Yuhang Dai†
*a,
Chengyi Zhang†c,
Xinyu Zhangb,
Peie Jiangab,
Jie Chen
b,
Wei Zong
a,
Sicheng Zhengb,
Xuan Gao*b,
Thomas J. Macdonaldd and
Guanjie He
*b
aDepartment of Engineering Science, University of Oxford, Oxford OX1 3PJ, UK. E-mail: yuhang.dai@eng.ox.ac.uk
bChristopher Ingold Laboratory, Department of Chemistry, University College London, London WC1H 0AJ, UK. E-mail: xuan.gao.21@ucl.ac.uk; g.he@ucl.ac.uk
cSchool of Chemical Sciences, The University of Auckland, Auckland 1010, New Zealand
dDepartment of Electronic and Electrical Engineering, University College London, London WC1E 7JE, UK
First published on 9th September 2025
Aqueous zinc-ion batteries (AZIBs) are attractive for large-scale energy storage due to their intrinsic safety, low cost, and environmental compatibility. However, the high charge-to-radius (q/r) ratio of Zn2+ leads to strong solvation and sluggish solid-state diffusion, which hinder efficient charge transport across solid–solid and solid–liquid interfaces. These limitations reduce both cycling stability and rate performances. In this review, we summarize interfacial transport regulation strategies, including solid–solid interfacial modulation via electrostatic fields, interfacial bonding, and ion–electron decoupling to enhance solid-state Zn2+ mobility. We further discuss solid–liquid interfacial desolvation regulation including water activity control, solvation structure tuning, and selective ion channels to mitigate desolvation barriers. We also describe emerging mechanisms involving water dissociation at interfaces, where protons and hydroxide ions act as alternative charge carriers. These unconventional pathways can complement or even outperform traditional Zn2+ intercalation. Collectively, these interfacial strategies not only accelerate Zn2+ transport but also introduce new electrochemical phenomena that boost capacity and rate performances of AZIBs. Advancing the deliberate design and mechanistic understanding of such interfacial processes will be essential to unlocking the full potential of next-generation AZIBs.
Broader contextThe development of safe, long-lasting, and cost-effective energy storage technologies is essential for the widespread integration of renewable energy. Aqueous zinc-ion batteries (AZIBs) have attracted increasing interests due to their intrinsic safety, non-flammable water-based electrolytes, and the low cost and abundance of their electrode materials. These advantages make AZIBs promising for grid-scale storage, peak-shaving, and fast-charging applications. However, their practical performance remains limited by interfacial challenges, as Zn2+ ions interact strongly with both solid electrodes and water molecules in the solvation shell. Problems such as dendrite growth, parasitic water decomposition reactions, and sluggish ion transport have not yet been fully resolved. While advances in material and electrolyte design have yielded significant improvements, interfacial regulation is increasingly recognized as a key factor that directly governs Zn2+ migration and overall electrochemical performance. Recent studies suggest that rationally engineered interfaces can not only suppress degradation but also boost capacity and charging rates. Notably, emerging studies revealed that water dissociation products, such as protons and hydroxide ions, can actively participate in charge storage, enabling unconventional mechanisms beyond Zn2+ insertion. Continued advances in understanding and controlling interfacial processes will be essential for realizing the full potential of AZIBs in practical energy storage systems. |
Despite these advantages, the practical deployment of AZIBs remains hindered by intrinsic scientific challenges. The key issue stems from the high charge density and small ionic radius of Zn2+ (+2, 0.74 Å),7 compared to Li+ (+1, 0.76 Å),8 resulting in much stronger electrostatic interactions with host lattices and coordinated water molecules in its solvation shell. Consequently, Zn2+ experiences sluggish solid-state diffusion and a high desolvation (removal of coordinated solvent molecules from ions prior to insertion into the host material) penalty at electrode–electrolyte interfaces. These limitations induce uneven Zn plating, dendrite formation, parasitic side reactions such as hydrogen evolution, and structural degradation of cathode materials through dissolution or distortion, ultimately impairing cycling stability and overall efficiency.
