Next Article in Journal
The Strategies for Increasing Grid-Integrated Share of Renewable Energy with Energy Storage and Existing Coal Fired Power Generation in China
Previous Article in Journal
Large-Power Transformers: Time Now for Addressing Their Monitoring and Failure Investigation Techniques
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Challenges and Perspectives for Doping Strategy for Manganese-Based Zinc-ion Battery Cathode

1
State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China
2
Department of Physical Science and Technology, School of Science, Wuhan University of Technology, Wuhan 430070, China
3
Hainan Institute, Wuhan University of Technology, Wuhan 430070, China
4
Laboratory of Advanced Separations, Ecole Polytechnique Federale de Lausanne, 1951 Sion, Switzerland
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(13), 4698; https://doi.org/10.3390/en15134698
Submission received: 31 May 2022 / Revised: 20 June 2022 / Accepted: 22 June 2022 / Published: 27 June 2022
(This article belongs to the Section D: Energy Storage and Application)

Abstract

:
As one of the most appealing options for large-scale energy storage systems, the commercialization of aqueous zinc-ion batteries (AZIBs) has received considerable attention due to their cost effectiveness and inherent safety. A potential cathode material for the commercialization of AZIBs is the manganese-based cathode, but it suffers from poor cycle stability, owing to the Jahn–Teller effect, which leads to the dissolution of Mn in the electrolyte, as well as low electron/ion conductivity. In order to solve these problems, various strategies have been adopted to improve the stability of manganese-based cathode materials. Among those, the doping strategy has become popular, where the dopant is inserted into the intrinsic crystal structures of electrode materials, which would stabilize them and tune the electronic state of the redox center to realize high ion/electron transport. Herein, we summarize the ion doping strategy from the following aspects: (1) synthesis strategy of doped manganese-based oxides; (2) valence-dependent dopant ions in manganese-based oxides; (3) optimization mechanism of ion doping in zinc-manganese battery. Lastly, an in-depth understanding and future perspectives of ion doping strategy in electrode materials are provided for the commercialization of manganese-based zinc-ion batteries.

1. Introduction

Due to the excessive consumption of fossil fuels and the growing problem of climate change caused by environmental pollution, renewable energy research and development, such as solar, wind, and tidal energy, has attracted worldwide attention [1,2,3,4]. However, renewable energy power generation is often intermittent and unpredictable, which requires large-scale energy storage systems to effectively buffer such fluctuations to achieve stable energy output for smart grid [5,6,7]. As an efficient and flexible energy storage device, lithium-ion batteries (LIBs) have not only been successfully applied in electronic consumer products such as cellphones and laptops, but are also expanding their range to electric vehicles and other fields [8,9,10,11]. However, the high production cost, limited lithium resource reserves, and the use of toxic and flammable organic electrolytes make lithium-ion batteries expensive, hazardous, and environmentally polluting, strongly impeding their further development and application in grid-scale energy storage [12,13,14]. Therefore, researchers are seeking new energy storage battery systems to replace LIBs in terms of cost, safety, and sustainability.
Among those energy storage devices, aqueous electrolytes have been widely employed in secondary battery systems for Na+, K+, Mg2+, Al3+, Ca2+, Zn2+, etc. in recent years because of their safety, high ionic conductivity, and ease of operation [15,16]. Among aqueous multivalent ion batteries, ZIBs stand out due to the following characteristics: (1) Zn metal reserves are abundant and the manufacture process of ZIBs occurs in an air environment, making it cost-effective; (2) Zn metal anode has a low redox potential of −0.76 V with respect to a standard hydrogen electrode and high theoretical gravimetric/volumetric capacity (820 mAh·g−1/5855 mAh·cm−3); (3) Zn metal can be directly applied as an anode due to its excellent electrochemical stability and reversibility in water; (4) ZIBs are highly safe because of the application of nontoxic aqueous electrolyte [17,18]. However, considering the large ionic radius of hydrated zinc (5.5 Å vs. 0.74 Å for Zn ion), the intercalation of hydrated zinc ions would either require large spacing to accommodate the large ions or withstand a large desolvation penalty for smaller dehydrated ions to intercalate, imposing a great challenge in the development of suitable cathode materials [19]. At present, manganese-based oxides, vanadium-based oxides, and Prussian blue analogs are mainly developed as cathode materials for ZIBs [20]. Among them, manganese-based oxides are widely recognized as candidates for the commercialization of ZIBs because of their mature synthesis process, abundant resources, lack of pollution, high specific capacity, and high operating voltage [21]. However, the further application of Mn-based cathodes is hindered by two major issues. The redox reaction involving Mn4+ is usually accompanied by the Jahn–Teller effect and leads to the formation of Mn2+, which tends to dissolve into the electrolyte and lead to irreversible capacity loss. On the other hand, the poor ion/electric conductivity of the transition metal oxide would sacrifice the rate capability of the zinc battery [9,22,23]. At present, various strategies such as nanostructure engineering [24,25,26], conductive agent coating [27], and ion doping are widely adopted to tackle the above problems [21,28,29]. Among the various strategies, ion doping involves a small number of guest ions being preinserted into the manganese-based oxide framework and interacting with the host atoms to achieve an inherent structure optimization, which significantly enhances the electrochemical performance from a fundamental thermodynamics and dynamics aspect. This approach is recognized as an efficient and straightforward optimization strategy, breaking through the limitations of the inherent crystallographic structure [30].
In this review, considering the current research progress of ion doping optimization strategies in manganese-based oxides, we first briefly summarize the synthesis strategy, discuss the dependence of the valence states of doping ions on the manganese-based oxide structure, and highlight the optimization mechanism of ion doping in terms of the electrochemical performance. Lastly, we provide unique insights and prospects for ion doping strategies in manganese-based oxides for the application of zinc batteries for large grid-scale energy storage devices.

