 52202239). Natural Science Foundation of Shandong Province (2023HWYQ-083, ZR2023QE019, ZR2022QB162). Jinan University New Talent Research Project (XRC 2402, XRC2412), Science and Technology Project of Jinan University (XKY2208).

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
SEMScanning electron microscope
TEMTransmission electron microscope
HRTEMHigh-resolution transmission electron microscope
SAEDSelected area electron diffraction
XRDPhase analysis of X-ray diffraction
PVAPolyvinyl alcohol

References

  1. Liu, Y.; Han, S.-G.; Li, X.; Luo, Y.; Wu, Y.; Lin, X.; Zhu, Q.-L. Manganese dioxide cathode materials for aqueous zinc ion battery: Development, challenges and strategies. EnergyChem 2025, 7, 100152. [Google Scholar] [CrossRef]
  2. Xu, H.; Yang, W.; Li, M.; Liu, H.; Gong, S.; Zhao, F.; Li, C.; Qi, J.; Wang, H.; Peng, W.; et al. Advances in Aqueous Zinc Ion Batteries based on Conversion Mechanism: Challenges, Strategies, and Prospects. Small 2024, 20, 2310972. [Google Scholar] [CrossRef] [PubMed]
  3. Deng, S.; Xu, B.; Zhao, J.; Fu, H. Advanced design for anti-freezing aqueous zinc-ion batteries. Energy Storage Mater. 2024, 70, 103490. [Google Scholar] [CrossRef]
  4. Yang, J.; Yao, G.; Li, Z.; Zhang, Y.; Wei, L.; Niu, H.; Chen, Q.; Zheng, F. Highly Flexible K-Intercalated MnO2/Carbon Membrane for High-Performance Aqueous Zinc-Ion Battery Cathode. Small 2023, 19, 2205544. [Google Scholar] [CrossRef]
  5. Zhao, L.-L.; Yin, J.-W.; Liu, B.-C.; Wang, P.-F.; Liu, Z.-L.; Xie, Y.; Shu, J.; Zhang, Q.-Y.; Yi, T.-F. Construction of high-performance aqueous zinc-ion batteries by guest pre-intercalation MnO2-based cathodes. Adv. Colloid Interface Sci. 2025, 341, 103499. [Google Scholar] [CrossRef]
  6. Yan, L.; Han, Y.; Zhu, C.; Luo, L.; Qin, Y.; Yu, D.; Liu, B.; Zou, X.; Zhou, Y.; Xiang, B. Achieving high-performance aqueous Zn-ion storage through strong hydrogen bonding and charge redistribution by in-situ induced insertion of non-metallic ion clusters. Nano Energy 2024, 122, 109331. [Google Scholar] [CrossRef]
  7. Yu, X.; Zhang, Z.; Chen, X.; Bao, L.; Peng, J.; Li, X. Engineering d-band center of Mn to strengthen Mn–O bonding for long cycle life zinc-ion battery. J. Mater. Sci. Technol. 2025, 229, 203–212. [Google Scholar] [CrossRef]
  8. Chen, Y.; Lin, C.; Chen, X.; Lu, Z.; Zhang, K.; Liu, Y.; Wang, J.; Han, G.; Xu, G. Modulating the Structure of Interlayer/Layer Matrix on δ-MnO2 via Cerium Doping-Engineering toward High-Performance Aqueous Zinc Ion Batteries. Adv. Energy Mater. 2024, 14, 2304303. [Google Scholar] [CrossRef]
  9. Li, Q.; Wang, C.; Zhu, Y.; Du, W.; Liu, W.; Yao, M.; Wang, Y.; Qian, Y.; Feng, S. Unlocking the critical role of Mg doping in α-MnO2 cathode for aqueous zinc ion batteries. Chem. Eng. J. 2024, 485, 150077. [Google Scholar] [CrossRef]
  10. Choi, S.; Seo, J.K.; Park, H.; Park, J.H.; Kim, S.W.; Shin, S.W.; Bang, J.A.; Jung, K.-N.; Kim, B.G. A Multifunctional Crosslinked F-Free Binder for MnO2 Microparticle Thick Cathode in Aqueous Zinc-Ion Batteries. Adv. Funct. Mater. 2025, 35, 2417179. [Google Scholar] [CrossRef]
  11. Zhou, Z.; Tong, J.; Zou, X.; Wang, Y.; Bai, Y.; Yang, Y.; Li, Y.; Wang, C.; Liu, S. Ultra-low diffusion barrier tetramethyl ammonium cation-intercalated layered MnO2 for high-performance aqueous zinc-ion batteries. J. Mater. Chem. A 2024, 12, 10923–10931. [Google Scholar] [CrossRef]
  12. Ma, K.; Ge, S.; Fu, R.; Feng, C.; Zhao, H.; Shen, X.; Liang, G.; Zhao, Y.; Jiao, Q. Uncovering Se, P co-doping effect in MnO2 toward high-performance aqueous zinc-ion batteries. Chem. Eng. J. 2024, 484, 149525. [Google Scholar] [CrossRef]
