Next Article in Journal
Fracture Initiation in Aluminum Alloys Under Multiaxial Loading at Various Low Strain Rates
Previous Article in Journal
Rotational Bending Fatigue Crack Initiation and Early Extension Behavior of Runner Blade Steels in Air and Water Environments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Grain Boundary Engineering for Reversible Zn Anodes in Rechargeable Aqueous Zn-Ion Batteries

1
State Environmental Protection Key Laboratory of Eco-Industry, Northeastern University, Shenyang 110819, China
2
Engineering Research Center of Frontier Technologies for Low-Carbon Steelmaking (Ministry of Education), School of Metallurgy, Northeastern University, Shenyang 110819, China
*
Authors to whom correspondence should be addressed.
Metals 2025, 15(7), 784; https://doi.org/10.3390/met15070784
Submission received: 30 April 2025 / Revised: 27 June 2025 / Accepted: 9 July 2025 / Published: 11 July 2025

Abstract

Rechargeable aqueous zinc-ion batteries (AZIBs) have garnered significant research attention in the energy storage field owing to their inherent safety, cost-effectiveness, and environmental sustainability. Nevertheless, critical challenges associated with zinc anodes—including dendrite formation, hydrogen evolution corrosion, and mechanical degradation—substantially impede their practical implementation. Grain boundary engineering (GBE) emerges as an innovative solution for zinc anode optimization through the precise regulation of grain boundary density, crystallographic orientation, and chemical states in metallic materials. This study comprehensively investigates the fundamental mechanisms and application prospects of GBE in zinc-based anodes, providing pivotal theoretical insights and technical methodologies for designing highly stable electrode architectures. The findings are expected to promote the development of aqueous zinc batteries toward a high energy density and long cycle life.

1. Introduction

The prolonged global reliance on fossil fuels has led to severe climate issues, with the accelerating trend of global warming posing an urgent threat. Furthermore, existing fossil fuel-based energy infrastructure fails to meet the demands of sustainable societal development, presenting significant challenges in enhancing energy supply capacity and security [1]. Consequently, the global energy transition has become an imperative goal for nations worldwide. As the energy landscape shifts toward cleaner, more environmentally friendly, and low-carbon solutions, efficient and safe energy storage technologies have emerged as a critical enabler for large-scale renewable energy integration. Currently, lead-acid and nickel-cadmium batteries dominate the energy storage market due to their low cost and durability. However, their limitations such as their low energy density and environmental concerns stemming from toxic electrodes hinder their long-term viability [2]. In contrast, lithium-ion batteries (LIBs) have garnered widespread attention for their lightweight design, high energy density, and superior specific power. Unfortunately, LIBs also suffer from drawbacks, including scarcity of lithium resources, energy density limitations, high costs, and safety risks associated with toxic and flammable organic electrolytes [3,4]. Aqueous zinc-ion batteries (AZIBs) have emerged as a promising alternative to LIBs, distinguished by their inherently safe aqueous electrolytes [5], Table 1 compiles and contrasts the fundamental properties of LIBs and AZIBs, high-lighting their respective advantages and limitations. In comparison to alternative metal anodes, including Li, Na, Mg, and Al (detailed in Table 2), zinc demonstrates distinct advantages including lower redox potential, optimal ionic radius, and superior volumetric capacity. Moreover, zinc resources are abundant, with identified global reserves exceeding 1.9 billion tons [6]. Additional advantages, such as their low cost and environmental friendliness, position AZIBs as one of the most promising electrochemical energy storage systems, demonstrating significant commercial potential for grid-scale applications [7].
As a zinc-based secondary battery, AZIBs typically employ metallic zinc as the anode, a weakly acidic aqueous electrolyte containing Zn2+ ions [20,21], and cathode materials with multi-ion channels (e.g., tunnel or layered structures with large interlayer spacing) [22]. Recent breakthroughs in cathode materials—including manganese-based compounds, vanadium-based compounds, MXenes (2D transition metal carbides/nitrides), Prussian blue analogs (PBAs), and organic compounds which have expanded the possibilities for AZIBs [23]. Future research will continue to clarify the ion insertion mechanism and develop advanced anode materials [24].
However, the practical application of AZIB anodes faces critical challenges (Figure 1). Non-uniform Zn2+deposition/dissolution leads to irregular surface morphology and zinc dendrite growth, a phenomenon known as the “tip effect” [25]. During initial nucleation, Zn2+ tends to deposit in regions with higher ion concentrations. Subsequent zinc nuclei, driven by high surface energy, preferentially grow on existing protrusions [26], eventually forming sharp, needle-like dendrites. These dendrites can detach during cycling, penetrate separators, and cause internal short circuits or battery failure [27], severely curtailing cycle life. Cheng et al. [28] developed a phase-field model to study the effects of applied voltage and Zn2+ concentration gradients on dendrite growth. Their simulations revealed that beyond a critical overpotential, zinc deposition transitions into dendritic morphology. Dendrite growth kinetics depend on localized overpotential, which is influenced by the anode-cathode voltage difference. Higher overpotentials accelerate dendrite initiation and propagation. While mild overcharging in aqueous systems can partially dissolve dendrites via oxygen-induced reactions, this process consumes water from the electrolyte and generates passivating oxide layers on the zinc surface. Thus, developing effective dendrite suppression strategies is crucial.
Another major challenge for AZIBs involves hydrogen evolution reactions (HERs) and corrosion. These parasitic processes arise from potential differences across the heterogeneous zinc electrode surface, forming numerous micro-corrosion cells [30]. HER generates gas bubbles, disrupting solid–liquid contact and increasing polarization voltage [31,32], often leading to battery swelling (“bulging”). Furthermore, HER and corrosion produce insoluble byproducts that passivate the electrode surface [33]. Collectively, these issues degrade cycling performance and impede the large-scale deployment of AZIBs. Addressing them is essential to realizing AZIBs as a viable, next-generation energy storage solution.
In recent years, GBE has demonstrated encouraging success in lithium alloy anodes, offering a novel strategy for modifying AZIB anodes [34,35]. As a critical microstructural feature in polycrystalline materials, grain boundaries (GBs), the fundamental structural units connecting crystalline grains, directly influence the mechanical properties, electrochemical activity, and ion transport behavior of metals through their density, orientation, and chemical states. Compared to crystalline planes, GBs exhibit heightened reactivity [36]. Studies reveal that localized corrosion typically initiates at GBs before propagating into grain interiors, ultimately leading to irreversible intergranular corrosion [37,38]. Consequently, tailoring GB characteristics can significantly enhance the uniformity of zinc deposition. High-density GBs provide abundant nucleation sites, suppressing dendritic preferential growth; controlled GB orientations reduce diffusion energy barriers for zinc atoms, promoting planar deposition; alloying element doping modulates GB energy states, mitigating parasitic reactions. Furthermore, synergistic integration of grain boundary engineering (GBE) with surface modification and electrolyte additive strategies has substantially improved the overall performance of AZIB anodes.
This paper systematically reviews the application mechanisms and research progress of grain boundary engineering in AZIB anodes. It begins by elaborating on the fundamental theory of grain boundaries and the strategies/methods of grain boundary engineering, followed by a focused analysis of its optimization effects on pure zinc and zinc alloy anodes specifically in suppressing dendrite growth, mitigating hydrogen evolution reactions, and enhancing mechanical properties. The study further explores the synergistic effects between grain boundary engineering and other modification strategies, while providing insights into future research directions. By establishing a correlation framework of “microstructural design-macroscopic performance enhancement,” this work aims to offer theoretical guidance and technical references for developing high-performance zinc anodes, thereby advancing the practical application of aqueous zinc-ion batteries.

2. Fundamentals of Grain Boundary Engineering

As a critical structural feature of polycrystalline materials, GBs exhibit disordered atomic arrangements and high-energy states that profoundly influence not only mechanical properties but also the stability of electrochemical interfaces. In AZIBs, the failure mechanisms of Zn anodes—including dendrite growth and hydrogen evolution corrosion—are intrinsically linked to GB characteristics. These boundaries serve as preferential pathways for rapid Zn2+ migration while simultaneously acting as active sites for defect accumulation and parasitic side reactions. A fundamental understanding of GBs is essential for modulating their properties, and targeted optimization of grain boundary character distribution (GBCD) through GBE enables atomic-scale reconstruction of the interfacial microenvironment in Zn anodes. The following sections will systematically elucidate the structural nature and classification of GBs, laying the theoretical groundwork for subsequent discussions on how GBE strategies, such as dendrite suppression and corrosion mitigation can enhance the performance of Zn-based anodes.