To tackle these challenges, extensive efforts have been devoted to both electrode and electrolyte design. On the cathode side, phase engineering, pre-intercalation strategies,9 and anion redox mechanisms10 have been explored to stabilize host structures and promote Zn2+ storage. Electrolyte engineering has primarily targeted the bulk solvation environment of Zn2+ through highly concentrated electrolytes (water-in-salt systems11), eutectic formulations,12 solvation structure modulation via donor number adjustments,13 and diverse additives, whose functional groups can critically modulate solvation structure and interfacial chemistry to mitigate side reactions.14–16 For zinc metal anodes, crystallographic orientation control has been employed to suppress dendrite growth and improve reversibility.17
In parallel, interfacial regulation has increasingly emerged as a systematic strategy to address multiple bottlenecks that fundamentally govern Zn2+ transport and reaction kinetics. Integrated with the European initiative Battery 2030+, which underscores the importance of interface control for future battery innovation, recent interfacial designs in AZIBs encompass solid–solid composite architectures within cathodes,18 protective interphases at electrode–electrolyte interfaces,19 and Zn anode modifications such as cation-adsorbed shielding layers.20 These approaches collectively enhance Zn2+ mobility, facilitate desolvation, suppress parasitic reactions, and enable additional reaction pathways related to interfacial water dissociation, thereby improving energy density and cycling stability especially under high-rate cycling conditions.
Herein, recent advances in interfacial regulation of Zn2+ transport are systematically summarized. We specifically focus on interfacial phenomena that govern charge storage and reversibility, rather than providing a comprehensive survey of all aspects of AZIBs such as Zn anode morphology or full-cell engineering. In contrast to previous reviews that largely emphasize Zn anodes or electrolyte design, our work complements them by highlighting cathode-side interfacial regulation.21–26 Within this scope, we integrate mechanistic insights to uncover emerging interfacial processes in which water dissociation products directly participate in charge storage. These insights reveal unconventional capacity gains and exceptionally fast kinetics. A deeper mechanistic understanding of interfacial Zn2+ transport is expected to inform novel design principles for next-generation high-performance aqueous zinc-ion batteries.
![]() | ||
Fig. 1 Solid–solid interfacial regulation of Zn2+ transport through electrostatic fields and chemical bonding. (a) DFT-calculated Zn2+ adsorption energies on H2V3O8–MXene interface with various configurations. (b) Charge density difference showing interfacial charge redistribution. (c) Zn2+ migration barriers on pristine H2V3O8 and H2V3O8–MXene interface. Reproduced from ref. 27. Copyright (2022), with permission from Elsevier. (d) Schematic of interfacial activation and Zn2+ transport pathways enabled by V–O–Ti–C bonding. (e) Dynamic evolution of V–O–Ti bonds during cycling, facilitating reversible Zn2+ insertion. Reproduced from ref. 33. Copyright (2023), with permission from Wiley-VCH. |
Similar electrostatic modulation effects have been reported in various heterointerface systems. For instance, the V2C–V2O5 heterostructure forms a highly polarized interface that strengthens the local electric field, enhancing Zn2+ adsorption and reducing migration barriers.28 Likewise, the Cu-HHTP–MXene (HHTP = 2,3,6,7,10,11-hexahydroxytriphenylene) hybrid forms extended two-dimensional channels with open coordination environments that facilitate Zn2+ transport under the influence of interfacial electric fields.29 In another example, MnO–nitrogen-doped graphene aerogel (NGA) composites leverage interfacial charge redistribution from MnO–NGA interactions to promote Zn2+ migration.30 Interfacial polarization has also been exploited in MoS2–MXene heterostructures, where strong interlayer interactions contribute to reduced Zn2+ diffusion barrier.31 In another case, poly(3,4-ethylenedioxythiophene) (PEDOT) modified V2O5 nanosheets exhibit improved Zn2+ storage kinetics by facilitating capacitive behaviors through an enhanced surface charge density.32
Collectively, these studies demonstrate that engineering built-in electric fields at heterointerfaces provides an efficient means to regulate Zn2+ adsorption and migration, offering improved charge transfer and fast ion transport pathways.
Various other heterostructures have implemented similar bonding strategies to regulate Zn2+ transport. Specifically, in VO2@Co–N–C composites, atomically dispersed Co–N coordination centers form Co–O–V bonds with VO2, which modulate local charge distribution and facilitate Zn2+ adsorption and directional diffusion.34 Similarly, the V2O5@NC system introduces nitrogen-doped carbon interfaces that generate V–O–C bonds, guiding Zn2+ migration while stabilizing structural frameworks.35 In another case, the IC@MnO2 material introduces positively charged interstitial carbon (IC) into the MnO2 lattice, leading to the formation of Mn–O–C bonds. These interfacial bonds facilitate simultaneous Zn2+ and H+ intercalation, improving both capacity and reversibility.36 In addition to these examples, the V2C–V2O5 and Cu-HHTP–MXene heterostructures also exhibit interfacial bonding features that contribute to the Zn2+ transport enhancement alongside their electrostatic effects.28,29
These interfacial bonding configurations offer tunable energy landscapes that enable fast, directional Zn2+ migration with suppressed structural degradation, providing an attractive pathway for high-rate AZIBs.