  13. Moon, H.; Ha, K.-H.; Park, Y.; Lee, J.; Kwon, M.-S.; Lim, J.; Lee, M.-H.; Kim, D.-H.; Choi, J.H.; Choi, J.-H.; et al. Direct Proof of the Reversible Dissolution/Deposition of Mn2+/Mn4+ for Mild-Acid Zn-MnO2 Batteries with Porous Carbon Interlayers. Adv. Sci. 2021, 8, 2003714. [Google Scholar] [CrossRef] [PubMed]
  14. Xia, X.; Zhao, Y.; Zhao, Y.; Xu, M.; Liu, W.; Sun, X. Mo doping provokes two electron reaction in MnO2 with ultrahigh capacity for aqueous zinc ion batteries. Nano Res. 2023, 16, 2511–2518. [Google Scholar] [CrossRef]
  15. Ren, Y.; Meng, F.; Zhang, S.; Ping, B.; Li, H.; Yin, B.; Ma, T. CNT@MnO2 composite ink toward a flexible 3D printed micro-zinc-ion battery. Carbon Energy 2022, 4, 446–457. [Google Scholar] [CrossRef]
  16. Huang, H.; Feng, H.; He, Z.; Huang, Y.; Xu, J.; Hu, C.; Chen, Z.; Yang, Z.; Zhang, W. Enhancing Zn2+/H+ Joint Charge Storage in MnO2: Concurrently Tailoring Mn’s Local Electron Density and O’s p-Band Center. ACS Nano 2024, 18, 25601–25613. [Google Scholar] [CrossRef]
  17. Yu, X.; Qian, K.; Du, L.; Zhang, J.; Lu, N.; Miao, Z.; Li, Y.; Kobayashi, H.; Yan, X.; Li, R. A strong Jahn–Teller distortion in Mn3O4–MnO heterointerfaces for enhanced silver catalyzed formaldehyde reforming into hydrogen. Sustain. Energy Fuels 2022, 6, 3068–3077. [Google Scholar] [CrossRef]
  18. Dai, L.; Wang, Y.; Sun, L.; Ding, Y.; Yao, Y.; Yao, L.; Drewett, N.E.; Zhang, W.; Tang, J.; Zheng, W. Jahn–Teller Distortion Induced Mn2+-Rich Cathode Enables Optimal Flexible Aqueous High-Voltage Zn-Mn Batteries. Adv. Sci. 2021, 8, 2004995. [Google Scholar] [CrossRef]
  19. 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 2024, 12, 25491–25503. [Google Scholar] [CrossRef]
  20. Wei, T.; Ren, Y.; Li, Z.; Zhang, X.; Ji, D.; Hu, L. Bonding interaction regulation in hydrogel electrolyte enable dendrite-free aqueous zinc-ion batteries from −20 to 60 °C. Chem. Eng. J. 2022, 434, 134646. [Google Scholar] [CrossRef]
  21. Dong, L.; Yang, W.; Yang, W.; Tian, H.; Huang, Y.; Wang, X.; Xu, C.; Wang, C.; Kang, F.; Wang, G. Flexible and conductive scaffold-stabilized zinc metal anodes for ultralong-life zinc-ion batteries and zinc-ion hybrid capacitors. Chem. Eng. J. 2020, 384, 123355. [Google Scholar] [CrossRef]
  22. Ma, Y.; Xu, M.; Liu, R.; Xiao, H.; Liu, Y.; Wang, X.; Huang, Y.; Yuan, G. Molecular tailoring of MnO2 by bismuth doping to achieve aqueous zinc-ion battery with capacitor-level durability. Energy Storage Mater. 2022, 48, 212–222. [Google Scholar] [CrossRef]
  23. 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 MnO Nanosheets as High Performance Cathode for Zinc-Ion Batteries. Energy Environ. Mater. 2023, 6, e12378. [Google Scholar] [CrossRef]
  24. Wang, W.; Black, A.P.; Liu, C.; Martin-Diaconescu, V.; Simonelli, L.; Tonti, D. High performance N-doped carbon nanosheet/MnO2 cathode derived from bacterial cellulose for aqueous Zn-ion batteries. J. Mater. Chem. A 2023, 11, 17272–17281. [Google Scholar] [CrossRef]
  25. Gou, L.; Li, J.; Liang, K.; Zhao, S.; Li, D.; Fan, X. Bi-MOF Modulating MnO2 Deposition Enables Ultra-Stable Cathode-Free Aqueous Zinc-Ion Batteries. Small 2023, 19, 2208233. [Google Scholar] [CrossRef]
  26. Shi, X.; Zhou, C.; Yang, F.; Shan, L.; Tang, B.; Zhang, J.; Nan, Q.; Xie, Y.; Li, J.; Li, H.; et al. Spontaneous Interface Layer Engineering of Ag2Mn8O16 Cathode via Anodic Oxidation Strategy toward High-Performance Aqueous Zinc-Ion Batteries. ACS Energy Lett. 2024, 9, 1063–1072. [Google Scholar] [CrossRef]
  27. Zhang, K.; Chen, H.; Xiao, B.; Chen, J.; Yan, W.; Liu, Y.; Hu, C.; Chen, L.; Huang, J.; Peng, S. Interlayer Ag+ and In Situ Ag Metal Particles Synergistically Strengthening MnO2 Cathode Performance for Aqueous Zinc-Ion Batteries. ACS Appl. Energy Mater. 2023, 6, 1467–1474. [Google Scholar] [CrossRef]
  28. Zhao, Y.; Zhang, P.; Liang, J.; Xia, X.; Ren, L.; Song, L.; Liu, W.; Sun, X. Unlocking Layered Double Hydroxide as a High-Performance Cathode Material for Aqueous Zinc-Ion Batteries. Adv. Mater. 2022, 34, 2204320. [Google Scholar] [CrossRef]