2. Synthesis Strategy for Ion-Doped Manganese-Based Oxides

Manganese oxides are highly dependent on dopant ions due to their effect on the crystalline phase, crystal structure, and average valence of the manganese oxides [31,32,33]. Daniil et al. built an ab initio model using the SCAN function to reveal the effect of doping different guest ions on the formation of manganese dioxide with different phases, as depicted in Figure 1a [31]. As shown in Figure 1b, the doping of Na+, K+, and Ca2+ was more likely to form α-MnO2, whereas Li+ and Mg2+ favored the formation of γ-MnO2, and δ-MnO2 was easily stabilized by Na+. The mechanistic basis was that distinct metal ion doping resulted in varied formation free energies required for the different phases of manganese dioxide. Hu et al. demonstrated that partial Mg2+ intercalation resulted in tunnels of various sizes, such as 3 × 3, 4 × 3, and 5 × 3 tunnels in T-MnO2, not just 3 × 3 tunnels [32]. Furthermore, the synthesis methods and synthesis circumstances also have a significant impact on the electrochemical performance of manganese oxides. Up to now, various synthesis methods have been widely applied in ion doping, including the hydrothermal method, ion penetration/exchange method, electrodeposition method, and calcination treatment.

2.1. Hydrothermal Method

Because of its simple controllability over the diverse crystal phases of manganese-based oxides, hydrothermal synthesis is the most frequently used approach for ion doping. Ion doping can be controlled by the addition of the different ions in the raw solutions before hydrothermal treatment. Zhang et al. prepared α-K0.19MnO2 nanotubes via decomposition of KMnO4 combined with carbon nanofibers as templates in 2019 [34]. Under the hydrothermal conditions of 140 °C for 10 h, a typical K+-doped tunnel structure was achieved. The reaction mechanism was as follows: K+ + MnO4 + C + H2O → δ-K0.19MnO2·nH2O + CO32− + HCO3. Through a similar route, MnO2H0.16(H2O)0.27 nanolayers were synthesized via the reaction of KMnO4 with acetylene black at 120 °C for 24 h by Pan et al. [35]. Interestingly, this new phase exhibited excellent rate capability (115.1 mAh·g−1 at 10 °C) with robust structural stability even with an interlayer spacing of less than 0.3 nm. Moreover, they developed a layered K0.36H0.26MnO2·0.28H2O via the neutralization reaction of KMnO4 and MnSO4 with K2SO4 additive as an excellent cathode in ZIBs [36]. The layer-type structure of monoclinic birnessite phase was obtained following a hydrothermal reaction of 120 °C for 12 h with a large interlayer spacing (7.12 Å). Shi et al. synthesized a cathode (K0.29MnO2·0.67H2O) with a larger interplanar spacing (7.4 Å) through a hydrothermal potassium insertion strategy [37]. Peng et al. developed a Na+-incorporated layered δ-MnO2 by subjecting K+-containing δ-MnO2 to hydrothermal treatment using a 0.5 mol·L−1 Na2SO4 solution at 180 °C for 3 h [38]. Layered Ca0.28MnO2·0.5H2O was synthesized by Tao et al. through a hydrothermal method at 160 °C for 12 h using CaCl2, KMnO4, and MnSO4 as reactants with a molar ratio of 1:6:1 [39]. This work demonstrated that divalent alkaline earth metal ions could also support layered manganese oxides, resulting in excellent electrochemical performance. Li et al. successfully incorporated Ni2+ into α-MnO2 for boosting the diffusion kinetics of protons in the tunnels, which proved the substitution of divalent metal ions for Mn sites in tunnel-type manganese based oxides [40]. Du et al. found that the addition of Ce3+ ions during hydrothermal synthesis could induce a phase transition of MnO2 from β to α, which resulted in a larger tunnel structure (2.3 × 2.3 Å2 vs. 4.6 × 4.6 Å2) [41]. Wang et al. developed a Bi3+-doped α-MnO2 cathode with an enlarged lattice spacing [42]. First, Bi(NO3)3 and MnSO4 were mixed uniformly; then, KMnO4 was added and stirred for 2 h, before being transferred to a 100 mL autoclave and reacting at 120 °C for 12 h. Yan at al. also designed Al-intercalated α-MnO2 using a hydrothermal approach with a narrower electronic bandgap [43]. Interestingly, Al-doped MnO2 exhibited a sea urchin-like morphology with a size of 4.5–5.0 µm and an enlarged interlayer spacing (0.24 nm vs. 0.29 nm). Xiong et al. reported that Al-doped α-MnO2 coated with lignin was formed through a hydrothermal reaction involving KMnO4, NH4F, Al2(SO4)3, and lignosulfonate at 200 °C, as illustrated in Figure 2a [44].