![]() | ||
Fig. 2 Solid–solid interfacial regulation of Zn2+ transport through electron–ion decoupling. (a) STEM image of VOx–rGO showing VOx clusters and Zn2+ storage at the interface. (b) Schematic of spatially separated Zn2+ storage and electron conduction enabling pseudocapacitive behavior. (c) Raman spectra tracking V–O–C bond changes upon cycling. (d) EELS analysis of vanadium valence evolution. (e) C 1s XPS spectra showing V–O–C bonding and π–π* signals. Reproduced from ref. 38. Copyright (2023), with permission from Wiley-VCH. |
Together, these findings reveal a decoupled transport mechanism in which Zn2+ ions are stored via redox reactions on VOx, while electrons are conducted through rGO (Fig. 2(b)). Galvanostatic intermittent titration technique (GITT) measurements in this work further reveal that the Zn2+ diffusion coefficient of VOx–rGO is over two orders of magnitude higher than that of bulk VO2, corroborating the markedly accelerated ion transport enabled by the decoupled electron/ion pathways. This system delivers a high specific capacity of 443 mAh g− 1 at 100 mA g−1 and maintains excellent rate capability, retaining 174.4 mAh g−1 at 100 A g−1. Moreover, the charge storage process is governed by pseudocapacitive (surface-controlled fast charge storage process) behavior, and such an interfacial Zn2+ storage strategy sustains high-rate performance even at practical mass loadings up to 12.72 mg cm−2 (3.67 mAh cm−2 at 1.27 mA cm−2). In addition, its facile synthesis process offers promising scalability for large-scale electrode fabrication.
The job-sharing principle has been extensively explored in VOx–graphene oxide (GO) composites by Liu et al., where hydrogen-bonded N⋯H–O interfaces further enhance charge separation and ion storage efficiency.39 In this system, Zn2+ ions are selectively adsorbed onto the VOx domains, while electrons are delocalized within the GO sheets, collectively establishing interfacial intercalation pseudocapacitance characterized by rapid kinetics and minimal vanadium redox fluctuation. Quantitative analyses revealed that more than 90% of Zn2+ storage occurred at the interface, while bulk contributions remained negligible. The hydrogen-bonded interface also imparted a self-healing capability, enabling stable long-term cycling even at high mass loadings.
These interface-governed job-sharing designs offer a unique pathway to simultaneously enhance Zn2+ diffusion, electron conductivity, and structural stability, thereby enabling AZIBs with both high capacity and fast charge–discharge characteristics.
Despite their advantages, these solid–solid interfacial strategies also present trade-offs. Electrostatic field designs often require complex heterostructures, interfacial bonding may hinder Zn2+ transport if overly strong, and ion–electron decoupling relies on added conductive frameworks that can increase material complexity. Balancing structural benefits with synthetic feasibility remains essential for their practical deployment.
One representative example is the work by Zhong et al., who introduced an interfacial desolvation enhancer (IDE) molecule that self-assembles as a monolayer on hydrated vanadium oxide cathodes.40 As shown in Fig. 3(a)–(c), molecular dynamics simulations and coordination number analyses reveal that IDE molecules preferentially adsorb on the cathode surface, displacing interfacial H2O molecules and reducing the number of coordinated water molecules in the Zn2+ solvation shell—from five in the control to four in the IDE20 system. Importantly, IDE does not participate in the Zn2+ solvation shell, thereby minimizing its influence on ion size and preserving transport kinetics. This selective interfacial masking weakens Zn2+–H2O interactions and lowers the desolvation energy, resulting in enhanced rate capability and cycling stability. The IDE-coated cathodes maintain structural integrity and deliver stable performances even under high current densities.