2.1. Basic Concepts of Grain Boundaries

Most crystalline materials are composed of multiple grains, where the interface between grains of the same solid phase but different orientations is termed a GB. The atomic arrangement at GBs differs from the regular lattice within grains, typically exhibiting greater disorder. Consequently, GB atoms possess higher energy than their intragranular counterparts, making GBs a type of planar defect that is extensive in two dimensions but confined in the third dimension.
GBs serve as fast diffusion pathways for atoms and are prone to segregation of impurity/solute atoms and the precipitation of secondary phases. As a critical microstructural feature in polycrystalline materials, GBs significantly influence material properties, including intergranular corrosion resistance [39,40], intergranular stress corrosion cracking behavior [41,42], and creep performance [43,44].
To describe GB geometry, both the boundary orientation and the relative crystallographic orientation of adjacent grains must be specified. In 2D lattices, a GB’s position can be defined by: The misorientation angle θ between two grains, The inclination angle ϕ between the GB plane and a reference lattice plane. For 3D lattices, five degrees of freedom are required: Three parameters define the misorientation between adjacent grains; two parameters determine the crystallographic orientation of the GB plane.
GBs are typically classified by the misorientation angle θ between constituent grains. When disregarding special symmetries, misorientation is conventionally expressed using the minimum-angle–axis-angle pair <uvw>/θ, representing a θ-degree rotation about the <uvw> crystal axis. Based on θ values: low-angle GBs (θ < 10°): predominant in subgrain structures, high-angle GBs (θ > 10°): more common in polycrystals.
Low-angle grain boundaries can be further classified based on their misorientation characteristics, Figure 2 shows the formation processes of different types of low-angle grain boundaries. Symmetric tilt boundaries consist of parallel arrays of edge dislocations, representing the simplest low-angle boundary configuration. More complex asymmetric tilt boundaries form through the intersection of non-parallel edge dislocations. A distinct class of low-angle boundaries, known as twist boundaries, results from the relative rotation of two crystal lattices about an axis perpendicular to the boundary plane. The presence of these low-angle boundaries directly influences the material’s dislocation density, typically leading to improved mechanical properties including enhanced strength, hardness, and ductility when compared to single crystal materials of identical composition. In contrast to the well-defined structure of low-angle boundaries, high-angle grain boundaries exhibit substantially more complex atomic arrangements characterized by non-periodic structures that cannot be adequately described using conventional dislocation models, making their fundamental understanding significantly more challenging.

2.2. Overview of Grain Boundary Engineering

The concept of GBE originated in 1984 when Watanabe from Tohoku University proposed the hypothesis of “grain boundary design and control” during his investigation of intergranular cracking [45], based on the coincidence site lattice (CSL) model and extensive experimental studies. In the 1990s, Palumbo et al. [46] further advanced this concept by demonstrating that increasing the proportion of “special grain boundaries” could significantly enhance a material’s resistance to intergranular stress corrosion cracking. They emphasized that this effect only became pronounced when the fraction of special grain boundaries exceeded a critical threshold, thereby establishing GBE as a distinct research field. Fundamentally, GBE aims to optimize material performance by controlling GBCD through thermomechanical processing, with the dual objectives of increasing the proportion of low-ΣCSL boundaries (where Σ ≤ 29) and disrupting the connectivity of random high-angle grain boundary networks.
According to the CSL classification theory, grain boundaries can be categorized into low-ΣCSL boundaries and general high-angle boundaries. Low-Σ CSL grain boundaries, defined by a high fraction of coincident lattice sites at particular misorientation angles between neighboring grains, represent a distinct category of crystallographic interfaces with reduced interfacial energy and enhanced structural periodicity [47]. Research has shown that low-ΣCSL boundaries, particularly Σ3 boundaries, exhibit exceptional functional properties due to their lower free energy. These boundaries demonstrate superior resistance to various degradation phenomena including corrosion, stress corrosion cracking [48], sensitization [49], and elemental segregation [50]. In contrast, general high-angle boundaries, characterized by higher energy and greater mobility, often serve as preferential pathways for crack nucleation and propagation, leading to intergranular corrosion or stress corrosion failures. This fundamental difference has driven GBE research to focus on manipulating the fraction of low-ΣCSL boundaries to optimize GBCD.
Current GBE strategies primarily target face-centered cubic (FCC) metals with medium-to-low stacking fault energy, such as copper alloys [51] and nickel-based alloys [52]. These materials readily form annealing twins (corresponding to Σ3 boundaries) during thermomechanical processing through recovery and recrystallization, while simultaneously promoting the generation of geometrically related Σ9 and Σ27 boundaries [53]. By carefully optimizing deformation and annealing parameters, the proportion of low-ΣCSL boundaries can be significantly increased, thereby disrupting the percolation of random boundary networks. This twin-induced boundary control approach not only enables targeted optimization of grain boundary structure but also provides an effective pathway for enhancing comprehensive material properties, including corrosion resistance and crack propagation resistance.

2.3. Strategies and Methods of Grain Boundary Engineering

GBE is typically achieved through deformation and heat treatment processes to optimize the GBCD of materials [54]. This requires precise control of key processing parameters including deformation degree [55], annealing temperature [56], annealing duration [57], and processing passes. Based on process characteristics, GBE methods can be classified into two primary categories: strain-recrystallization processes and strain-annealing processes.
The strain-recrystallization approach encompasses both iterative strain-recrystallization and single-step strain-recrystallization techniques. The iterative method involves repeated cycles (typically 3–7 times) of cold rolling deformation (20–30%) followed by short-duration annealing (≤20 min) at elevated temperatures (≥0.7 Tm), ultimately yielding a fine-grained, texture-free microstructure. For instance, this process has been shown to significantly increase the proportion of Σ3 and Σ9 boundaries in nickel-based alloys (Ni-16Cr-9Fe). In contrast, the single-step variant employs moderate deformation coupled with brief high-temperature annealing (1–2 min), effectively enhancing twin boundary density while maintaining fine grain structure [58].
Strain-annealing processes similarly include both iterative and single-step implementations. The iterative approach alternates between minor deformation (2–7%) and short high-temperature annealing, or alternatively combines small deformations with prolonged low-temperature annealing (6–14 h) over multiple cycles. The single-step version utilizes limited deformation (5–10%) followed by extended low-temperature annealing (typically tens of hours) to optimize GBCD. Research shows that single-step strain-annealing promotes dislocation slip and climb, driving grain boundaries toward lower-energy configurations while effectively disrupting random high-angle boundary networks and improving intergranular corrosion resistance [59]. Applied to 304 stainless steels [60], this method increased the low-ΣCSL boundary fraction from 63% to 85% in the base material.
While both process categories effectively enhance special boundary fractions, their operational emphases differ substantially. Randle et al. [52] found that single-pass thermomechanical processing more efficiently increases low-ΣCSL boundary proportions and improves boundary-related properties, with minor single-pass deformation proving particularly effective in disrupting boundary connectivity. Conversely, multi-pass processing facilitates development of more stable boundary structures. Further research by Randle’s group [61] revealed that multi-pass treatments can additionally optimize low-ΣCSL boundary distributions and more thoroughly disrupt random boundary networks. This enhancement stems from deformation energy storage at annealing twins formed during initial processing, which subsequently provides the driving force for boundary migration and reaction, thereby promoting additional low-ΣCSL boundary formation. Consequently, material-specific thermomechanical processing strategies must be carefully selected based on application requirements to maximize special boundary fractions and resultant material performance [47,62].

3. Application of Grain Boundary Engineering in Zinc Anodes

Grain boundary engineering provides a multiscale solution to address critical challenges in zinc anodes, including dendrite formation and parasitic reactions, through precise control of crystallographic orientation, grain boundary density, and spatial distribution (Figure 3). In pure zinc electrodes, preferential enhancement of (002) plane orientation significantly reduces zinc deposition energy barriers while effectively suppressing hydrogen evolution corrosion, whereas the introduction of high-density grain boundaries promotes uniform nucleation for dendrite-free deposition. The zinc alloying approach further extends these advantages by leveraging synergistic interactions between solute atoms and grain boundaries, simultaneously stabilizing boundary structures and modulating interfacial charge distribution to achieve superior cycling stability across extended electrochemical windows. When synergistically combined with complementary strategies such as surface coating and electrolyte engineering, grain boundary engineering demonstrates remarkable versatility in anode optimization. These collective findings establish a solid scientific foundation for developing high-performance zinc anodes.