2.2. Ion Penetration/Exchange Method

The use of manganese oxides as precursors and the subsequent introduction of guest ions through a post-treatment process are more straightforward ideas for the ion doping strategy. Dai et al. developed a porous HxMn2O4 cathode using a cation exchange strategy, which exhibited a novel crystal structure with an excellent cycle stability (1000 cycles at 1 A·g−1) [45]. First, they used ZnSO4 as the zinc source, MnSO4 as the manganese source, and NH4HCO3 for the coprecipitation reaction to obtain ZnCO3-MnCO3 composites, followed by high-temperature treatment at 600 °C for 3 h to obtain the ZnMn2O4 precursor, which was finally dispersed into 0.5 M H2SO4 ion exchange solution for 12 h to get HxMn2O4. The mechanism of zinc-ion extraction by H+ was the disproportionation of Mn3+ and [ZnO4] tetragonal distortion. This distinctive spinel-type cathode offered new opportunities for long-life span ZIBs. Banerjee et al. reported that a Cu-intercalated MnO2 layered cathode was attained by mixing the prepared MnO2 powder with a 1 M CuSO4 solution for 48 h [46]. The penetration of Cu2+ into δ-MnO2 resulted in an enlarged lattice spacing, thereby lowering the charge transfer resistance. Furthermore, they exploited the redox potential of Cu for full capacity using two electrons. A cobalt-modified δ-MnO2 with a redox-active surface showed superior self-recovery capability, as reported by Shao et al. [47]. As shown in Figure 2b, a molten-salt method was adopted to synthesize δ-MnO2 using MnSO4·H2O and NaNO3 as the reactants, and then δ-MnO2 powder was mixed with 1 M CoCl2 aqueous solution by constant stirring for 8 h at room temperature. The deposition–dissolution mechanism was proven by the electrochemical performance (over 500 mAh·g−1), and Co2+ played a catalytic role in the electrochemical deposition of Mn2+. In addition, a Mn2+ additive was introduced into the electrolyte for enhanced cycle-stability.

2.3. Electrodeposition Method

Electrodeposition methods often involve depositing an electrolyte onto a conductive substrate by applying a certain current or voltage, which has the advantage of outstanding conductivity because of the highly conductive substrate. Dai et al. prepared a Na+-doped MnO2@GCF cathode via the electrodeposition of 0.1 M Na2SO4 and 0.05 M MnSO4·H2O onto a graphene-like carbon film (GCF) [48]. The cathode was synthesized through a two-step procedure. As illustrated in Figure 2c, they first transformed the raw graphite paper into GCF by electrodepositing it into H2SO4 electrolyte, which had a 2D–3D hybrid network composed of graphene sheets. Then, the H2SO4 electrolyte was replaced by NaSO4 and MnSO4·H2O to obtain Na+-doped MnO2@GCF. The prepared cathode achieved excellent energy density (511.9 Wh·kg−1 at 137 W·kg−1). A Co–MnO2 membrane was electrodeposited onto N-decorated carbon cloth (N-CC) by Nakayama et al. [49] in 2020, using an electrolyte consisting of MnSO4, ZnSO4, and CoSO4. Furthermore, the cathode delivered an impressive capacity of 280 mAh·g−1, even at 1.2 A·g−1. Wang et al. reported the electrodeposition synthesis of multivalence cobalt-doped Mn3O4 (Co-Mn3O4) [50]. Similarly, a pretreated carbon cloth was applied as the substrate, while the cobalt and manganese sources were Co(CH3COO)2·4H2O and Mn(CH3COO)2·4H2O, respectively. Moreover, cobalt was present in multiple valence forms in the manganese oxides and played different roles, resulting in improved charge/ion transport and enhanced structure stability.