  29. Peng, H.; Xia, F.; Zhang, C.; Zhuo, H.; Peng, X.; Song, P.; Sun, C.; Wu, J. The Coupling of Local Strain and K+-Ion Release Induced Phase Transition Heterogeneity in Tunnel MnO2. Adv. Funct. Mater. 2022, 32, 2113424. [Google Scholar] [CrossRef]
  30. Cui, Y.-f.; Zhuang, Z.-b.; Xie, Z.-l.; Cao, R.-f.; Hao, Q.; Zhang, N.; Liu, W.-q.; Zhu, Y.-h.; Huang, G. High-Energy and Long-Lived Zn–MnO2 Battery Enabled by a Hydrophobic-Ion-Conducting Membrane. ACS Nano 2022, 16, 20730–20738. [Google Scholar] [CrossRef]
  31. 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]
  32. Hua, J.; Wu, X.; Xie, F.; Wu, M. Cr-doping δ-MnO2 as the cathode material for aqueous zinc-ion batteries. Surf. Interfaces 2025, 62, 106312. [Google Scholar] [CrossRef]
  33. Li, X.; He, D.; Zhou, Q.; Zhou, X.; Wang, Z.; Wei, C.; Shi, Y.; Hu, X.; Huang, B.; Yang, Z.; et al. Deciphering anomalous zinc ion storage in intermediate-state MnO2 during layer-to-tunnel structural transition. Energy Environ. Sci. 2024, 17, 9195–9204. [Google Scholar] [CrossRef]
  34. Lv, X.; Wang, Y.; Zhang, W.; Wang, T. High valence MnO2 as an aqueous zinc ion battery cathode prepared using a secondary hydrothermal method. Dalton Trans. 2024, 54, 1182–1190. [Google Scholar] [CrossRef]
  35. Wang, H.; Guo, R.; Ma, Y.; Zhou, F. Cross-Doped Mn/Mo Oxides with Core-Shell Structures Designed by a Self-Template Strategy for Durable Aqueous Zinc-Ion Batteries. Adv. Funct. Mater. 2023, 33, 2301351. [Google Scholar] [CrossRef]
  36. Liu, H.; Huang, X.; Du, X.; Zhang, M.; Cui, X.; Wang, Q.; Wang, H. CTAB controlled electrochemical deposition of manganese dioxide with enhanced performance in aqueous Zn-ion battery. Mater. Sci. Eng. B 2022, 282, 115777. [Google Scholar] [CrossRef]
  37. Qiaohui, D.; Yao, W.; Shuyu, D.; Denis, Y.W.Y. Facile electrode additive stabilizes structure of electrolytic MnO2 for mild aqueous rechargeable zinc-ion battery. J. Power Sources 2022, 528, 231194. [Google Scholar] [CrossRef]
  38. Carbery, W.P.; Farfan, C.A.; Ulbricht, R.; Turner, D.B. The phonon-modulated Jahn–Teller distortion of the nitrogen vacancy center in diamond. Nat. Commun. 2024, 15, 8646. [Google Scholar] [CrossRef]
  39. Wang, P.-F.; Jin, T.; Zhang, J.; Wang, Q.-C.; Ji, X.; Cui, C.; Piao, N.; Liu, S.; Xu, J.; Yang, X.-Q.; et al. Elucidation of the Jahn-Teller effect in a pair of sodium isomer. Nano Energy 2020, 77, 105167. [Google Scholar] [CrossRef]
  40. Waters, M.D.J.; Ng, Z.X.; Monahan, N.R.; Wörner, H.J. Ultrafast Imaging of the Jahn–Teller Topography in Carbon Tetrachloride. J. Am. Chem. Soc. 2023, 145, 7659–7666. [Google Scholar] [CrossRef]
  41. Miñarro, A.S.; Villa, M.; Casals, B.; Plana-Ruiz, S.; Sánchez, F.; Gázquez, J.; Herranz, G. Spin-orbit entanglement driven by the Jahn-Teller effect. Nat. Commun. 2024, 15, 8694. [Google Scholar] [CrossRef]
  42. Streltsov, S.V.; Khomskii, D.I. Jahn-Teller Effect and Spin-Orbit Coupling: Friends or Foes? Phys. Rev. X 2020, 10, 031043. [Google Scholar] [CrossRef]
  43. Zheng, Y.; Xie, H.; Li, J.; Hui, K.S.; Yu, Z.; Xu, H.; Dinh, D.A.; Ye, Z.; Zha, C.; Hui, K.N. Insights into the Jahn-Teller Effect in Layered Oxide Cathode Materials for Potassium-Ion Batteries. Adv. Energy Mater. 2024, 14, 2400461. [Google Scholar] [CrossRef]
  44. Jin, J.; Wang, Y.; Zhao, X.; Hu, Y.; Li, T.; Liu, H.; Zhong, Y.; Jiao, L.; Liu, Y.; Chen, J. Intrinsic Distortion against Jahn-Teller Distortion: A New Paradigm for High-Stability Na-Ion Layered Mn-Rich Oxide Cathodes. Angew. Chem. Int. Ed. 2025, 64, e202423728. [Google Scholar] [CrossRef] [PubMed]