3.1. Grain Boundary Engineering in Pure Zinc Electrodes

The crystallographic orientation of electrode surfaces critically governs metal deposition pathways, highlighting the importance of uniform surface crystallinity for dendrite-free metal plating [63]. Zinc metal, possessing a hexagonal close-packed (HCP) structure, demonstrates significant crystallographic dependence in deposition behavior, with the (002) and (001) planes exhibiting particularly strong texture coefficients. Notably, the (100) plane’s heterogeneous interfacial charge distribution readily induces dendritic growth [64], while the (002) plane facilitates uniform Zn deposition due to its homogeneous charge distribution (Figure 4a). Furthermore, comparative analysis reveals the (002) plane’s superior thermodynamic stability, exhibiting more positive free energies for both hydrogen evolution (Figure 4b) and stripping energy (Figure 4c), indicative of stronger chemical bonding that enhances corrosion resistance. These characteristics collectively enable the (002) plane to promote parallel Zn plating into lamellar structures, whereas the (100) plane preferentially induces both corrosion and dendritic formations (Figure 4d) [65]. Consequently, zinc anodes with exclusive (002) crystallographic orientation have emerged as a premium strategy for high-performance battery systems [66,67].
Research has demonstrated that precisely controlled annealing processes can effectively modulate the crystallographic orientation of zinc anodes. Wang et al. [68] developed a transformative approach by annealing commercial zinc foil at 300 °C for 60 min under 5% Ar/H2 atmosphere, leveraging recrystallization to engineer (002)-textured foils (Figure 4e). Subsequent immersion in ZnSO4 electrolyte spontaneously formed a dense zinc hydroxide sulfate (ZHS) layer [69], functioning as an artificial solid-electrolyte interphase (SEI) with high Zn2+ conductivity but negligible electronic conduction. Electrochemical evaluation demonstrated exceptional performance: (002)-Zn symmetric cells achieved over 7000 stable cycles, while full cells paired with MnO2 cathodes maintained 92.7% capacity retention with 99.9% average Coulombic efficiency. This breakthrough originates from dual mechanisms: the (002) plane’s low self-diffusion barrier enables two-dimensional Zn migration and lateral deposition, while the ZHS layer physically isolates zinc from electrolyte, synergistically suppressing both dendrite formation and parasitic corrosion reactions.
Chen et al. [70] developed an innovative rapid melting–solidification technique to transform commercial zinc foil into millimeter-scale single (002)-textured zinc (designated as HG-002-Zn), significantly enhancing anode stability and cycling performance. The process involved heating zinc foil to 600 °C for 1 min to achieve complete melting, followed by natural cooling to room temperature (Figure 4f). Thermodynamically driven by the (002) plane’s minimal surface energy [71], this method spontaneously produced (002)-oriented crystallites. Electrochemical evaluation revealed HG-002-Zn’s exceptional performance: symmetric cells demonstrated stable cycling for over 3280 h at 1 mA cm−2 and 830 h at 10 mA cm−2. Even under demanding 75% depth of discharge (DOD) conditions, HG-002-Zn maintained stable operation for 180 h, outperforming commercial zinc foil’s 25 h lifespan. Full-cell configurations with MnO2 cathodes exhibited 74% capacity retention after 1500 cycles at 1 A g−1, contrasting sharply with the 40% retention of untreated zinc foil.
Further investigation by Chen’s group [72] established the critical relationship between initial grain size and (002) texture development during cold rolling. Through optimized annealing (350 °C, 10 h), they controlled commercial foil’s grain structure to approximately 30 μm while minimizing sub-10 μm grains. Subsequent cold rolling effectively induced (002) texture formation while substantially increasing grain size and reducing grain boundary density, thereby decreasing parasitic reactions. The processed foil achieved remarkable stability, sustaining over 2800 h of cycling in symmetric cells at 0.1 mA cm−2.
Zhang et al. [73] proposed and fabricated a [0001]-uniaxial oriented Zn metal anode exhibiting exclusive (0002) texture. Utilizing a direct current electrodeposition technique with pure zinc as the starting material and copper foil as the substrate, the researchers successfully prepared Zn anodes with a single (0002) texture in a zinc sulfate solution containing boric acid. Through precise control of electrolyte composition and deposition parameters, the oriented growth of Zn crystals along the (0002) plane was achieved. The electrodeposited [0001]-uniaxial oriented Zn metal anodes with a single (0002) texture fundamentally eliminate the lattice mismatch and achieve ultra-sustainable homoepitaxial growth. Notably, the Zn (0002) plane demonstrates a remarkably low surface energy of 0.84 J/m2, significantly lower than other crystallographic planes. Experimental results demonstrate exceptional electrochemical performance: even at a high DODZn of 75.2%, the Zn (0002) symmetric cells maintain stable operation for 250 h. Furthermore, when assembled in Zn||NH4V4O10 pouch cells, these anodes deliver a high specific capacity of 220 mAh g−1 for over 450 cycles with an impressive capacity retention of 80%. The fundamental insights into the [0001]-oriented Zn metal anode and its persistent homoepitaxial growth mechanism presented in this work provide a valuable framework for developing other highly reversible metal electrodes.
The controlled crystal growth engineering demonstrates that precisely oriented single-crystalline zinc anodes fundamentally address the challenges of dendritic growth and interfacial instability while exhibiting superior electrochemical performance compared to conventional polycrystalline zinc counterparts. Shen et al. [74] successfully grew single-crystalline Zn with preferential [0001] orientation using the Bridgman method, obtaining a Zn(002) single-crystal anode. This single-crystal Zn (002) anode, featuring a perfect atomic arrangement and exclusively low-energy facets, not only enables epitaxial Zn deposition to suppress dendrite formation kinetically but also maintains thermodynamic stability to minimize side reactions. In contrast, polycrystalline Zn (Zn(poly)) anodes exhibit randomly oriented facets with varying surface energies, leading to inhomogeneous Zn2+ deposition and dendritic growth. Electrochemical tests demonstrated exceptional performance: Zn(002)‖Zn(002) symmetric cells achieved ultralong cycling stability over 2800 h at 1 mA cm−2 and 1 mAh cm−2, over 4800 h at 2 mA cm−2 and 0.5 mAh cm−2, and over 700 h at 5 mA cm−2 and 2 mAh cm−2, while Zn(poly)‖Zn(poly) cells failed rapidly under identical conditions with significant voltage polarization. The Zn(002)‖Cu asymmetric cell maintained an average Coulombic efficiency of 99.92% over 500 cycles at 10 mA cm−2 and 10 mAh cm−2, far outperforming its polycrystalline counterpart. In full-cell tests with MnVO cathode, the Zn(002)‖MnVO cell delivered higher capacity (302 vs. 295 mAh g−1 at 2 A g−2) with lower polarization and faster kinetics, and maintained superior performance (190 vs. 180 mAh g−1) even at 8 A g−1.
Complementing crystallographic orientation approaches, Lian et al. [75] employed femtosecond laser processing (500 fs pulse width) to generate ultrahigh-proportion grain boundaries (UP-GBs) via Coulomb explosion-induced electron emission (Figure 4g). These engineered boundaries provided abundant nucleation sites, facilitating uniform zinc deposition and dendrite suppression [76,77]. The femtosecond-treated zinc (Fs-Zn) demonstrated unprecedented performance metrics: symmetric cells achieved 6.5 Ah cm−2 cumulative plating capacity (versus 1.2 Ah cm−2 for commercial foil), with stable voltage profiles at 1 mA cm−2/1 mAh cm−2 (0.548 Ah cm−2 cumulative capacity). Full cells paired with MnO2 cathodes exhibited excellent rate capability (0.1–2 A g−1) and cycling stability, showcasing the versatility of grain boundary engineering approaches for advanced zinc anodes.
Figure 4. (a) The surface atomic arrangement and electron equipotential plane of different crystal planes (reprinted with permission from Ref. [65]; 2025 Elsevier). (b) The free-energy diagram for HER of (002) and (100) planes (reprinted with permission from Ref. [65]; 2025 Elsevier). (c) The stripping energy of Zn from (002) and (100) planes (reprinted with permission from Ref. [65]; 2025 Elsevier). (d) Schematic illustration of the nucleation and growth process of (002) and (100) planes (reprinted with permission from Ref. [65]; 2025 Elsevier). (e) Fabrication of (002)-Zn electrode via annealing in Ar/H2 atmosphere (reprinted with permission from Ref. [68]; 2025 Elsevier). (f) Schematic illustration of fast melting–solidification approach for single (002)-textured Zn preparation (reprinted with permission from Ref. [70]; 2024 Wiley). (g) The principle diagram of constructing UP-GBs by Fs laser bombarding (reprinted with permission from Ref. [75]; 2024 Wiley).
Figure 4. (a) The surface atomic arrangement and electron equipotential plane of different crystal planes (reprinted with permission from Ref. [65]; 2025 Elsevier). (b) The free-energy diagram for HER of (002) and (100) planes (reprinted with permission from Ref. [65]; 2025 Elsevier). (c) The stripping energy of Zn from (002) and (100) planes (reprinted with permission from Ref. [65]; 2025 Elsevier). (d) Schematic illustration of the nucleation and growth process of (002) and (100) planes (reprinted with permission from Ref. [65]; 2025 Elsevier). (e) Fabrication of (002)-Zn electrode via annealing in Ar/H2 atmosphere (reprinted with permission from Ref. [68]; 2025 Elsevier). (f) Schematic illustration of fast melting–solidification approach for single (002)-textured Zn preparation (reprinted with permission from Ref. [70]; 2024 Wiley). (g) The principle diagram of constructing UP-GBs by Fs laser bombarding (reprinted with permission from Ref. [75]; 2024 Wiley).
Metals 15 00784 g004

3.2. Application of Grain Boundary Engineering in Zinc Alloy Electrodes

Recent years have witnessed growing interest in alloying strategies as an effective approach for zinc anode optimization [78]. Various zinc-based alloys, including Zn-Cu [79], Zn-Al [80], and Zn-In systems [81], have demonstrated remarkable success in regulating zinc deposition behavior and corrosion mechanisms, enabling highly reversible zinc electrodes.
Zhao et al. [82] developed a Zn-Ti alloy (99.5:0.5 weight ratio) where titanium promotes the formation of TiZn16 intermetallic compounds (IMCs) at grain boundaries. The preferential distribution of these IMCs at grain boundaries significantly reduces boundary activity, suppresses intergranular corrosion, and improves zinc nucleation and growth behavior, resulting in uniform and dense zinc deposition, as confirmed by computational simulations. The engineered alloy exhibited exceptional stability, with experimental results showing that symmetric cells employing the Zn-Ti alloy achieve stable cycling for over 1100 h at 2 mA cm−2 and 2 mAh cm−2, four times longer than those using pure Zn. In the Zn-Ti//NH4V4O10 full-cell configuration, the Zn-Ti alloy anode demonstrates exceptional cycling stability over approximately 3500 cycles while maintaining an average operating voltage above 3.3 V. Furthermore, the scaled-up 34 mAh multilayer pouch cell exhibits remarkable stability through 500 cycles, retaining nearly 85% of its initial capacity.
Moreover, straightforward chemical displacement reactions can effectively achieve alloying doping objectives. Li’s group [83] employed a chemical replacement method to create Sn@Zn alloys, achieving planar zinc deposition morphology (Figure 5a) compared to dendritic growth on pristine zinc. The alloy’s elevated hydrogen evolution overpotential effectively mitigated gas evolution and battery swelling. Sn@Zn symmetric cells achieved 1500 h lifespans at 1 mA cm−2/1 mAh cm−2, while full cells with CNT@MnO2 and NH4V4O10 cathodes delivered 150 mAh g−1 (0.5 A g−1, 300 cycles) and 212 mAh g−1 (5 A g−1, 1000 cycles), respectively.
Rare-earth-element doping confers unique advantages for grain boundary engineering, such as with cerium. Chen et al. [84] introduced cerium into zinc grain boundaries via electrodeposition-assisted Zener pinning, creating an ultra-refined alloy layer (URAL) that reduced grain size from 3 to 1 μm (Figure 5b). Scanning Electron Microscopy(SEM)/Electron Backscatter Diffraction(EBSD) analysis revealed exceptionally uniform deposition on ZnCe electrodes, enabling unprecedented 4000 h symmetric cell operation without short-circuiting at 2 mA cm−2/2 mAh cm−2. The ZnCe||NH4V4O10 full cell maintained 96% capacity after 6000 cycles, with Linear Sweep Voltammetry (LSV) and insitu pH monitoring confirming superior HER suppression through Ce segregation at boundaries.
Further advancing this approach, Chen’s team [85] utilized indium’s low HER activity and melting point to create Zn-In alloys via 450 °C thermal processing, where liquid indium penetrated (002)-textured zinc boundaries (Figure 5c,d). DFT calculations by Yuan et al. [81] revealed indium’s strong zinc adsorption energy, enabling uniform nucleation. Liu et al. [86] extended this strategy by designing Ga-In-Zn ternary coatings based on phase diagram analysis, further enhancing anode stability and lifespan through synergistic grain boundary engineering.