![]() | ||
Fig. 3 Representative cathode–electrolyte interfacial desolvation regulation strategies in AZIBs. (a), (b) Radial distribution function and coordination number of Zn2+ with OTf− and H2O in 2 M Zn(OTf)2 electrolyte (a) and 2 M Zn(OTf)2 with 20 mM isosorbide dimethyl ether (IDE) (b). (c) Snapshot of IDE monolayer on hydrated VOH surface showing distributions of H2O, OTf−, and Zn2+. Reproduced from ref. 40 Copyright (2024), with permission from Royal Society of Chemistry. (d) Energy barriers for the desolvation process of [Zn(H2O)5]2+(CF3SO3)− on TiS2 and TiS2–TiO2 surfaces. (e), (f) Mass change per mole of electron for TiS2–TiO2 electrode (TSO-2) (e) and pure TiS2 electrode (TS) (f) electrodes during cyclic voltammetry (CV) scans. Reproduced from ref. 43 Copyright (2025), with permission from Wiley-VCH. (g) Mean square displacement of different ions in Znx(OTf)y(OH)2x−y·nH2O (ZnOTf-LDH) from ab initio molecular dynamics (AIMD) simulations. (h) Adsorption energy of Zn2+ and H2O within the channel of ZnOTf-LDH. (i) CV curves of V6O13 and V6O13 coated with ZnOTf-LDH at 0.1 mV s−1. Reproduced from ref. 44 under the Creative Commons CC BY license. |
Other studies have further illustrated water masking approaches. Gou et al. constructed a hydrophobic PEDOT layer inspired by biological membranes on MnO2 cathodes, which reduced water access and promoted Zn2+ desolvation.41 Furthermore, Ding et al. utilized F-coordination to regulate local water activity and enhance desolvation on MnO2 surfaces.42 These results demonstrate the broad applicability of water masking strategies to improve cathode interfacial kinetics in AZIBs.
Chen et al.43 constructed a TiS2–TiO2 heterostructure that induces a built-in electric field (BIEF) via interfacial electronic redistribution between the two components. As shown in Fig. 3(d), DFT calculations reveal that the desolvation energy barrier for [Zn(H2O)5]2+ complexes is significantly reduced at the TiS2–TiO2 interface (2.97 eV) compared to pure TiS2 (3.42 eV), highlighting the desolvation-facilitating role of the heterojunction.
To experimentally validate this effect, the authors compared a TiS2–TiO2 electrode (TSO-2, prepared via 2 min O2 plasma treatment) and a pure TiS2 electrode (TS) using electrochemical quartz crystal microbalance (EQCM). As shown in Fig. 3(e) and (f), the TSO-2 electrode exhibits a mass change per electron (∼27.4 g mol−1) close to the theoretical value for bare Zn2+ insertion, indicating effective desolvation at the interface. In contrast, the TS electrode shows a higher mass change (∼39.6 g mol−1), consistent with hydrated Zn2+ ([Zn(H2O)]2+) co-intercalation. These findings confirm that the built-in electric field at the TiS2–TiO2 interface promotes Zn2+ desolvation and alters the nature of the intercalating species. Altogether, this heterostructure design mitigates strain from hydrated ion insertion, promotes fast Zn2+ transport, and enables high-rate cycling (e.g., 160.9 mAh g−1 at 5 A g−1).
Complementary work by Sun et al. employed Ca2+ pre-intercalation in V2O5 to widen the interlayer spacing, which facilitates partial desolvation of Zn2+ upon insertion and enhances both capacity and cycling stability.45 These lattice-level adjustments further highlight the role of interfacial field effects in regulating desolvation kinetics.
Dai et al. designed a Znx(OTf)y(OH)2x−y·nH2O (ZnOTf-LDH) interphase that functions as a selective ion-conducting layer on vanadium oxide cathodes.44 As shown in Fig. 3(g), molecular dynamics simulations reveal that Zn2+ exhibits much higher diffusivity than VO2+ or VO2+ across the ZnOTf-LDH layer, as confirmed by their distinct mean square displacements and diffusion coefficients. This selectivity stems from stronger electrostatic interactions between ZnOTf-LDH and vanadium species, which limit their mobility. Meanwhile, Zn2+ is efficiently transported and partially desolvated at the interface. Fig. 3(h) shows that Zn2+ exhibits stronger adsorption to the ZnOTf-LDH surface compared to H2O, reflecting the interphase's combined zincophilic and hydrophobic character. This facilitates Zn2+ capture while excluding water molecules and promoting the desolvation of hydrated Zn2+, as evidenced by electrochemical impedance spectroscopy (EIS) measurements at different temperatures and supporting simulations, which revealed reduced charge-transfer resistance and lower activation energy for interfacial Zn2+ transport. Cyclic voltammetry (CV) measurements further confirm that the ZnOTf-LDH layer stabilizes interfacial ion flux, suppresses vanadium dissolution, and maintains redox activity during cycling (Fig. 3(i)). As a result, the cathode retains capacity over extended operation. This interphase exemplifies how structural confinement and selective ion affinity can enhance both Zn2+ kinetics and interfacial stability.