  45. Sakamoto, M.; Hada, M.; Ota, W.; Uesugi, F.; Sato, T. Localised surface plasmon resonance inducing cooperative Jahn–Teller effect for crystal phase-change in a nanocrystal. Nat. Commun. 2023, 14, 4471. [Google Scholar] [CrossRef] [PubMed]
  46. Drosou, M.; Zahariou, G.; Pantazis, D.A. Orientational Jahn–Teller Isomerism in the Dark-Stable State of Nature’s Water Oxidase. Angew. Chem. Int. Ed. 2021, 60, 13493–13499. [Google Scholar] [CrossRef]
  47. Zhu, S.; Wang, Y.; Zhang, J.; Sheng, J.; Yang, F.; Wang, M.; Ni, J.; Jiang, H.; Li, Y. Jahn-Teller Effect Directed Bandgap Tuning of Birnessite for Pseudocapacitive Application. Energy Environ. Mater. 2022, 6, e12382. [Google Scholar] [CrossRef]
  48. Rui-Jie, L.; Xun-Lu, L.; Jie-Ying, D.; Jian, B.; Cui, M.; Chong-Yu, D.; Xin-Yin, C.; Xiao-Jing, W.; Yong-Ning, Z. Suppressing Jahn-Teller distortion and phase transition of K0.5MnO2 by K-site Mg substitution for potassium-ion batteries. Energy Storage Mater. 2022, 47, 408–414. [Google Scholar] [CrossRef]
  49. Zuo, C.; Hu, Z.; Qi, R.; Liu, J.; Li, Z.; Lu, J.; Dong, C.; Yang, K.; Huang, W.; Chen, C.; et al. Double the Capacity of Manganese Spinel for Lithium-Ion Storage by Suppression of Cooperative Jahn–Teller Distortion. Adv. Energy Mater. 2020, 10, 2000363. [Google Scholar] [CrossRef]
  50. Zuo, C.; Xiong, F.; Shao, Y.; Li, M.; Zhu, D.; Zhu, J.; Wang, J.; An, Q. Mitigating Jahn–Teller effect of MnO2 via charge regulation of Mn-local environment for advanced calcium storage. Energy Storage Mater. 2024, 72, 103763. [Google Scholar] [CrossRef]
  51. Zhao, X.; Zhang, F.; Li, H.; Dong, H.; Yan, C.; Meng, C.; Sang, Y.; Liu, H.; Guo, Y.-G.; Wang, S. Dynamic heterostructure design of MnO2 for high-performance aqueous zinc-ion batteries. Energy Environ. Sci. 2024, 17, 3629–3640. [Google Scholar] [CrossRef]
  52. Zhang, Z.; Zheng, J.; Chen, X.; Yu, X.; Li, L.; Bao, L.; Peng, J.; Li, X. Achieving ultra-long cycling life for MnO2 cathode: Modulating Mn3+ spin state to suppress Jahn–Teller distortion and manganese dissolution. Energy Storage Mater. 2025, 76, 104128. [Google Scholar] [CrossRef]
  53. Xie, M.; Zhang, X.; Wang, R.; Jiao, Y.; Shu, Z.; Shan, S.; Bian, Y.; Lin, H.; Chen, J.; Xu, Y. Mn-O bond Engineering Mitigating Jahn-Teller Effects of Manganese Oxide for Aqueous Zinc-ion Battery Applications. Chem. Eng. J. 2024, 494, 152908. [Google Scholar] [CrossRef]
  54. Yang, B.; Li, K.; Yang, C.; Zhang, Y.; Gong, Y. Orbital hybridization and electron transfer triggered by V-dopant and symmetry breakage to stabilize low-valent Mn for zinc-ion batteries. Chem. Eng. J. 2025, 513, 163032. [Google Scholar] [CrossRef]
  55. Dai, Q.; Li, L.; Tu, T.; Zhang, M.; Song, L. An appropriate Zn2+/Mn2+ concentration of the electrolyte enables superior performance of AZIBs. J. Mater. Chem. A 2022, 10, 23722–23730. [Google Scholar] [CrossRef]
  56. Han, M.; Qin, L.; Liu, Z.; Zhang, L.; Li, X.; Lu, B.; Huang, J.; Liang, S.; Zhou, J. Reaction mechanisms and optimization strategies of manganese-based materials for aqueous zinc batteries. Mater. Today Energy 2020, 20, 100626. [Google Scholar] [CrossRef]
  57. Wang, Y.; Zhang, Y.; Duan, Q.; Lee, P.-K.; Wang, S.; Yu, D.Y.W. Engineering cathode-electrolyte interface of graphite to enable ultra long-cycle and high-power dual-ion batteries. J. Power Sources 2020, 471, 228466. [Google Scholar] [CrossRef]
  58. Li, X.; Rong, H.; Zhang, J.; Wang, D.; Li, Y. Modulating the local coordination environment of single-atom catalysts for enhanced catalytic performance. Nano Res. 2020, 13, 1842–1855. [Google Scholar] [CrossRef]
  59. 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]
  60. Qu, W.; Cai, Y.; Chen, B.; Zhang, M. Heterointerface Engineering-Induced Oxygen Defects for the Manganese Dissolution Inhibition in Aqueous Zinc Ion Batteries. Energy Environ. Mater. 2024, 7, e12645. [Google Scholar] [CrossRef]