2.4. Calcination Treatment

Calcination treatment can provide high kinetics for guest ion intercalation, which is also applicable for manganese-based oxide doping strategies. Low-bandgap NixMn3−xO4 nanoparticles were synthesized by Guo et al. through different calcination processes using manganese acetate as the manganese source and nickel acetate as the additive [51]. A Ni–Mn-layered double hydroxide-derived Ni-doped Mn2O3 (NM) was developed by Huang et al. [52]. First, the precursor Ni–Mn-LDH was formed by adding ammonia to a mixture of Ni(NO3)2·6H2O, MnSO4·H2O, and NH4F for coprecipitation at room temperature, and then Ni-doped Mn2O3 was obtained by calcining the precursor at 450 °C, as depicted in Figure 2d. A metal–organic framework template strategy was adopted by Sun et al. to synthesize a N-doped Mn-based cathode (MnOx@N-C) [53]. Firstly, MnO2 was generated by decomposing potassium permanganate under an acidic environment, and then MnO2 was mixed with PVP, Zn(NO3)2·6H2O, and 2-methylimidazole at room temperature to produce PVP-modified MnO2@ZIF-8. Finally, MnOx@N-C was obtained via calcination of MnO2@ZIF-8 at 700 °C. Xia et al. fabricated a N-doped MnO2–x cathode by calcining MnO2 at 200 °C under an NH3 atmosphere [54]. MnO2 was deposited on TiC/C via KMnO4 decomposition, while N doping was processed by NH3 treatment at low temperature. Li et al. designed a N-doped Na2Mn3O7 (N-NMO) in combination with sodium pre-intercalation and nitrification strategies [55]. In the first step, they used a chemical reaction involving KMnO4, C6H12O6, and NaKC4H4O6 to synthesize rugby-type MnCO3 particles as precursors. Next, Na2Mn3O7 (NMO) was obtained by calcining MnCO3 and NaNO3 with a molar ratio of 3:2 at 600 °C for 4 h. Finally, N was introduced into NMO via further calcination under an ammonia atmosphere. Sun et al. reported that sulfur-doped MnO2 (S-MnO2) nanosheets were obtained using a two-zone furnace for application as a high-performance cathode [56]. The S powder was placed on the upstream side under a temperature of 450 °C, while MnO2 was placed on the downstream side under a temperature of 250 °C. This process was maintained for 1 h under Ar atmosphere.

2.5. Other Methods

In addition to the synthesis methods summarized above, several other strategies have been reported. Zn2+-doped MnO2 was also constructed by Wang et al. using KMnO4 as the oxidant and zinc powders as the reducing agent [57]. Lu et al. adopted a facile one-step liquid coprecipitation to compose La–Ca co-doped ε-MnO2 using CaCl2 and La(NO3)3·6H2O as the additives combined with a neutralization reaction of potassium permanganate and manganese sulfate [58]. K0.41MnO2·0.5H2O was synthesized via a conventional sol–gel method by Tao et al. [59]. Mai et al. reported that Ti–MnO2 was obtained by calcining core–shell MnO2@TiO2 nanowires at 450 °C in air for 4 h. MnO2@TiO2 nanowires were synthesized with MnO2 as the substrate and TiCl4 as the additive in a homemade atomic layer deposition system [60]. Cu–MnO was prepared by dissolving δ-MnO2 and Cu(CH3CN)4PF6 in acetone, followed by refluxing at 45 °C for 4 h by Wu et al. [61]. Barpanda et al. combined simple sonochemical and high-temperature calcination techniques to convert MnCO3 into a novel K1.33Mn8O16 [62]. Liu et al. reported a novel Mn3O4@NC cathode, which showed superior cycling stability for ZIBs [63]. As shown in Figure 2e, they first prepared the MnOOH precursor via a KMnO4 decomposition reaction, and then the precursor @PPy was obtained by mixing MnOOH with C18H29SO3Na and ammonium persulfate in a cold bath (0 °C). Finally, the as-prepared sample was sintered at 400 °C for 2 h under Ar flow.

3. Valence-Dependent Dopant Ions in Manganese-Based Oxides

Due to the variable electronegativity of the dopant ions in different valence states, the intercalation sites in the host structure are distinct, performing diverse modifying functions. In this section, pertinent publications about the ion doping of manganese-based oxides reports are discussed and summarized considering the multiple valence states of the doped ions, with the goal of delving into various manganese-based ion doping rules and gaining novel insights. The valence states of dopant ions can be roughly divided into the following categories: (1) monovalent cation doping, (2) multivalent cation doping, and (3) anion doping. In order to facilitate an understanding of the role of different dopant ions in various synthesis strategies, we present their electrochemical performance in Table 1.