Zheng et al.46 applied polydopamine (PDA) coatings to NH4V4O10 to create zincophilic layers that attract Zn2+, lower desolvation barriers, and inhibit parasitic reactions. In a related system, Jiang et al.47 reported PEDOT-modified V5O12·6H2O cathodes that integrate hydrophobic and zincophilic properties, enabling balanced Zn2+ interfacial behavior and robust cycling under various conditions.
These studies illustrate that regulating Zn2+ desolvation at the cathode–electrolyte interface, whether through water masking, electric field effects, or artificial interphases, offers a coherent strategy to enhance Zn2+ transport kinetics, suppress material degradation, and extend the operational lifetime of AZIBs.
![]() | ||
Fig. 4 Representative anode–electrolyte interfacial desolvation regulation strategies in AZIBs. (a) DFT-calculated desolvation energy profile for stepwise Zn2+ dehydration on pristine Zn and SAFe@MXene-Zn surfaces (SA denotes single-atom). (b) Zn2+ adsorption energies on pristine Zn, MXene-Zn, and SAFe@MXene-Zn surfaces, indicating enhanced Zn2+ affinity at the catalytic interface. Reproduced from ref. 48 Copyright (2025), with permission from American Chemical Society. (c) DFT-calculated adsorption energies of Zn(H2O)62+ cluster on pure carbon and various nitrogen-doped carbon surfaces. (d) Dissociation energy of Zn(H2O)62+ cluster in various chemical environments. Reproduced from ref. 52 under the Creative Commons CC BY license. (e) Experimental pair distribution function (PDF) G(r) (top) and simulated radial distribution function (RDF) g(r) (bottom) of 1 and 4 M ZnSO4 electrolytes. (f) Deconvolution of the stretching vibration of H2O. The pie chart shows the proportions of the local weak hydrogen-bond/W-H (high wavenumber) and strong hydrogen-bond/S-H (low wavenumber). DDAA refers to a double donor–double acceptor configuration, DDA to a double donor–single acceptor, DAA to a single donor–double acceptor, and DA to a single donor–single acceptor. (g) In situ FTIR spectra were collected during the Zn-electroplating/stripping with the variation of spectra color from blue (Zn-plating, bottom) to orange (Zn-stripping, top). (h) Evolution of the relative absorbance A(tx) − A(t0) of 1 M ZnSO4 during Zn-plating (bottom) and Zn stripping (top). Reproduced from ref. 55 Copyright (2024), with permission from American Chemical Society. |
Additional strategies for electronic-structure modulation include the use of molecular cage additives that offer multi-site Zn2+ coordination,53 and sulfonic acid-functionalized frameworks that locally tune charge density at the interface.54
![]() | ||
Fig. 5 Representative mechanisms of proton storage regulation in AZIBs. (a) CV curves of the aqueous Zn||S-MoO2 at 1.5 mV s−1 and the simultaneous response of the mass change of S-MoO2 electrode as recorded by EQCM. ZHS denotes basic zinc sulfate. (b) Mass weight versus charge curve during the discharging and charging process. Reproduced from ref. 58 Copyright (2023), with permission from Wiley-VCH. (c) CV curves of the Zn||2M ZnSO4 + 0.5 M MnSO4||ZnO cell at 0.2 mV s−1, from 0.8 to 1.8 V vs. Zn/Zn2+. Reproduced from ref. 63 Copyright (2022), with permission from Wiley-VCH. (d) Quasi-elastic neutron scattering (QENS) analysis of various hybrid aqueous-non aqueous (HANE) systems. (e) Typical voltage profile of VPO4F between 0.2 and 2.1 V in the 4 m Zn(OTf)2·H2O electrolyte. Inset: H/Zn atom ratio in the discharged VPO4F electrode. (f) Typical voltage profile of ZnxHyVPO4 between 0.2 and 2.1 V in 2 m Zn(OTf)2·2H2O–propylene carbonate (PC) electrolyte. Reproduced from ref. 69 Copyright (2021), with permission from Wiley-VCH. |
Similar Zn2+/H+ co-insertion behavior has been widely observed in hydrated vanadium oxides such as V2O5·nH2O and VO2. In these systems, the presence of pre-intercalated water layers provides expanded interlayer spacing and abundant hydrogen bonding sites, which effectively stabilize the co-inserted Zn2+ and H+ ions. For example, in V2O5·nH2O, quantitative analyses suggest that H+ contributes nearly 44% of the total capacity due to its faster insertion kinetics and smaller ionic size.59 In VO2, Li et al. demonstrated proton insertion with Zn2+ being mostly involved in surface byproduct formation.60