  61. Zhao, Y.; Zhang, S.; Zhang, Y.; Liang, J.; Ren, L.; Fan, H.J.; Liu, W.; Sun, X. Vacancy-rich Al-doped MnO2 cathodes break the trade-off between kinetics and stability for high-performance aqueous Zn-ion batteries. Energy Environ. Sci. 2024, 17, 1279–1290. [Google Scholar] [CrossRef]
  62. Wu, B.; Zhang, G.; Yan, M.; Xiong, T.; He, P.; He, L.; Xu, X.; Mai, L. Graphene Scroll-Coated α-MnO2 Nanowires as High-Performance Cathode Materials for Aqueous Zn-Ion Battery. Small 2018, 14, 1703850. [Google Scholar] [CrossRef]
  63. Ye, J.-J.; Li, P.-H.; Hou, Z.; Zhang, W.; Zhu, W.; Jin, S.; Ji, H. Se-dopant Modulated Selective Co-Insertion of H+ and Zn2+ in MnO2 for High-Capacity and Durable Aqueous Zn-Ion Batteries. Angew. Chem. Int. Ed. 2024, 63, e202410900. [Google Scholar] [CrossRef]
  64. Cai, K.; Luo, S.-h.; Qian, L.; Meng, X.; Yan, S.-x.; Guo, J.; Wang, Q.; Ji, X.-b.; Zhou, X.-y. Three-dimensional porous composite Mn2O3@PPy as cathode material for zinc ion battery with high energy density. J. Power Sources 2023, 564, 232854. [Google Scholar] [CrossRef]
  65. Liu, L.; Ding, S.; Yao, L.; Liu, M.; Li, S.; Zhao, Q.; Qin, R.; Pan, F. In situ BaSO4 coating enabled activation-free and ultra-stable δ-MnO2 for aqueous Zn-ion batteries. Chem. Commun. 2023, 59, 6227–6230. [Google Scholar] [CrossRef] [PubMed]
  66. Lai, G.; Zhao, Z.; Zhang, H.; Hu, X.; Lu, B.; Liang, S.; Zhou, J. In-situ positive electrode-electrolyte interphase construction enables stable Ah-level Zn-MnO2 batteries. Nat. Commun. 2025, 16, 2194. [Google Scholar] [CrossRef]
  67. Chen, C.; Shi, M.; Zhao, Y.; Yang, C.; Zhao, L.; Yan, C. Al-Intercalated MnO2 cathode with reversible phase transition for aqueous Zn-Ion batteries. Chem. Eng. J. 2021, 422, 130375. [Google Scholar] [CrossRef]
  68. Liu, X.; Yang, F.; Xu, W.; Zeng, Y.; He, J.; Lu, X. Zeolitic Imidazolate Frameworks as Zn2+ Modulation Layers to Enable Dendrite-Free Zn Anodes. Adv. Sci. 2020, 7, 2002173. [Google Scholar] [CrossRef] [PubMed]
  69. Ding, Y.; Cai, C.; Ma, L.; Wang, J.; Mercer, M.P.; Liu, J.; Kramer, D.; Yu, X.; Xue, D.; Zhi, C.; et al. Tailoring MnO2 Cathode Interface via Organic–Inorganic Hybridization Engineering for Ultra-Stable Aqueous Zinc-Ion Batteries. Adv. Energy Mater. 2025, 15, 2402819. [Google Scholar] [CrossRef]
  70. Yang, L.; Zhang, J.; Zhou, X.; Liu, Y. Amorphous aluminum-doped manganese oxide cathode with strengthened performance for aqueous zinc-ion batteries. J. Alloys Compd. 2025, 1012, 178502. [Google Scholar] [CrossRef]
  71. Xiong, T.; Zhang, Y.; Lee, W.S.V.; Xue, J. Defect Engineering in Manganese-Based Oxides for Aqueous Rechargeable Zinc-Ion Batteries: A Review. Adv. Energy Mater. 2020, 10, 2001769. [Google Scholar] [CrossRef]
  72. Zhang, L.; Yang, S.; Fu, W.; Cui, Y.; Wang, J.; Zhao, D.; Yang, C.; Wang, X.; Cao, B. Plasma-induced ε-MnO2 based aqueous zinc-ion batteries and their dissolution-deposition mechanism. J. Mater. Sci. Technol. 2022, 127, 206–213. [Google Scholar] [CrossRef]
  73. Song, Y.; Li, J.; Qiao, R.; Dai, X.; Jing, W.; Song, J.; Chen, Y.; Guo, S.; Sun, J.; Tan, Q.; et al. Binder-free flexible zinc-ion batteries: One-step potentiostatic electrodeposition strategy derived Ce doped-MnO2 cathode. Chem. Eng. J. 2022, 431, 133387. [Google Scholar] [CrossRef]
  74. Dong, H.; Liu, R.; Hu, X.; Zhao, F.; Kang, L.; Liu, L.; Li, J.; Tan, Y.; Zhou, Y.; Brett, D.J.L.; et al. Cathode–Electrolyte Interface Modification by Binder Engineering for High-Performance Aqueous Zinc-Ion Batteries. Adv. Sci. 2023, 10, 2205084. [Google Scholar] [CrossRef]
  75. Li, J.; Ma, D.; Zhang, Q.; Cao, Y.; Wang, J.; Wang, T.; Liu, N.; Ma, C.; Zhang, T.; Zhao, Q.; et al. Optimized integration of intercalation and conversion behaviors for Cu2O/MnO2 hybrid cathodes of zinc ion batteries. J. Energy Chem. 2025, 107, 212–220. [Google Scholar] [CrossRef]