3.3. Synergistic Integration of Grain Boundary Engineering with Multi-Factor Strategies

Recent advances have demonstrated the efficacy of combining grain boundary engineering with complementary approaches to further enhance zinc anode performance. Li et al. [87] pioneered an electrolyte additive strategy using sodium gluconate (SG) in 2 M ZnSO4 electrolyte (10–100 mM concentrations). Their systematic study revealed that SG molecules preferentially adsorb on Zn (100) and (101) planes, forming a protective layer that redirects deposition to the (002) plane (Figure 6a). This crystallographic control yielded remarkably uniform plating morphology, enabling Zn||Zn symmetric cells with SG20 electrolyte to achieve 2000 h cycling at 1 mA cm−2/1 mAh cm−2 and 900 h operation at 5 mA cm−2/2.5 mAh cm−2. The NH4V4O10||Zn full cell maintained 95% capacity retention after 2000 cycles at 2 A g−1 with near 100% Coulombic efficiency.
Building on this work, the same group [88] investigated L-leucine (Leu) additives, discovering stronger adsorption on (100)/(101) versus (002) planes (Figure 6b,c). The Leu20 formulation not only promoted (002)-oriented deposition but also suppressed parasitic reactions, as evidenced by the absence of ZHS byproducts in post-cycled anodes through X-ray Diffraction (XRD)/Energy-Dispersive X-ray Spectroscopy (EDS) analysis.
Parallel developments in interfacial engineering have yielded complementary solutions. Chen’s team [89] demonstrated that polyvinyl alcohol (PVA) coatings simultaneously enhance corrosion resistance and induce (002)-preferential deposition. The PVA@Zn anode exhibited intensified (002) diffraction after plating, correlating with dendrite-free operation for over 5000 h.
Addressing mechanical aspects, Chen et al. [90] developed an innovative microgroove architecture (30 μm width/25 μm depth) via metal mesh-assisted calendaring, combined with Nafion coating (N@P-Zn). This dual-functional design effectively releases plating-induced stress while minimizing electrolyte contact (Figure 6d,e). The N@P-Zn||MnO2 full cells delivered 186 mAh g−1 capacity with exceptional cycling stability and mechanical flexibility, establishing new benchmarks for practical applications.
These findings reveal that GBE, through synergistic integration with electrolyte regulation and interfacial coatings, can substantially enhance the performance boundaries of zinc anodes. To systematically and comprehensively evaluate the effectiveness of various modification approaches, Table 3 compares key electrochemical performance metrics across four critical aspects: grain boundary density control, crystallographic orientation optimization, alloy design, and multifactor coupling strategies. This comprehensive analysis provides experimental evidence supporting the refinement of the “microstructure-to-macroscopic performance” design framework.
Table 3. Comparative evaluation of grain boundary engineering strategies for zinc anodes.
Table 3. Comparative evaluation of grain boundary engineering strategies for zinc anodes.
StrategyMethodologyKey Performance MetricsEffectiveness EvaluationRefs.
Grain Boundary Engineering in Pure Zinc ElectrodesAnnealing commercial Zn foil at 300 °C for 60 min under 5% Ar/H2 atmosphere.Symmetric cell cycling: >7000 cycles;
Full-cell capacity retention: 92.7% (MnO2 cathode);
Coulombic efficiency: 99.9%.
Effectively suppresses dendrite growth and HER corrosion, significantly enhancing cycling stability.[68]
Melting Zn foil at 600 °C for 1 min followed by natural cooling.Symmetric cell cycling: 3280 h (1 mA cm−2);
High-rate performance: 830 h (10 mA cm−2);
75% DOD cycling: 180 h.
Reduces grain boundary density, mitigates side reactions, and improves deep cycling capability.[70]
Mechanical grinding to control crystallographic orientation.75.2% DOD cycling: 250 h (symmetric cell);
Full-cell capacity retention: 80% (450 cycles).
Low-cost, scalable method achieving high reversibility.[73]
Bridgman method to grow single-crystalline Zn with [0001] orientation.Symmetric cell cycling: >2800 h (1 mA cm−2);
- Coulombic efficiency: 99.92% (500 cycles).
Perfect atomic arrangement inhibits dendrites and side reactions, outperforming polycrystalline Zn.[74]
Coulomb explosion via femtosecond laser (500 fs pulse width) to create UP-GBs.Cumulative plating capacity: 6.5 Ah cm−2 (vs. 1.2 Ah cm−2 for commercial foil);
Stable cycling: 0.548 Ah cm−2 (1 mA cm−2).
Provides abundant nucleation sites for uniform deposition, effectively suppressing dendrites.[75]
Grain Boundary Engineering in Zinc Alloy ElectrodesZn-Ti Alloy (TiZn16 at GBs): Melting Zn with 0.5 wt% Ti to form TiZn16 intermetallic compounds at GBs.Symmetric cell cycling: 1100 h (2 mA cm−2);
Full-cell capacity retention: 85% (500 cycles).
Reduces GB reactivity, suppresses corrosion, and promotes uniform nucleation.[82]
Zn-Sn Alloy (Sn@Zn): Chemical displacement to create Sn-based heterogeneous nucleation sites on Zn surfacesSymmetric cell cycling: 1500 h (1 mA cm−2);
Full-cell capacity: 212 mAh g−1 (5 A g−1).
Sn increases HER overpotential, prevents battery swelling, and enables planar deposition.[83]
Zn-Ce Alloy (Ce Doping): Electrodeposition-assisted Zener pinning to introduce Ce, forming URAL.Symmetric cell cycling: 4000 h (2 mA cm−2);
Full-cell capacity retention: 96% (6000 cycles).
Ce refines grains to ~1 μm, suppresses HER, and enhances cycling stability.[84]
Zn-In Alloy (In Infiltration): Thermal treatment at 450 °C to infiltrate liquid In into (002)-textured Zn GBs.DFT calculations confirm high Zn adsorption energy on In, enabling uniform nucleation.In stabilizes GBs against corrosion and dendrite formation.[85]
Ga-In-Zn Ternary Alloy Coating: Phase diagram-guided design of liquid alloy coating.Extends cycling life and improves anode stability.Synergistically optimizes GB and surface properties for enhanced performance.[86]
Multi-Factor Coupling StrategiesElectrolyte Additive: Adding SG (20 mM) to 2 M ZnSO4 to induce (002)-oriented deposition.Symmetric cell cycling: 2000 h (1 mA cm−2);
Full-cell capacity retention: 95% (2000 cycles).
SG selectively passivates non-(002) facets, suppressing dendrites while maintaining high-rate capability.[87]
Electrolyte Additive: Adding Leu (20 mM) to 2 M ZnSO4 to inhibit side reactions.No ZHS byproducts detected post-cycling;
High Coulombic efficiency.
Leu preferentially adsorbs on high-energy facets, reducing corrosion and passivation.[88]
Coating Zn anode with PVA layer.Cycling life: >5000 h;
Enhanced corrosion resistance.
PVA isolates Zn from electrolyte while promoting (002)-oriented deposition.[89]
Fabricating microgrooves (30 μm width/25 μm depth) via mesh-assisted calendaring, followed by Nafion coating.Full-cell capacity: 186 mAh g−1 (MnO2 cathode);
Excellent flexibility.
Microgrooves relieve plating stress; Nafion minimizes electrolyte contact, suitable for flexible devices.[90]
Figure 6. (a) Schematic illustration of zinc electrodeposition in 2 M ZnSO4 electrolyte (reprinted from [87]). (b,c) Schematic illustrations of Zn electrodeposition in different electrolytes: (b) 2 M ZnSO4 and (c) 2 M ZnSO4 with the presence of 20 mM Leu molecules (reprinted with permission from Ref. [88]; 2025 Royal Society of Chemistry). (d,e) Schematic illustration of Zn-plating-induced stress and its effect on Zn plating/stripping behavior on PL-Zn (d) and P-Zn (e) (reprinted with permission from Ref. [90]; 2022 Elsevier).
Figure 6. (a) Schematic illustration of zinc electrodeposition in 2 M ZnSO4 electrolyte (reprinted from [87]). (b,c) Schematic illustrations of Zn electrodeposition in different electrolytes: (b) 2 M ZnSO4 and (c) 2 M ZnSO4 with the presence of 20 mM Leu molecules (reprinted with permission from Ref. [88]; 2025 Royal Society of Chemistry). (d,e) Schematic illustration of Zn-plating-induced stress and its effect on Zn plating/stripping behavior on PL-Zn (d) and P-Zn (e) (reprinted with permission from Ref. [90]; 2022 Elsevier).
Metals 15 00784 g006