A unique H+ “lubrication” effect has also been proposed in MoS2-based hosts by Li et al., where pre-inserted protons significantly lower the migration barrier for subsequent Zn2+ intercalation. Theoretical calculations reveal that the Zn2+ insertion energy barrier can be reduced from 0.97 eV to 0.41 eV upon prior H+ insertion, thus enhancing Zn2+ diffusion kinetics and reversibility.61
Beyond MnO2, proton-dominant behavior extends to other hosts. For instance, recent studies on Na0.12Zn0.25V2O5·2.5H2O layered structures highlight highly dynamic proton conduction pathways in V2O5-based hosts, wherein H+ migration proceeds via Zundel pair-dancing mechanisms, enabling ultrafast pseudocapacitive response even at extremely high current densities, effectively excluding Zn2+ intercalation at ultrahigh rates.65 In parallel, in TiS2, facet engineering enabled H+-dominated storage with suppressed hydrogen evolution under (011) orientation.66 Furthermore, pH and electrolyte volume have been shown to directly influence H+ participation. Using V2O5 electrodes, Lee et al.67 demonstrated that smaller electrolyte volumes or higher pH significantly suppress H+ insertion, whereas lower pH and larger electrolyte reservoirs facilitate greater H+ contribution.
Moreover, co-doping strategies have emerged as a powerful approach to further enhance H+ insertion behavior. Chen et al.68 showed that Co2+ pre-insertion into MnO2 tunnels stabilizes the lattice, allowing more favorable H+ intercalation while excluding Zn2+ insertion. DFT calculations revealed significantly lower migration barriers for H+ compared to Zn2+ within the Co-stabilized tunnel structures, underscoring the proton-dominant storage mechanism.
It is worth noting that despite widespread discussion of “H+ insertion,” the actual origin of these protons is rarely examined in detail. At the coin cell level, quantitative analysis reveals that substantial proton involvement cannot plausibly arise from pre-existing free H+ ions in the bulk electrolyte alone. For example, in a Zn||VO2 cell using a typical lab-level cathode mass of 1–5 mg and 80–200 μL of 2 M ZnSO4 or 3 M Zn(OTf)2 electrolyte (pH ≈ 4),60,70–74 the total amount of H+ in the bulk electrolyte is only 8–20 nmol. In contrast, experimental capacities of 250–400 mAh g−1 require 4–37 μmol of proton charge storage, which exceeds the available H+ supply by more than two to three orders of magnitude. This discrepancy clearly indicates that continuous interfacial water dissociation must occur to generate additional protons during cycling. As elaborated in the following section, similar interfacial dissociation processes also underpin hydroxide-based charge storage mechanisms.
Initial evidence for OH− involvement emerged from systems where Zn2+ hydrolysis leads to the formation of solid hydroxide phases such as basic zinc sulfate (ZHS), which is particularly intriguing under mildly acidic electrolytes because the generation of seemingly alkaline compounds is not straightforward. These observations indicated that Zn2+ could polarize adjacent water molecules, triggering water dissociation and local accumulation of OH−. This process not only produces H+ that may participate in intercalation mentioned above but simultaneously generates OH− species that contribute to charge storage by being captured or fixed as solid byproducts.75,76
Moving beyond these indirect pathways, Dai et al.77 recently established a catalytic model in which Zn2+ ions polarize adjacent water molecules, thereby facilitating continuous water dissociation at the electrode–electrolyte interface, as shown in Fig. 6(a). In this model, the OH− species generated reversibly adsorbed at the electrode surface, with their storage kinetics primarily governed by the binding strength of *OH to the electrode surface. Through systematic screening of various combinations of cations and cathode materials, vanadium nitride (VN) was identified as the optimal candidate, featuring the most balanced Zn2+–H2O and VN–H2O interactions, showing a moderate *OH binding energy that ensures fast kinetics while minimizing side reactions, as illustrated in Fig. 6(b).