  76. Zhang, R.; Liang, P.; Yang, H.; Min, H.; Niu, M.; Jin, S.; Jiang, Y.; Pan, Z.; Yan, J.; Shen, X.; et al. Manipulating intercalation-extraction mechanisms in structurally modulated δ-MnO2 nanowires for high-performance aqueous zinc-ion batteries. Chem. Eng. J. 2022, 433, 133687. [Google Scholar] [CrossRef]
  77. Guo, C.; Tian, S.; Chen, B.; Liu, H.; Li, J. Constructing α-MnO2@PPy core-shell nanorods towards enhancing electrochemical behaviors in aqueous zinc ion battery. Mater. Lett. 2020, 262, 127180. [Google Scholar] [CrossRef]
  78. Li, Y.; Zhang, N.; Wang, J.-Y.; Wang, P.; Liu, Z.; Zhu, Y.-R.; Shu, J.; Yi, T.-F. Construction of high-voltage aqueous Zn-MnO2 batteries based on polar small-molecule organic acid-induced MnO2/Mn2+ reactions. Chem. Eng. J. 2025, 507, 160415. [Google Scholar] [CrossRef]
  79. Liang, X.; Liu, X.; Wang, P.; Guo, Z.; Chen, X.; Yao, J.; Li, J.; Gan, Y.; Lv, L.; Tao, L.; et al. Ion-exchange induced Ni doping of α-MnO2 cathode with structural modification for aqueous zinc ion batteries. J. Power Sources 2025, 635, 236518. [Google Scholar] [CrossRef]
  80. Zhao, Y.; Zhang, P.; Liang, J.; Xia, X.; Ren, L.; Song, L.; Liu, W.; Sun, X. Uncovering sulfur doping effect in MnO2 nanosheets as an efficient cathode for aqueous zinc ion battery. Energy Storage Mater. 2022, 47, 424–433. [Google Scholar] [CrossRef]
  81. Ding, Y.; Xue, W.; Chen, K.; Yang, C.; Feng, Q.; Zheng, D.; Xu, W.; Wang, F.; Lu, X. Sodium Ion Pre-Intercalation of δ-MnO2 Nanosheets for High Energy Density Aqueous Zinc-Ion Batteries. Nanomaterials 2023, 13, 1075. [Google Scholar] [CrossRef] [PubMed]
  82. Jiao, Y.; Kang, L.; Berry-Gair, J.; McColl, K.; Li, J.; Dong, H.; Jiang, H.; Wang, R.; Corà, F.; Brett, D.J.L.; et al. Enabling stable MnO2 matrix for aqueous zinc-ion battery cathodes. J. Mater. Chem. A 2020, 8, 22075–22082. [Google Scholar] [CrossRef]
  83. Wu, H.; Yan, C.; Xu, L.; Xu, N.; Wu, X.; Jiang, Z.; Zhu, S.; Diao, G.; Chen, M. Super Flexible Cathode Material with 3D Cross-Linking System Based on Polyvinyl Alcohol Hydrogel for Boosting Aqueous Zinc Ion Batteries. ChemElectroChem 2022, 9, e202200288. [Google Scholar] [CrossRef]
  84. Zhang, A.; Zhao, R.; Wang, Y.; Yue, J.; Yang, J.; Wang, X.; Wu, C.; Bai, Y. Hybrid Superlattice-Triggered Selective Proton Grotthuss Intercalation in δ-MnO2 for High-Performance Zinc-Ion Battery. Angew. Chem. Int. Ed. 2023, 62, e202313163. [Google Scholar] [CrossRef] [PubMed]
  85. Jiang, W.; Xu, X.; Liu, Y.; Tan, L.; Zhou, F.; Xu, Z.; Hu, R. Facile plasma treated β-MnO2@C hybrids for durable cycling cathodes in aqueous Zn-ion batteries. J. Alloys Compd. 2020, 827, 154273. [Google Scholar] [CrossRef]
  86. Zeng, X.; Liu, J.; Mao, J.; Hao, J.; Wang, Z.; Zhou, S.; Ling, C.D.; Guo, Z. Toward a Reversible Mn4+/Mn2+ Redox Reaction and Dendrite-Free Zn Anode in Near-Neutral Aqueous Zn/MnO2 Batteries via Salt Anion Chemistry. Adv. Energy Mater. 2020, 10, 1904163. [Google Scholar] [CrossRef]
  87. Huang, J.; Chi, X.; Wu, J.; Liu, J.; Liu, Y. High-concentration dual-complex electrolyte enabled a neutral aqueous zinc-manganese electrolytic battery with superior stability. Chem. Eng. J. 2022, 430, 133058. [Google Scholar] [CrossRef]
Figure 1. Mn 3d orbital energy level splitting in an octahedral field [43].
Figure 1. Mn 3d orbital energy level splitting in an octahedral field [43].
Materials 18 02817 g001
Figure 2. Schematic illustration of strategies to suppress the Jahn–Teller effect [61,63,76,77,78]. Ref. [77] Copyright 2020, Elsevier. Ref. [78] Copyright 2025, Elsevier.
Figure 2. Schematic illustration of strategies to suppress the Jahn–Teller effect [61,63,76,77,78]. Ref. [77] Copyright 2020, Elsevier. Ref. [78] Copyright 2025, Elsevier.