3.1. Monovalent Cation-Doped Manganese-Based Oxides

Monovalent ions can often be easily embedded into the host structure because of their low charge density and their large ionic radius tending to enlarge the interlayer spacing after inserting, thereby improving the electrochemical performance and the stability of manganese-based oxides. The monovalent ions used to modify manganese-based oxides are mainly sodium and potassium ions, while other monovalent ions remain largely unexplored.
In 2018, Wang et al. synthesized poorly crystalline MnO2–birnessite at room temperature with a good rate performance (110 mAh·g−1 at 10 °C and 308 mAh·g−1 at 1 °C) and outstanding cycle performance (2000 cycles at 2 A·g−1). However, the roles of Na+ and H2O were not clearly illustrated [64]. In 2019, Zhi et al. discovered that the doping of Na+ ions and water molecules could actively improve the zinc storage capability of δ-MnO2, which delivered excellent electrochemical performance (106 mAh·g−1 at 6 A·g−1 after 10,000 cycles) [65]. Furthermore, they succeeded in making self-healing Zn–δ-MnO2 cells by via with a self-healing substance for electrodes, which could be repaired and restored even after repeated destructive cutting. Carboxylated polyurethane (CPU) and polyacrylamide (PAM) hydrogel electrolytes were used to design flexible batteries, whereby the former was used as the substrate for the electrodes and the latter was used as the quasi-solid-state electrolyte. The assembly structure of the self-healing device and the self-healing process are illustrated in Figure 3a. The self-healing ability was mainly due to carboxyl groups and hydrogen bonds of the CPU substrate being rebuilt after contact, as shown in Figure 3b. Dai et al. fabricated a Na-MnO2@GCF cathode via electrochemical deposition [48]. DFT was used to investigate the effect of sodium ion doping on layered manganese dioxide. The introduction of sodium ions could reduce the binding energy between zinc ions and manganese dioxide, allowing more zinc ions to be combined. The electron domains of oxygen atoms around the zinc ions became more concentrated, further indicating that the electrode material was better for zinc absorption due to the improved electrochemical performance, as illustrated in Figure 3c,d. In 2020, Peng et al. reported the fabrication of Na+ incorporated into layered δ-MnO2 with a uniform array arrangement and high conductivity by adopting nickel foam as a substrate, which delivered an excellent rate performance (131 mAh·g−1 at 6 A·g−1) [38]. Zhu et al. proposed that regulating the coordination of solvent water molecules with Zn2+ could significantly improve the performance of a birnessite cathode (Na0.1MnO2·0.5H2O) [66]. Urea was applied to control the coordination number and geometry of Zn2+ because of its strong coordination, preventing the dissolution of Mn via the interface effect between the cathode and the electrolyte. Wang et al. employed a mechanical cell grinder for doping Na+ into MnO2 and obtained a Na0.44MnO2 cathode, which had a large specific surface area [67]. In 2022, Li et al. reported a N-doped Na2Mn3O7 cathode that displayed a capacity of 300 mAh·g−1 at a current density of 0.2 A·g−1, which even retained a capacity of 100 mAh·g−1 at 10 A·g−1 [55]. The modification principle was that the doping of sodium ions could enlarge the interlayer spacing and serve as a pillar to stabilize the host structure. Nitrogen doping could accelerate electron conduction in the electrodes and suppress the dissolution of Mn during repeated cycling.
In addition to Na+, K+ is usually used for manganese-based oxide doping. In 2019, α-K0.19MnO2 was synthesized through a self-sacrificing template method by Zhang et al. [34]. In order to determine the relationship between the potassium content of the host structure and the electrochemical performance, a concentrated HNO3 treatment was employed to adjust the content of K ions in MnO2. The results showed that the cathode with high potassium content possessed better electrochemical performance, shown in Figure 4a. Moreover, a K-salt was added to the electrolyte to prevent the deintercalation of K+ from the host material, further enhancing the battery performance (Figure 4b). Barpanda et al. reported that K1.33Mn8O16 was prepared via a simple sonochemical technique, adding a new member to the family of manganese-based oxides [62]. Liang et al. reported a K+-stabilized K0.8Mn8O16 with oxygen defects, whereby the incorporation of K+ for suppressing manganese dissolution was confirmed by ICP-OES (Figure 4c), and the oxygen defects improved the ability of electron transfer and opened the [MnO6] octahedral walls for ion diffusion, as illustrated in Figure 4d,e [68]. In 2021, Pan et al. provided a new viewpoint that generating hybrid phases could enhance electrochemical performance, which was confirmed by a layered K0.36H0.26MnO2·0.28H2O (K36) cathode. Via a proton and Zn2+ co-intercalation mechanism, this cathode underwent a gradual phase transition from original K36 to hybrid layer-type KxHyZnzMnO2·nH2O and spinel-type ZnMn2O4 nanocrystals after several cycles [36]. In 2022, Shi et al. reported a K0.29MnO2·0.67H2O cathode with ultralong cycles, whose capacity remained at 158 mAh·g−1 after 12,000 cycles at a high current density of 2000 mA·g−1 [37].
Additional monovalent ion doping work has also been reported. In 2019, Ag1.5Mn8O16 nanorods were prepared via a facile reflux method and successfully applied as a cathode in ZIBs by Takeuchi et al. [69]. A structurally stable newfangled phase of manganese-based oxide MnO2H0.16(H2O)0.27 (MOH) was designed by Pan et al. [35]. As shown in the chemical formula, the intercalation of protons and water molecules into the manganese oxides offered a possibility for improved performance. In 2022, a manganese dioxide (HxMn2O4) with protons inserted was derived from ZnMn2O4 templates via a cation exchange method by Dai et al., which delivered a high capacity of 281 mAh·g−1 at 100 mA·g−1, as well as a good rate performance of 133.4 mAh·g−1 at 1 A·g−1 for 1000 cycles [45].