![]() | ||
Fig. 6 Zn2+-mediated interfacial OH− storage in VN cathodes. (a) Schematic showing Zn2+ polarization-induced water dissociation and reversible OH− adsorption at the electrode surface. (b) Contour plot of water dissociation activity as a function of *OH adsorption energy, comparing different metal ions and cathode materials. (c) Operando extended X-ray absorption fine structure (EXAFS) fitting results indicating surface-confined *OH adsorption on VN@rGO during discharge, with no lattice expansion. (d) Comparison of rate performance between water dissociation rendered OH− storage and conventional Zn2+ intercalation. Reproduced from ref. 77 Copyright (2024), with permission from Springer Nature. |
This mechanism was experimentally validated through operando (real-time and in situ) extended X-ray absorption fine structure (EXAFS) measurements, which confirmed that OH− adsorption occurs without any detectable lattice distortion in the VN@rGO electrode. This observation supports the conclusion that charge storage takes place predominantly at the surface rather than within the bulk lattice, as depicted in Fig. 6(c). In addition, electrochemical testing revealed that VN@rGO achieves a high specific capacity of 577.1 mAh g−1 even at an ultrahigh current density of 300 A g−1, significantly outperforming Zn2+-dominated systems under comparable conditions, as shown in Fig. 6(d).
These results suggest a fundamental shift in the design principles for AZIBs. Instead of relying solely on traditional ion migration through the solid phase, this approach emphasizes interfacial catalysis as the primary mechanism governing charge storage. Careful tuning of interfacial chemistry to optimize *OH adsorption enables water dissociation products, previously considered undesirable, to serve as highly effective and kinetically favorable charge carriers. It is also important to note that OH− storage mechanisms are typically sensitive to local pH fluctuations, which may trigger side reactions such as precipitation or phase instability over extended cycling.
Beyond classical Zn2+ intercalation mechanisms, recent studies have identified that products derived from interfacial water dissociation, specifically H+ and OH−, offer alternative and highly efficient charge storage pathways. The interfacial water dissociation mediated by Zn2+ polarization provides a continuous supply of H+ and OH−, enabling mechanisms such as H+-dominated intercalation and surface-confined OH− pseudocapacitance. These emerging pathways enhance battery capacity and rate capability, while redefining the role of interfacial reactions in overall performance.
Looking forward, further progress will rely on deepening mechanistic understanding of interfacial chemical dynamics under realistic operating conditions such as high mass loadings and a balanced negative-to-positive (N/P) capacity ratio. In practice, achieving a rational N/P ratio critically depends on preparing high-mass-loading cathodes, yet maintaining electrochemical performance under such conditions remains a major challenge in AZIBs.78 Rational interfacial design is therefore expected to play a pivotal role in mitigating transport limitations and enabling stable operation at high areal capacities. To fully unravel these interfacial functions under practically relevant conditions, advanced operando characterization, multiscale simulations, and interface-specific probes are essential to resolve the intricate interplay among Zn2+ transport, water dissociation dynamics, and interfacial phase evolution. Moreover, deliberate design of catalytic interfaces that precisely regulate water dissociation and capture its dissociation products offers promising opportunities to break conventional trade-offs between capacity, rate capability, and long-term stability.
While many of the interfacial strategies discussed in this review show strong potential, most remain at the laboratory or proof-of-concept stage. Further studies on scalability, compatibility, and long-term stability will be essential to assess their commercial relevance and guide the transition toward practical applications.
If successfully translated into practical systems, interfacial regulation holds great promise to enable unprecedented levels of fast charging, long cycling life, and real-world viability for large-scale energy storage using AZIBs. A conceptual roadmap outlining promising interfacial design trends over the next 5–10 years is presented in Fig. 7 to guide future research priorities. Furthermore, the insights and strategies reviewed herein are expected to provide valuable guidance for advancing not only other multivalent-ion battery systems (where ion size, solvation structures, and interfacial phenomena similarly dictate performance) but also conversion-type chemistries such as Zn–S and Zn–I2 batteries, in which interfacial storage processes play an increasingly dominant role.
Footnote |
† These authors contribute equally. |
This journal is © The Royal Society of Chemistry 2025 |