4. Conclusions and Outlook

AZIBs have emerged as a highly promising next-generation energy storage technology, demonstrating unique advantages for large-scale energy storage applications due to the zinc metal anode’s high theoretical capacity, low redox potential and abundant natural reserves. However, zinc anodes face two critical challenges in practical applications: (1) interfacial side reactions caused by hydrogen evolution and corrosion processes, which not only consume active materials but also alter the electrolyte environment; (2) dendrite growth resulting from uneven deposition, severely compromising cycling stability and safety performance. While conventional strategies like electrolyte modification and interface coating have achieved limited success, they often fail to address the intrinsic limitations of polycrystalline zinc materials at a fundamental level.
Grain boundary engineering has emerged as an innovative approach to overcome these challenges through atomic-scale microstructure regulation. This technique enables simultaneous achievement of multiple objectives—dendrite suppression, corrosion mitigation, and enhanced ion transport—by precisely controlling grain boundary density, chemical composition, and atomic arrangement. Specifically, high-density low-energy grain boundaries (e.g., coherent twin boundaries) can effectively homogenize Zn2+ flux distribution and reduce localized current density concentration. The preferred Zn(002) crystal orientation reduces surface energy by 40%, promoting horizontal layered zinc deposition. Introducing intermetallic compounds like TiZn16 at grain boundaries forms passivation layers that significantly decrease corrosion current density from 3.4 to 1.8 mA cm2. Meanwhile, defect channels along grain boundaries accelerate Zn2+ transport, reducing nucleation overpotential from 51 to 24 mV. Notably, grain boundary engineering exhibits synergistic effects when combined with conventional strategies like surface modification and alloying, enabling further performance enhancement. Current research demonstrates that grain-boundary-engineered zinc anodes can extend battery cycle life by 3–7 times while achieving Coulombic efficiency exceeding 99.9%.
Future research should focus on three key directions: (1) elucidating the dynamic correlation between grain boundary evolution and electrochemical behavior during cycling through in situ characterization and theoretical calculations; (2) developing cost-effective, scalable grain boundary engineering methods; (3) establishing quantitative structure–property relationships between grain boundary characteristics and electrochemical performance to guide material design. The development of grain boundary engineering not only provides a new pathway for practical AZIB applications but also offers valuable insights for optimizing lithium, sodium, and other metal anodes. With advancing research and maturing fabrication techniques, grain boundary engineering is poised to accelerate the commercialization of AZIBs in smart grids and distributed energy storage systems, contributing significantly to achieving carbon neutrality goals.

Author Contributions

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

Funding

This research was funded by the National Natural Science Foundation of China (No. U24A20195 and No. 52270177), Liaoning Province Science and Technology Plan Joint Program (Key Research and Development Program Project) (2023JH2/101800058), Postdoctoral Fund of Northeastern University (20240204) and the Fundamental Research Funds for the Central Universities (No. N2425035).

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 conflicts of interest.