3.2. Multivalent Cation-Doped Manganese-Based Oxides

Multivalent ions are widely applied in the doping of manganese-based oxides because of their wide variety and variable valence state, providing more possibilities for doping modification. They are classified as cations or anions and summarized according to the presence of single-ion doping or multi-ion co-doping, where the valence state of dopant ions ranges from low to high.
In 2017, Banerjee et al. developed a Cu2+ plug-in Bi–δ-MnO2 that delivered near-full dual-electron capacity after cycling more than 6000 times. The excellent electrochemical performance was attributed to the valence state changes of Cu and Bi during charge and discharge, as shown in Figure 5a [46]. In 2020, Cu–MnO and Cu–Mn2O3 were respectively reported by Hwang et al. and Zhang et al., both of which improved performance by introducing oxygen defects through copper doping [61,70]. Following the reports of NixMn3−xO4 by Guo et al. [51] in 2019 and Ni-doped Mn2O3 by Huang et al. [52] in 2021, Pan et al. proposed a Grotthuss proton transport mechanism for the first time, which was verified by Ni doping of substitute Mn in α-MnO2 [40]. According to this mechanism, a greater disordered degree of tetragonal–orthorhombic (TO) states results in better diffusion kinetics of protons in the tunnels. Ni dopants could promote the TO distortion of the lattice during discharge, resulting in the improvement of electrochemical performance, as illustrated in Figure 5b. A Zn2+-stabilized layer-type MnO2 structure was reported by Wang et al. in 2019 [57], where hydrated Zn2+ worked as pillars to enhance structural stability. In 2020, Shao et al. reported a cobalt-modified δ-MnO2 cathode that assisted the deposition–dissolution mechanism, and they proposed that cobalt doping played a role in catalysis [47]. In 2021, Wang et al. explained the effect and mechanism of Co with different valence states in Mn-based materials for zinc storage, which was confirmed by constructing a multivalent cobalt (Co2+, Co3+)-doped Mn3O4 nanosheet [50]. Tao et al. designed a Ca0.28MnO2·0.5H2O cathode material in 2020, which exhibited an ultralong life of 5000 cycles with no apparent capacity decay [39]. In 2019, Du et al. found that cerium doping could transform β-MnO2 to α-MnO2, which was accompanied by a significant improvement in conductivity and stability [41]. In the same year, birnessite δ-MnO2 nanoflowers with La3+ doping were reported by Lu et al. [71]. The intercalation of La3+ into δ-MnO2 (LMO) could enlarge the interlamellar spacing and lower the resistance of Zn2+ (de)insertion, thereby remarkably enhancing the performance of LMO. A Bi-doped α-MnO2 electrode was reported by Wang et al. in 2021 [42]. They insisted that Bi doping could catalyze the conversion of MnOOH to Mn2+, which improved the deposition and dissolution efficiency of manganese, attaining a more stable second discharge platform. Mesoporous Al0.35Mn2.52O4 (AMO) was prepared through Al substitution in Mn3O4 by Shi et al. in 2021 [72]. As shown in Figure 5c, Al was first introduced into Mn3O4 to replace Mn sites through a simple hydrothermal method. Subsequently, cation defects were obtained by selectively etching Al away from Al0.35Mn2.52O4. Thus, the electronic conductivity and diffusion kinetics of H+ and Zn2+ were both significantly improved as a result of the high specific capacity (302 mAh·g−1 at 0.1 A·g−1) and outstanding cycling performance (147 mAh·g−1 after 1000 cycles at 1.5 A·g−1). Yan et al. constructed an Al-intercalated MnO2 cathode, proposing the critical role of Al doping, which was confirmed by DFT calculations and in situ Raman investigation [43]. Recently, an Fe-doped α-MnO2 coated with polypyrrole was created by Kong et al. [73]. The doping of Fe3+ could effectively enlarge the lattice spacing of α-MnO2 and improve the diffusion kinetics of Zn2+, which led to a better electrochemical performance. In 2019, Ti doping was introduced to build an oriented electric field by Mai et al., achieving impressive performance [60]. As early as 2017, V5+ doping was adopted by Kim et al. to improve the conductivity of manganese-based oxides [74].
While the single-ion doping modification strategy has achieved good results, dual-ion doping has also emerged. In 2020, a La–Ca co-doped ε-MnO2 cathode was obtained through a simple one-step liquid coprecipitation process by Lu et al. [58]. Their experimental results revealed that both La3+ and Ca2+ played a role in enhancing capacity and reversibility. Wang et al. reported that Ni and Co co-substitution of spinel ZnMn2O4@N-rGO was an effective approach to suppress the Jahn–Teller distortion of Mn3+, thereby stabilizing the structure [75].