Materials 18 02817 g002
Figure 3. (a) Schematic illustration of the synthesis process for Al-doped MnO2. (b) Atomic structure model of Al0.1MnO2. (c) Comparative cycling performance of Al0.1MnO2 versus pristine MnO2. (d) Schematic illustration of the structure model for Ni-doped MnO2. (e) Optimized structure and charge density localization diagram. (f) Cycle performance of MnO2 and Ni-MnO2. Ref. [61] Copyright 2024, Royal Society of Chemistry. Ref. [79] Copyright 2025, Elsevier.
Figure 3. (a) Schematic illustration of the synthesis process for Al-doped MnO2. (b) Atomic structure model of Al0.1MnO2. (c) Comparative cycling performance of Al0.1MnO2 versus pristine MnO2. (d) Schematic illustration of the structure model for Ni-doped MnO2. (e) Optimized structure and charge density localization diagram. (f) Cycle performance of MnO2 and Ni-MnO2. Ref. [61] Copyright 2024, Royal Society of Chemistry. Ref. [79] Copyright 2025, Elsevier.
Materials 18 02817 g003
Figure 4. (a) Differential charge density analysis of MnO2 and Se-MnO2. (b) Changes in Mn-O bond length at the discharged state of MnO2 and Se-MnO2. (c) Cyclic performance (3.0 A g−1). (d) Adsorption energies of Zn2+ and H+ migration in MnO2 and Se-MnO2. (e) Differential charge density analysis of MnO2 and Se-MnO2 after Zn2+/H+ co-intercalation. (f) The atomic structure model of S-MnO2 and the side view of a schematic illustration of Zn migration in S-MnO2. (g) The Zn ion diffusion barrier profiles for MnO2 and S-MnO2. (h) Charge density difference distribution diagrams for Zn ion in MnO2 and S-MnO2. (i) Cycling performance of MnO2 and S-MnO2 electrodes at a current density of 200 mA g−1. Ref. [63] Copyright 2024, Wiley-VC. Ref. [80] Copyright 2025, Elsevier.
Figure 4. (a) Differential charge density analysis of MnO2 and Se-MnO2. (b) Changes in Mn-O bond length at the discharged state of MnO2 and Se-MnO2. (c) Cyclic performance (3.0 A g−1). (d) Adsorption energies of Zn2+ and H+ migration in MnO2 and Se-MnO2. (e) Differential charge density analysis of MnO2 and Se-MnO2 after Zn2+/H+ co-intercalation. (f) The atomic structure model of S-MnO2 and the side view of a schematic illustration of Zn migration in S-MnO2. (g) The Zn ion diffusion barrier profiles for MnO2 and S-MnO2. (h) Charge density difference distribution diagrams for Zn ion in MnO2 and S-MnO2. (i) Cycling performance of MnO2 and S-MnO2 electrodes at a current density of 200 mA g−1. Ref. [63] Copyright 2024, Wiley-VC. Ref. [80] Copyright 2025, Elsevier.
Materials 18 02817 g004
Figure 5. (a) Schematic illustration of the synthetic process of CuMO electrode. (b) Long-cycling performance (5 A g−1). (c) Formation energies of δ-MnO2 and CuMO. (d) Schematic illustration of the electrochemical reaction mechanism of Zn//CuMO batteries. (e) Schematic illustration of H+ and Zn2+ transport in PVP-MnO2 electrode. (f) Calculation model of H transmission state. (g) TEM images of PVP-MnO2. (h) Schematic diagram of PVP intercalation in δ-MnO2. (i) Galvanostatic charge/discharge profiles. (j) Rate performance of MnO2 and PVP-MnO2. Ref. [76] Copyright 2021, Elsevier. Ref. [84] Copyright 2023, Wiley-VC.
Figure 5. (a) Schematic illustration of the synthetic process of CuMO electrode. (b) Long-cycling performance (5 A g−1). (c) Formation energies of δ-MnO2 and CuMO. (d) Schematic illustration of the electrochemical reaction mechanism of Zn//CuMO batteries. (e) Schematic illustration of H+ and Zn2+ transport in PVP-MnO2 electrode. (f) Calculation model of H transmission state. (g) TEM images of PVP-MnO2. (h) Schematic diagram of PVP intercalation in δ-MnO2. (i) Galvanostatic charge/discharge profiles. (j) Rate performance of MnO2 and PVP-MnO2. Ref. [76] Copyright 2021, Elsevier. Ref. [84] Copyright 2023, Wiley-VC.
Materials 18 02817 g005
Figure 6. (a) TEM images (SAED inset in (a)). (b) Schematic illustration of the two-step intercalation mechanism of MGS cathode. (c) Charge and discharge curves of MGS at current densities ranging from 0.1 to 3 A g−1. (d) Long-term cycling performances at 7 and 3 A g−1 (inset). (e) Preparation processes of β-MnO2@C composites. (f) Rate performance at current rates from 100 mA g−1 to 2 A g−1. (g) Schematic showing the reactions during the discharge process for β-MnO2/Zn cell. (h) Cycle performance of β-MnO2@C. Ref. [62] Copyright 2018, Wiley-VCH. Ref. [85] Copyright 2018, Wiley-VCH.