References

  1. Zhang, H. Progress and perspectives of flow battery technologies. Curr. Opin. Electrochem. 2019, 18, 123–125. [Google Scholar] [CrossRef]
  2. Etacheri, V.; Marom, R.; Elazari, R.; Salitra, G.; Aurbach, D. Challenges in the development of advanced Li-ion batteries: A review. Energy Environ. Sci. 2011, 4, 3243–3262. [Google Scholar] [CrossRef]
  3. Wang, L.; Qiu, J.; Wang, X.; Chen, L.; Cao, G.; Wang, J.; Zhang, H.; He, X. Insights for understanding multiscale degradation of LiFePO4 cathodes. eScience 2022, 2, 125–137. [Google Scholar] [CrossRef]
  4. Lu, W.; Li, X.; Zhang, H. The next generation vanadium flow batteries with high power density—A perspective. Phys. Chem. Chem. Phys. 2017, 20, 23–35. [Google Scholar] [CrossRef]
  5. Al-Amin, M.; Islam, S.; Shibly, S.U.A.; Iffat, S. Comparative Review on the Aqueous Zinc-Ion Batteries (AZIBs) and Flexible Zinc-Ion Batteries (FZIBs). Nanomaterials 2022, 12, 3997. [Google Scholar] [CrossRef]
  6. Li, C.; Jin, S.; Archer, L.A.; Nazar, L.F. Toward practical aqueous zinc-ion batteries for electrochemical energy storage. Joule 2022, 6, 1733–1738. [Google Scholar] [CrossRef]
  7. Zhao, C.; Zhang, Y.; Gao, J.; Guo, Z.; Chen, A.; Liu, N.; Lu, X.; Zhang, X.; Zhang, N. Zincophilic Design and the Electrode/Electrolyte Interface for Aqueous Zinc-Ion Batteries: A Review. Batter. Supercaps 2023, 6, e202200478. [Google Scholar] [CrossRef]
  8. Zhou, T.; Huang, R.; Lu, Q.; Liu, P.; Hu, L.; Zhang, K.; Bai, P.; Xu, R.; Cao, X.; Sun, Z.; et al. Recent progress and perspectives on highly utilized Zn metal anode-towards marketable aqueous Zn-ion batteries. Energy Storage Mater. 2024, 72, 103689. [Google Scholar] [CrossRef]
  9. Al-Abbasi, M.; Zhao, Y.; He, H.; Liu, H.; Xia, H.; Zhu, T.; Wang, K.; Xu, Z.; Wang, H.; Zhang, W.; et al. Challenges and protective strategies on zinc anode toward practical aqueous zinc-ion batteries. Carbon. Neutralization 2024, 3, 108–141. [Google Scholar] [CrossRef]
  10. Xie, W.; Zhu, K.; Jiang, W.; Yang, H.; Ma, M.; Zhao, L.; Yang, W. Highly 002-Oriented Dendrite-Free Anode Achieved by Enhanced Interfacial Electrostatic Adsorption for Aqueous Zinc-Ion Batteries. ACS Nano 2024, 18, 21184–21197. [Google Scholar] [CrossRef]
  11. Dai, H.; Sun, T.; Zhou, J.; Wang, J.; Chen, Z.; Zhang, G.; Sun, S. Unraveling chemical origins of dendrite formation in zinc-ion batteries via in situ/operando X-ray spectroscopy and imaging. Nat. Commun. 2024, 15, 8577. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, J.; Mao, L.; Xia, Z.; Fan, M.; Deng, Y.; Chen, Z. Zincophilic Design for Highly Stable and Dendrite-Free Zinc Metal Anodes in Aqueous Zinc-Ion Batteries. Adv. Funct. Mater. 2025, 35, 2412547. [Google Scholar] [CrossRef]
  13. Zhang, X.; Hu, J.-P.; Fu, N.; Zhou, W.-B.; Liu, B.; Deng, Q.; Wu, X.-W. Comprehensive review on zinc-ion battery anode: Challenges and strategies. InfoMat 2022, 4, e12306. [Google Scholar] [CrossRef]
  14. Hu, Y.; Du, H.; Lu, J.; Li, Z.; Ma, Y.; Song, S. Achieving dendrite-free and long cyclic stability of zinc anode via CeO2 doped PEO interface modification. Fuel 2023, 354, 129417. [Google Scholar] [CrossRef]
  15. Li, F.; Zhang, H.; Liu, X.; Xu, B.; Wang, C. Dendrite-Free Zn Anode Modified by Organic Coating for Stable Aqueous Zinc Ion Batteries. Batteries 2024, 10, 420. [Google Scholar] [CrossRef]
  16. Zhang, Z.; Said, S.; Smith, K.; Zhang, Y.S.; He, G.; Jervis, R.; Shearing, P.R.; Miller, T.S.; Brett, D.J.L. Dendrite suppression by anode polishing in zinc-ion batteries. J. Mater. Chem. A 2021, 9, 15355–15362. [Google Scholar] [CrossRef]
  17. Gao, Q.; He, W.; Liu, C.; You, Y.; Zhang, P.; Liu, L.; Xu, G.; Gong, K.; Zhang, A.; Sun, Z. Dendrite-Free Zinc Anodes via a Three-Dimensional Ti2AlC Coating for High-Performance Zinc-Ion Batteries. ACS Appl. Energy Mater. 2025, 8, 1526–1534. [Google Scholar] [CrossRef]
  18. Das, A.; Balakrishnan, N.T.M.; Sreeram, P.; Fatima, M.J.J.; Joyner, J.D.; Thakur, V.K.; Pullanchiyodan, A.; Ahn, J.-H.; Raghavan, P. Prospects for magnesium ion batteries: A compreshensive materials review. Coord. Chem. Rev. 2024, 502, 215593. [Google Scholar] [CrossRef]
  19. Das, S.K.; Mahapatra, S.; Lahan, H. Aluminium-ion batteries: Developments and challenges. J. Mater. Chem. A 2017, 5, 6347–6367. [Google Scholar] [CrossRef]
  20. Wang, J.; Wang, J.-G.; Liu, H.; You, Z.; Wei, C.; Kang, F. Electrochemical activation of commercial MnO microsized particles for high-performance aqueous zinc-ion batteries. J. Power Sources 2019, 438, 226951. [Google Scholar] [CrossRef]
  21. Zhang, C.; Holoubek, J.; Wu, X.; Daniyar, A.; Zhu, L.; Chen, C.; Leonard, D.P.; Rodríguez-Pérez, I.A.; Jiang, J.-X.; Fang, C.; et al. A ZnCl2 water-in-salt electrolyte for a reversible Zn metal anode. Chem. Commun. 2018, 54, 14097–14099. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, F.-H.; Liu, H.-B. Research Progress of Zinc Anode Materials for Aqueous Zinc Ion Recharge Battery. Chin. J. Inorg. Chem. 2019, 35, 1999–2012. [Google Scholar] [CrossRef]
  23. Chen, L.-N.; Yan, M.-Y.; Mei, Z.-W.; Mai, L.-Q. Research Progress and Prospect of Aqueous Zinc Ion Battery. J. Inorg. Mater. 2017, 32, 225–234. [Google Scholar] [CrossRef]
  24. Wang, J.; Liu, Z.; Wang, H.; Cui, F.; Zhu, G. Integrated pyrazine-based porous aromatic frameworks/carbon nanotube composite as cathode materials for aqueous zinc ion batteries. Chem. Eng. J. 2022, 450, 138051. [Google Scholar] [CrossRef]
  25. Zhang, Q.; Luan, J.; Tang, Y.; Ji, X.; Wang, H. Interfacial Design of Dendrite-Free Zinc Anodes for Aqueous Zinc-Ion Batteries. Angew. Chem. Int. Ed. 2020, 59, 13180–13191. [Google Scholar] [CrossRef]
  26. Yufit, V.; Tariq, F.; Eastwood, D.S.; Biton, M.; Wu, B.; Lee, P.D.; Brandon, N.P. Operando Visualization and Multi-scale Tomography Studies of Dendrite Formation and Dissolution in Zinc Batteries. Joule 2019, 3, 485–502. [Google Scholar] [CrossRef]
  27. Lu, H.; Zhang, X.; Luo, M.; Cao, K.; Lu, Y.; Xu, B.B.; Pan, H.; Tao, K.; Jiang, Y. Amino Acid-Induced Interface Charge Engineering Enables Highly Reversible Zn Anode. Adv. Funct. Mater. 2021, 31, 2103514. [Google Scholar] [CrossRef]
  28. Luo, C.; Yuan, H. Measurement and modeling of deformation-induced martensitic transformation in a metastable austenitic stainless steel under cyclic loadings. Acta Mater. 2022, 238, 118202. [Google Scholar] [CrossRef]
  29. Feng, Z.; Feng, Y.; Fan, F.; Deng, D.; Dong, H.; Liu, S.; Kang, L.; Jun, S.C.; Wang, L.; Zhu, J.; et al. Functionalization design of zinc anode for advanced aqueous zinc-ion batteries. SusMat 2024, 4, e184. [Google Scholar] [CrossRef]
  30. Sajjadnejad, M.; Ghorbani, M.; Afshar, A. Microstructure-corrosion resistance relationship of direct and pulse current electrodeposited Zn-TiO2 nanocomposite coatings. Ceram. Int. 2015, 41, 217–224. [Google Scholar] [CrossRef]
  31. Jia, X.; Liu, C.; Neale, Z.G.; Yang, J.; Cao, G. Active Materials for Aqueous Zinc Ion Batteries: Synthesis, Crystal Structure, Morphology, and Electrochemistry. Chem. Rev. 2020, 120, 7795–7866. [Google Scholar] [CrossRef] [PubMed]
  32. Jeong, D.Y.; Chang, W.J.; Jang, S.; Kim, M.; Kim, Y.; Kim, B.; Park, W.I. Controlling dendrite growth and side reactions in anode-free Zn-ion aqueous batteries with PMMA:Zn coated electrode. J. Energy Storage 2024, 76, 109791. [Google Scholar] [CrossRef]
  33. Cao, Z.; Zhuang, P.; Zhang, X.; Ye, M.; Shen, J.; Ajayan, P.M. Strategies for Dendrite-Free Anode in Aqueous Rechargeable Zinc Ion Batteries. Adv. Energy Mater. 2020, 10, 2001599. [Google Scholar] [CrossRef]
  34. Shi, P.; Fu, Z.-H.; Zhou, M.-Y.; Chen, X.; Yao, N.; Hou, L.-P.; Zhao, C.-Z.; Li, B.-Q.; Huang, J.-Q.; Zhang, X.-Q.; et al. Inhibiting intercrystalline reactions of anode with electrolytes for long-cycling lithium batteries. Sci. Adv. 2022, 8, eabq3445. [Google Scholar] [CrossRef] [PubMed]
  35. Yan, Q.; Ko, S.-T.; Dawson, A.; Agyeman-Budu, D.; Whang, G.; Zhao, Y.; Qin, M.; Dunn, B.S.; Weker, J.N.; Tolbert, S.H.; et al. Thermodynamics-driven interfacial engineering of alloy-type anode materials. Cell Rep. Phys. Sci. 2022, 3, 100694. [Google Scholar] [CrossRef]
  36. Yule, L.C.; Shkirskiy, V.; Aarons, J.; West, G.; Bentley, C.L.; Shollock, B.A.; Unwin, P.R. Nanoscale Active Sites for the Hydrogen Evolution Reaction on Low Carbon Steel. J. Phys. Chem. C 2019, 123, 24146–24155. [Google Scholar] [CrossRef]
  37. Huang, S.; Wu, W.; Su, Y.; Qiao, L.; Yan, Y. Insight into the corrosion behaviour and degradation mechanism of pure zinc in simulated body fluid. Corros. Sci. 2021, 178, 109071. [Google Scholar] [CrossRef]
  38. Cai, Z.; Ou, Y.; Zhang, B.; Wang, J.; Fu, L.; Wan, M.; Li, G.; Wang, W.; Wang, L.; Jiang, J.; et al. A Replacement Reaction Enabled Interdigitated Metal/Solid Electrolyte Architecture for Battery Cycling at 20 mA cm−2 and 20 mAh cm−2. J. Am. Chem. Soc. 2021, 143, 3143–3152. [Google Scholar] [CrossRef]
  39. Kobayashi, S.; Kobayashi, R.; Watanabe, T. Control of grain boundary connectivity based on fractal analysis for improvement of intergranular corrosion resistance in SUS316L austenitic stainless steel. Acta Mater. 2016, 102, 397–405. [Google Scholar] [CrossRef]
  40. Shi, P.; Hu, R.; Zhang, T.; Yuan, L.; Li, J. Grain boundary character distribution and its effect on corrosion of Ni-23Cr-16Mo superalloy. Mater. Sci. Technol. 2017, 33, 84–91. [Google Scholar] [CrossRef]
  41. Liu, T.; Xia, S.; Bai, Q.; Zhou, B.; Zhang, L.; Lu, Y.; Shoji, T. Three-dimensional study of grain boundary engineering effects on intergranular stress corrosion cracking of 316 stainless steel in high temperature water. J. Nucl. Mater. 2018, 498, 290–299. [Google Scholar] [CrossRef]
  42. Stratulat, A.; Duff, J.A.; Marrow, T.J. Grain boundary structure and intergranular stress corrosion crack initiation in high temperature water of a thermally sensitised austenitic stainless steel, observed in situ. Corros. Sci. 2014, 85, 428–435. [Google Scholar] [CrossRef]
  43. Lehockey, E.M.; Palumbo, G. On the creep behaviour of grain boundary engineered nickel. Mater. Sci. Eng.-Struct. Mater. Prop. Microstruct. Process. 1997, 237, 168–172. [Google Scholar] [CrossRef]
  44. Xu, M.; Chen, J.; Lu, H.; Xu, J.; Yu, C.; Wei, X. Effects of residual stress and grain boundary character on creep cracking in 2.25Cr-1.6W steel. Mater. Sci. Eng.-Struct. Mater. Prop. Microstruct. Process. 2016, 659, 188–197. [Google Scholar] [CrossRef]
  45. Watanabe, T. A new era of grain boundary design and control for high temperature materials. In Creep and Fracture of Engineering Materials and Structures; Key Engineering Materials; Sakuma, T., Yagi, K., Eds.; Transtec Publications Ltd.: Zurich, Switzerland, 2000; pp. 237–244. [Google Scholar]
  46. Watanabe, T. Grain boundary engineering: Historical perspective and future prospects. J. Mater. Sci. 2011, 46, 4095–4115. [Google Scholar] [CrossRef]
  47. Lejček, P.; Paidar, V. Challenges of interfacial classification for grain boundary engineering. Mater. Sci. Technol. 2005, 21, 393–398. [Google Scholar] [CrossRef]
  48. Shimada, M.; Kokawa, H.; Wang, Z.J.; Sato, Y.S.; Karibe, I. Optimization of grain boundary character distribution for intergranular corrosion resistant 304 stainless steel by twin-induced grain boundary engineering. Acta Mater. 2002, 50, 2331–2341. [Google Scholar] [CrossRef]
  49. Singh, R.; Chowdhury, S.G.; Kumar, B.R.; Das, S.K.; De, P.K.; Chattoraj, I. The importance of grain size relative to grain boundary character on the sensitization of metastable austenitic stainless steel. Scr. Mater. 2007, 57, 185–188. [Google Scholar] [CrossRef]
  50. Kurban, M.; Erb, U.; Aust, K.T. A grain boundary characterization study of boron segregation and carbide precipitation in alloy 304 austenitic stainless steel. Scr. Mater. 2006, 54, 1053–1058. [Google Scholar] [CrossRef]
  51. Schwartz, A.J.; King, W.E.; Kumar, M. Influence of processing method on the network of grain boundaries. Scr. Mater. 2006, 54, 963–968. [Google Scholar] [CrossRef]
  52. Randle, V.; Coleman, M. A study of low-strain and medium-strain grain boundary engineering. Acta Mater. 2009, 57, 3410–3421. [Google Scholar] [CrossRef]
  53. Zhang, X.; Zhao, X.; Li, D.; Guo, S. Relationships Between the Σ9 and Σ27 Boundaries and the Connectivity of Random Boundary in Hastelloy C-276 Alloy. Rare Met. Mater. Eng. 2018, 47, 2318–2321. [Google Scholar]
  54. Huo, R.; Du, Z.; Cheng, J.; Sun, B.; Gong, T.; Du, X. The Influence of Deformation and Heat Treatment on the Grain Boundary Character Distribution in CoCrFeNi High-Entropy Alloy. J. Mater. Eng. Perform. 2025, 1–12. [Google Scholar] [CrossRef]