3.3. Anion-Doped Manganese-Based Oxides

In addition to multivalent cation doping, anion doping is regarded as an effective strategy, e.g., N, P, and S. N is always doped in a carbon-containing substrate and then compounded with manganese oxides to improve performance. In 2017, Li et al. reported a 3D porous MnO2 supported by an ideal N-doped carbon cloth substrate, which delivered a high capacity of 353 mAh·g−1 and remarkable cycle capability (93.6% retention after 1000 cycles) [76]. In 2018, a porous N-C@MnOx cathode derived through MOF-modified manganese oxides was reported for the first time by Sun et al. [53]. The cathode material exhibited a superior electrochemical performance of 305 mAh·g−1 at 500 mA·g−1 after 1600 cycles at 2000 mA·g−1, which benefited from several factors. On the one hand, the onion-shaped N-doped amorphous carbon layers established a high-speed path for electron conduction and maintained the stability of the material host structure. On the other hand, the porous structure of N-C@MnOx accelerated electrolyte infiltration and ion transfer. In 2019, Mn3O4@N-doped carbon matrix composite nanorods and a ZnMn2O4/N-doped graphene nanocomposite were reported by Kong et al. and Yang et al., respectively [63,77]. In 2020, ultrathin MnO2 nanoflakes grown on N-doped hollow carbon spheres were prepared by Yang et al., which displayed high capacity (349 mAh·g−1 at 0.1 A·g−1) and long-term cycling performance (2000 cycles at 2 A·g−1) [78]. Compared with N doping of carbon-based modification layers, the introduction of N into the bulk phase for modification is relatively rare. In 2019, Xia et al. fabricated N–MnO2−x by introducing N into the bulk phase of MnO2 via ammonia calcination at 200 °C, achieving excellent electrochemical performance [54]. In 2020, oxygen vacancies and P ion doping were generated simultaneously by Zheng et al. in a P–MnO2-x@VMG cathode via phosphorization [79]. The oxygen vacancies could increase the electrical conductivity of MnO2, while the P ions were able to expand its interlayer spacing, thereby accelerating ion transport. Recently, sulfur-doped MnO2 (S-MnO2) nanosheets were applied as a cathode for ZIBs by Sun et al., which showed a high capacity (324 mAh·g−1 at 0.2 A·g−1) and excellent rate performance (205 mAh·g−1 at 2 A·g−1) [56].

4. Optimization Mechanism of Ion Doping in Zinc–Manganese Battery

Ion doping alters the behavior of electrode materials in a variety of ways. It is vital to produce a complete overview to develop better electrode materials and identify knowledge gaps for in-depth research in the future. As far as the current research progress is concerned, the positive effects of ion doping can be roughly divided into three categories: (1) enlarged interlayer spacing for improved ion diffusion kinetics, (2) defect engineering for enhanced electrical conductivity, and (3) pillar effect for enhanced stability of the host structure.