Figure 6. (a) TEM images (SAED inset in (a)). (b) Schematic illustration of the two-step intercalation mechanism of MGS cathode. (c) Charge and discharge curves of MGS at current densities ranging from 0.1 to 3 A g−1. (d) Long-term cycling performances at 7 and 3 A g−1 (inset). (e) Preparation processes of β-MnO2@C composites. (f) Rate performance at current rates from 100 mA g−1 to 2 A g−1. (g) Schematic showing the reactions during the discharge process for β-MnO2/Zn cell. (h) Cycle performance of β-MnO2@C. Ref. [62] Copyright 2018, Wiley-VCH. Ref. [85] Copyright 2018, Wiley-VCH.
Materials 18 02817 g006
Figure 7. (a) Zn/MnO2 system design with acetate electrolyte. (b) Energy diagram of the dissolution reaction on MnO2 with a bare surface and with an acetate-rich surface. (c) Atomic structures for the dissolution reaction on MnO2 with a bare surface and with an acetate-rich surface. (d) Rate performance (1–70 mA cm−2) with acetate electrolyte. (e) Schematic illustration of the pH stabilization mechanism and reaction process of Zn–Mn electrolytic battery with the HCDCE electrolyte. The pH value changes with various additions of (f) KOH and (g) HAc into the electrolytes. (h) The long-term cycling performance at 2 mA cm−2 with a constant areal charge capacity of 1 mAh cm−2. (i) The rate performance at various current densities (2–10 mA cm−2). Ref. [86] Copyright 2020, Wiley-VCH. Ref. [87] Copyright 2021, Elsevier.
Figure 7. (a) Zn/MnO2 system design with acetate electrolyte. (b) Energy diagram of the dissolution reaction on MnO2 with a bare surface and with an acetate-rich surface. (c) Atomic structures for the dissolution reaction on MnO2 with a bare surface and with an acetate-rich surface. (d) Rate performance (1–70 mA cm−2) with acetate electrolyte. (e) Schematic illustration of the pH stabilization mechanism and reaction process of Zn–Mn electrolytic battery with the HCDCE electrolyte. The pH value changes with various additions of (f) KOH and (g) HAc into the electrolytes. (h) The long-term cycling performance at 2 mA cm−2 with a constant areal charge capacity of 1 mAh cm−2. (i) The rate performance at various current densities (2–10 mA cm−2). Ref. [86] Copyright 2020, Wiley-VCH. Ref. [87] Copyright 2021, Elsevier.
Materials 18 02817 g007
Table 1. Recent capacity retention and Mn dissolution of the cathode.
Table 1. Recent capacity retention and Mn dissolution of the cathode.
CathodeSpecific CapacityCycling PerformanceDissolved Mn2+Refs.
δ-MnO2125 mAh g−1 at 0.2 A g−114.3% after 200 cycles at 1 A g−12.5 mg L−1 after 50 cycles[61]
α-MnO2/MGS382.2 mAh g−1 at 0.3 A g−194% after 3000 cycles at 3 A g−10.42 mg L−1 after 1 cycle[62]
Se-MnO2386 mAh g−1 at 0.1 A g−178% after 5000 cycles at 3 A g−10.71 mg L−1 after 300 cycles[63]
Al-MnO2379 mAh g−1 at 0.2 A g−187% after 1000 cycles at 1 A g−10.12 mg L−1 after 50 cycles[61]
δa-MnO2175 mAh g−1 at 0.5 A g−191% after 500 cycles at 1 A g−10.54 mg L−1 after 100 cycles[60]
Mn2O3@PPy353.9 mAh g−1 at 0.5 A g−182% after 500 cycles at 1 A g−10.27 mg L−1 after 200 cycles[64]
BMO348 mAh g−1 at 0.1 A g−160% after 2000 cycles at 1 A g−10.015 mg L−1 after 100 cycles[65]
δ-MnO2
(ZS-DOP electrolyte)
160 mAh g−1 at 1 A g−180% after 70 cycles at 7.5 mA g−11.2 mg L−1 after 100 cycles[66]
AMO400 mAh g−1 at 0.1 A g−194.5% after 2000 cycles at 2 A g−10.25 mg L−1 after 300 cycles[67]
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

Duan, J.; Huang, M.; Song, M.; Zhou, W.; Tan, H. Suppressing Jahn–Teller Distortion in Manganese Oxides for High-Performance Aqueous Zinc-Ion Batteries. Materials 2025, 18, 2817. https://doi.org/10.3390/ma18122817

AMA Style

Duan J, Huang M, Song M, Zhou W, Tan H. Suppressing Jahn–Teller Distortion in Manganese Oxides for High-Performance Aqueous Zinc-Ion Batteries. Materials. 2025; 18(12):2817. https://doi.org/10.3390/ma18122817

Chicago/Turabian Style

Duan, Jiangfeng, Man Huang, Ming Song, Weijia Zhou, and Hua Tan. 2025. "Suppressing Jahn–Teller Distortion in Manganese Oxides for High-Performance Aqueous Zinc-Ion Batteries" Materials 18, no. 12: 2817. https://doi.org/10.3390/ma18122817

APA Style

Duan, J., Huang, M., Song, M., Zhou, W., & Tan, H. (2025). Suppressing Jahn–Teller Distortion in Manganese Oxides for High-Performance Aqueous Zinc-Ion Batteries. Materials, 18(12), 2817. https://doi.org/10.3390/ma18122817

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