  55. Li, Z.; Zhang, L.; Sun, N.; Sun, Y.; Shan, A. Effects of prior deformation and annealing process on microstructure and annealing twin density in a nickel based alloy. Mater. Charact. 2014, 95, 299–306. [Google Scholar] [CrossRef]
  56. Huang, W.; Chai, L.; Li, Z.; Yang, X.; Guo, N.; Song, B. Evolution of microstructure and grain boundary character distribution of a tin bronze annealed at different temperatures. Mater. Charact. 2016, 114, 204–210. [Google Scholar] [CrossRef]
  57. Tokita, S.; Kokawa, H.; Sato, Y.S.; Fujii, H.T. In situ EBSD observation of grain boundary character distribution evolution during thermomechanical process used for grain boundary engineering of 304 austenitic stainless steel. Mater. Charact. 2017, 131, 31–38. [Google Scholar] [CrossRef]
  58. Zhang, J.X.; Zhang, J.Y.; Xiao, W.C.; Hou, J.X.; Li, Q.; Xiao, B.; Yang, S.Y.; Wu, J.L.; Zhang, T.L.; Yang, T. Grain refinement induced unusually large shape memory effect in lightweight titanium alloy. Acta Mater. 2024, 272, 119936. [Google Scholar] [CrossRef]
  59. Shi, F.; Tian, P.C.; Jia, N.; Ye, Z.H.; Qi, Y.; Liu, C.M.; Li, X.W. Improving intergranular corrosion resistance in a nickel-free and manganese-bearing high-nitrogen austenitic stainless steel through grain boundary character distribution optimization. Corros. Sci. 2016, 107, 49–59. [Google Scholar] [CrossRef]
  60. Zhang, W.; Liu, F.; Liu, L.; Li, Q.; Liu, L.; Liu, F.; Huang, C. Effect of grain size and distribution on the corrosion behavior of Y2O3 dispersion-strengthened 304 stainless steel. Mater. Today Commun. 2022, 31, 103723. [Google Scholar] [CrossRef]
  61. Randle, V.; Jones, R. Grain boundary plane distributions and single-step versus multiple-step grain boundary engineering. Mater. Sci. Eng.-Struct. Mater. Prop. Microstruct. Process. 2009, 524, 134–142. [Google Scholar] [CrossRef]
  62. Seita, M.; Gao, S. Broadening the design space of engineering materials through “additive grain boundary engineering”. J. Mater. Sci. 2022, 57, 9530–9540. [Google Scholar] [CrossRef]
  63. Zheng, J.; Zhao, Q.; Tang, T.; Yin, J.; Quilty, C.D.; Renderos, G.D.; Liu, X.; Deng, Y.; Wang, L.; Bock, D.C.; et al. Reversible epitaxial electrodeposition of metals in battery anodes. Science 2019, 366, 645–648. [Google Scholar] [CrossRef]
  64. Zhou, M.; Guo, S.; Li, J.; Luo, X.; Liu, Z.; Zhang, T.; Cao, X.; Long, M.; Lu, B.; Pan, A.; et al. Surface-Preferred Crystal Plane for a Stable and Reversible Zinc Anode. Adv. Mater. 2021, 33, 2100187. [Google Scholar] [CrossRef]
  65. Wen, Q.; Fu, H.; Cui, R.; Chen, H.-Z.; Ji, R.-H.; Tang, L.-B.; Yan, C.; Mao, J.; Dai, K.-H.; Zhang, X.-H.; et al. Recent advances in interfacial modification of zinc anode for aqueous rechargeable zinc ion batteries. J. Energy Chem. 2023, 83, 287–303. [Google Scholar] [CrossRef]
  66. Zhang, Z.; Xia, S.; Dong, A.; Li, X.; Wang, F.; Yang, J.; Ruan, J.; Li, Q.; Sun, D.; Fang, F.; et al. Mechanical Grinding Formation of Highly Reversible (002)-Textured Zinc Metal Anodes. Adv. Energy Mater. 2025, 15, 2403598. [Google Scholar] [CrossRef]
  67. Zheng, Z.; Zhong, X.; Zhang, Q.; Zhang, M.; Dai, L.; Xiao, X.; Xu, J.; Jiao, M.; Wang, B.; Li, H.; et al. An extended substrate screening strategy enabling a low lattice mismatch for highly reversible zinc anodes. Nat. Commun. 2024, 15, 753. [Google Scholar] [CrossRef]
  68. Wang, J.; Zhang, B.; Cai, Z.; Zhan, R.; Wang, W.; Fu, L.; Wan, M.; Xiao, R.; Ou, Y.; Wang, L.; et al. Stable interphase chemistry of textured Zn anode for rechargeable aqueous batteries. Sci. Bull. 2022, 67, 716–724. [Google Scholar] [CrossRef]
  69. Yi, Z.; Chen, G.; Hou, F.; Wang, L.; Liang, J. Strategies for the Stabilization of Zn Metal Anodes for Zn-Ion Batteries. Adv. Energy Mater. 2021, 11, 2003065. [Google Scholar] [CrossRef]
  70. Chen, Z.; Wu, Q.; Han, X.; Wang, C.; Chen, J.; Hu, T.; He, Q.; Zhu, X.; Yuan, D.; Chen, J.; et al. Converting Commercial Zn Foils into Single (002)-Textured Zn with Millimeter-Sized Grains for Highly Reversible Aqueous Zinc Batteries. Angew. Chem.-Int. Ed. 2024, 63, e202401507. [Google Scholar] [CrossRef]
  71. Yu, X.; Li, Z.; Wu, X.; Zhang, H.; Zhao, Q.; Liang, H.; Wang, H.; Chao, D.; Wang, F.; Qiao, Y.; et al. Ten concerns of Zn metal anode for rechargeable aqueous zinc batteries. Joule 2023, 7, 1145–1175. [Google Scholar] [CrossRef]
  72. Chen, Z.; Zhao, J.; He, Q.; Li, M.; Feng, S.; Wang, Y.; Yuan, D.; Chen, J.; Alshareef, H.N.; Ma, Y. Texture Control of Commercial Zn Foils Prolongs Their Reversibility as Aqueous Battery Anodes. ACS Energy Lett. 2022, 7, 3564–3571. [Google Scholar] [CrossRef]
  73. Zhang, X.; Li, J.; Liu, Y.; Lu, B.; Liang, S.; Zhou, J. Single [0001]-oriented zinc metal anode enables sustainable zinc batteries. Nat. Commun. 2024, 15, 2735. [Google Scholar] [CrossRef] [PubMed]
  74. Shen, T.; Wei, T.; Zhang, S.; Liu, H.; Li, C.; Li, Z.; Yang, M.; Liu, C.; Pei, Y. Single-Crystalline Zn(002) Facet Enables Ultrastable Anode–Electrolyte Interface. Small Struct. 2025, 6, 2400325. [Google Scholar] [CrossRef]
  75. Lian, S.; Cai, Z.; Yan, M.; Sun, C.; Chai, N.; Zhang, B.; Yu, K.; Xu, M.; Zhu, J.; Pan, X.; et al. Ultra-High Proportion of Grain Boundaries in Zinc Metal Anode Spontaneously Inhibiting Dendrites Growth. Angew. Chem.-Int. Ed. 2024, 63, e202406292. [Google Scholar] [CrossRef] [PubMed]
  76. Zou, Q.; Liang, Z.; Wang, W.; Dong, D.; Lu, Y.-C. A nuclei-rich strategy for highly reversible dendrite-free zinc metal anodes. Energy Environ. Sci. 2023, 16, 6026–6034. [Google Scholar] [CrossRef]
  77. Li, Y.; Wu, P.; Zhong, W.; Xie, C.; Xie, Y.; Zhang, Q.; Sun, D.; Tang, Y.; Wang, H. A progressive nucleation mechanism enables stable zinc stripping–plating behavior. Energy Environ. Sci. 2021, 14, 5563–5571. [Google Scholar] [CrossRef]
  78. Zhou, L.-F.; Du, T.; Li, J.-Y.; Wang, Y.-S.; Gong, H.; Yang, Q.-R.; Chen, H.; Luo, W.-B.; Wang, J.-Z. A strategy for anode modification for future zinc-based battery application. Mater. Horiz. 2022, 9, 2722–2751. [Google Scholar] [CrossRef]
  79. Huang, W.; Huang, Y.; Huang, X.; Shao, F.; Liu, W.; Kang, F. 3D Leaf-Like Copper-Zinc Alloy Enables Dendrite-Free Zinc Anode for Ultra-Long Life Aqueous Zinc Batteries. Small 2024, 20, 2404294. [Google Scholar] [CrossRef]
  80. Li, J.; Liang, H.; Long, Y.; Yu, X.; Li, J.; Li, N.; Han, J.; Wang, J.; Yang, Z. Dual-function protective layer for highly reversible Zn anode. J. Energy Chem. 2024, 98, 12–23. [Google Scholar] [CrossRef]
  81. Yuan, D.; Zhao, J.; Ren, H.; Chen, Y.; Chua, R.; Jie, E.T.J.; Cai, Y.; Edison, E.; Manalastas, W.; Wong, M.W.; et al. Anion Texturing Towards Dendrite-Free Zn Anode for Aqueous Rechargeable Batteries. Angew. Chem. 2021, 60, 7213–7219. [Google Scholar] [CrossRef]
  82. Zhao, Y.; Guo, S.; Chen, M.; Lu, B.; Zhang, X.; Liang, S.; Zhou, J. Tailoring grain boundary stability of zinc-titanium alloy for long-lasting aqueous zinc batteries. Nat. Commun. 2023, 14, 7080. [Google Scholar] [CrossRef]
  83. Li, J.; Liu, Z.; Han, S.; Zhou, P.; Lu, B.; Zhou, J.; Zeng, Z.; Chen, Z.; Zhou, J. Hetero Nucleus Growth Stabilizing Zinc Anode for High-Biosecurity Zinc-Ion Batteries. Nano-Micro Lett. 2023, 15, 237. [Google Scholar] [CrossRef]
  84. Chen, M.; Gong, Y.; Zhao, Y.; Song, Y.; Tang, Y.; Zeng, Z.; Liang, S.; Zhou, P.; Lu, B.; Zhang, X.; et al. Spontaneous grain refinement effect of rare earth zinc alloy anodes enables stable zinc batteries. Natl. Sci. Rev. 2024, 11, nwae205. [Google Scholar] [CrossRef]
  85. Chen, Z.; Wang, Y.; Wu, Q.; Wang, C.; He, Q.; Hu, T.; Han, X.; Chen, J.; Zhang, Y.; Chen, J.; et al. Grain Boundary Filling Empowers (002)-Textured Zn Metal Anodes with Superior Stability. Adv. Mater. 2024, 36, 2411004. [Google Scholar] [CrossRef] [PubMed]
  86. Liu, C.; Luo, Z.; Deng, W.; Wei, W.; Chen, L.; Pan, A.; Ma, J.; Wang, C.; Zhu, L.; Xie, L.; et al. Liquid Alloy Interlayer for Aqueous Zinc-Ion Battery. ACS Energy Lett. 2021, 6, 675–683. [Google Scholar] [CrossRef]
  87. Li, X.; Chen, Z.; Ruan, P.; Hu, X.; Lu, B.; Yuan, X.; Tian, S.; Zhou, J. Inducing preferential growth of the Zn. (002) plane by using a multifunctional chelator for achieving highly reversible Zn anodes. Nanoscale 2024, 16, 2923–2930. [Google Scholar] [CrossRef] [PubMed]
  88. Li, X.; Chen, Z.; Ruan, P.; Hu, X.; Yuan, X.; Lu, B.; Qin, L.; Zhou, J. Guiding uniform Zn deposition with a multifunctional additive for highly utilized Zn anodes. Nanoscale 2024, 16, 18835–18842. [Google Scholar] [CrossRef]
  89. Chen, X.; Li, W.; Hu, S.; Akhmedov, N.G.; Reed, D.; Li, X.; Liu, X. Polyvinyl alcohol coating induced preferred crystallographic orientation in aqueous zinc battery anodes. Nano Energy 2022, 98, 107269. [Google Scholar] [CrossRef]
  90. Chen, J.; Qiao, X.; Han, X.; Zhang, J.; Wu, H.; He, Q.; Chen, Z.; Shi, L.; Wang, Y.; Xie, Y.; et al. Releasing plating-induced stress for highly reversible aqueous Zn metal anodes. Nano Energy 2022, 103, 107814. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of zinc anode dendrites and their associated side reaction mechanisms (adapted from [29]).
Figure 1. Schematic diagram of zinc anode dendrites and their associated side reaction mechanisms (adapted from [29]).
Metals 15 00784 g001
Figure 2. Schematic illustrations of formation processes for different low-angle grain boundaries. (a,b) Evolution of symmetric tilt boundaries showing (a) pre-tilt and (b) post-tilt configurations; (c) formation mechanism of twist boundaries through rotational misorientation.
Figure 2. Schematic illustrations of formation processes for different low-angle grain boundaries. (a,b) Evolution of symmetric tilt boundaries showing (a) pre-tilt and (b) post-tilt configurations; (c) formation mechanism of twist boundaries through rotational misorientation.
Metals 15 00784 g002
Figure 3. Schematic illustration of grain boundary engineering mechanisms in zinc anodes.
Figure 3. Schematic illustration of grain boundary engineering mechanisms in zinc anodes.
Metals 15 00784 g003
Figure 5. (a) Schematics of introducing the Sn hetero nucleus and its effects on mitigating dendrite growth and hydrogen evolution reaction (reprinted from [83]). (b) Schematic illustration of the preparation mechanism for URAL (reprinted from [84]). (c,d) Schematic illustration showing the design of (c) common (002)-textured Zn anodes and (d) the IM(002) Zn anodes (reprinted with permission from Ref. [85]; 2024 Wiley).
Figure 5. (a) Schematics of introducing the Sn hetero nucleus and its effects on mitigating dendrite growth and hydrogen evolution reaction (reprinted from [83]). (b) Schematic illustration of the preparation mechanism for URAL (reprinted from [84]). (c,d) Schematic illustration showing the design of (c) common (002)-textured Zn anodes and (d) the IM(002) Zn anodes (reprinted with permission from Ref. [85]; 2024 Wiley).
Metals 15 00784 g005
Table 1. Comparative analysis of LIBs and AZIBs properties (adapted from [8,9,10,11,12]).
Table 1. Comparative analysis of LIBs and AZIBs properties (adapted from [8,9,10,11,12]).
PropertiesAZIBsLIBs
Energy Density50–135 Wh/kg160 Wh/kg
SafetyExcellent: Non-flammable aqueous electrolyte; no dendritic explosion risksPoor: Flammable organic electrolyte; lithium dendrites may trigger thermal runaway
CostLow: Abundant zinc resourcesHigh: Dependent on scarce resources (e.g., lithium, cobalt)
Cycle LifeModerate: 100–1500 cyclesLong: 3000 cycles
Environmental FriendlinessHigh: Non-toxic componentsLow: Pollution risks from organic solvents
Rate CapabilitySuperior: Fast Zn2+ diffusion enables rapid chargingModerate: Li+ mobility limited by organic electrolytes
Key ChallengesZinc dendrites, hydrogen evolution reaction (HER), side reactionsLithium dendrites, thermal runaway, electrolyte decomposition
Table 2. Systematic comparison of fundamental properties of common metallic anodes (adapted from [2,13,14,15,16,17,18,19]).
Table 2. Systematic comparison of fundamental properties of common metallic anodes (adapted from [2,13,14,15,16,17,18,19]).
PropertiesZnLiNaMgAl
Ionic radius (Å)0.74 (Zn2+)0.76 (Li+)1.02 (Na+)0.72 (Mg2+)0.54 (Al3+)
Volumetric capacity (mAh/cm3)58552046112938338046
Electrode potential (V vs. SHE)−0.76−3.04−2.71−2.37−1.66
Battery systemAqueous zinc-ion batteriesOrganic lithium-ion batteriesSodium-ion batteriesMagnesium-ion batteriesAluminum-ion batteries
Key advantagesSafe with low cost and excellent rate capabilityEnergy-dense with low self-discharge and mature technologyResource-abundant with cost-effectiveness and good low-temperature performanceDivalent-enabled with dendrite-free nature and high volumetric energy densityTrivalent-advantaged with corrosion-resistant and lightweight properties
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

Liu, Y.-X.; Wang, J.-Z.; Cao, L.; Wang, H.; Cheng, Z.-Y.; Zhou, L.-F.; Du, T. Grain Boundary Engineering for Reversible Zn Anodes in Rechargeable Aqueous Zn-Ion Batteries. Metals 2025, 15, 784. https://doi.org/10.3390/met15070784

AMA Style

Liu Y-X, Wang J-Z, Cao L, Wang H, Cheng Z-Y, Zhou L-F, Du T. Grain Boundary Engineering for Reversible Zn Anodes in Rechargeable Aqueous Zn-Ion Batteries. Metals. 2025; 15(7):784. https://doi.org/10.3390/met15070784

Chicago/Turabian Style

Liu, Yu-Xuan, Jun-Zhe Wang, Lei Cao, Hao Wang, Zhen-Yu Cheng, Li-Feng Zhou, and Tao Du. 2025. "Grain Boundary Engineering for Reversible Zn Anodes in Rechargeable Aqueous Zn-Ion Batteries" Metals 15, no. 7: 784. https://doi.org/10.3390/met15070784

APA Style

Liu, Y.-X., Wang, J.-Z., Cao, L., Wang, H., Cheng, Z.-Y., Zhou, L.-F., & Du, T. (2025). Grain Boundary Engineering for Reversible Zn Anodes in Rechargeable Aqueous Zn-Ion Batteries. Metals, 15(7), 784. https://doi.org/10.3390/met15070784

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