4.1. Enlarged Interlayer Spacing for Improved Ion Diffusion Kinetics

Theoretically, Zn2+ ions have a small ionic radius (0.74 Å) and high ionic conductivity in aqueous solution (~1–10 mS·cm−1) [80]; however, in practice, due to their high charge density, Zn2+ ions combine with water molecules to form hydrated [Zn(H2O)6]2+, leading to an increment in ionic radius to 5.5 Å, slowing down the diffusion of Zn2+. Furthermore, the solid electrostatic effect between Zn2+ and the host structure of the cathode material also causes sluggish Zn2+ intercalation [81,82]. The diffusion rate of carriers has a linear negative relationship with the electrostatic repulsion (ƒ) between the carriers and the host structure. According to the formula ƒ 1 ε r r 0 2 , where ε r is the permittivity and r0 is the distance between Zn2+ and the closest ions, a larger value of r0 means faster diffusion kinetics [21,83]. In other words, a larger layer spacing leads to better diffusion dynamics. Ion doping is an efficient strategy to expand the layer spacing of the cathode material, thus enhancing performance.
Kim et al. reported that V-doped MnO2 (VMO) could enhance zinc storage properties by expanding the layer spacing [74]. The (211) peak in the X-ray diffraction (XRD) patterns of VMO showed a minor shift toward lower scanning angles, as shown in Figure 6a, confirming anisotropy of the unit cell parameters, which would facilitate the insertion of zinc ions. Lu’s group obtained a cathode with a larger interlamellar spacing by doping La3+ into δ-MnO2 (LMO), which showed lower resistance of Zn2+ (de)insertion and better structural stability [71]. The rate performance of LMO was significantly improved (121.8 mAh·g−1 at 1.6 A·g−1) compared to pristine δ-MnO2 (only 3.4 mAh·g−1 at 1.6 A·g−1). Zheng et al. reported that phosphate ion-doped MnO2 could expand the interlayer spacing of the (001) plane from 0.68 nm to 0.70 nm, accelerating ion transfer. Simultaneously, oxygen vacancies were introduced via phosphorization, enhancing the electrical conductivity of MnO2 [79]. Wang’s work revealed that the pre-intercalation of Bi3+ into α-MnO2 could effectively enlarge the lattice spacing and have a positive effect on the ion diffusion rates, resulting in a superior rate performance with a capacity retention of 150 mAh·g−1 at 5 A·g−1 [42]. K0.29MnO2·0.67H2O (KMO) with an interplanar spacing of 7.4 Å was synthesized via a simple hydrothermal strategy by Shi et al. [37], as shown in Figure 6b, exhibiting high capacity (300 mAh·g−1 at 0.2 A·g−1) and an ultralong cycle performance (158 mAh·g−1 after 12,000 cycles at 2 A·g−1). According to the XRD pattern in Figure 6c, this work calculated that the interlayer spacing corresponding to the (001) plane increased from 6.8 Å to 7.4 Å according to Bragg’s rule. The diffusion energy barriers of H+ and Zn2+ in MnO2 and KMO were explored using density functional theory (DFT)-based first-principles calculations, and the results revealed a lower value of KMO (0.11 eV and 0.19 eV) than MnO2 (0.32 eV and 0.49 eV), as shown in Figure 6d, indicating that the increased interlayer spacing indeed accelerated ion transfer. The kinetic behavior of the KMO sample was further investigated using the galvanostatic intermittence titration technique (GITT). As shown in Figure 6e,f, KMO displayed a smaller overpotential and higher diffusion coefficient than MnO2 during the discharge process, which indicated that the doping of K+ indeed promoted ion diffusion kinetics.

4.2. Defect Engineering for Enhanced Electrical Conductivity

For secondary batteries, electron transfer between the cathode and anode is an integral part of completing the whole electrochemical reaction; hence, the electrical conductivity of the cathode plays an important role in the electrochemical performance [84]. However, manganese-based oxides are usually semiconductors with poor electrical conductivity [85]. The strategy of complexing with conductive agents is generally adopted to accelerate electron transfer, while ion doping is another method to enhance the electronic conductivity of cathode materials [86]. For example, a distinctive N-doped MnO2−x cathode with numerous oxygen defects was prepared through NH3 treatment at 200 °C by Xia et al. in 2019 [54]. Oxygen vacancies were introduced at the same time as N doping, which increased the electron density and lowered the bandgap of manganese dioxide, resulting in a better electronic conductivity and activity. As shown in Figure 7a, the position of the absorption edge corresponding to the oxidation of N-doped MnO2−x in the XANES spectrum presented a shift toward a lower energy, indicating higher average electron density. On the other hand, the FT spectrum implied that N-doped MnO2 did not change phase but increased its level of disorder. The DFT calculation results (Figure 7b) showed that N-doped MnO2–x possessed a much smaller bandgap (0.12 eV) than pure MnO2 (1.83 eV), revealing a significant enhancement of electronic conductivity. Excellent electrochemical performance was achieved that 285 mAh·g−1 at 0.2 A·g−1 with 85.7% retention after 1000 cycles at 1 A·g−1. In the same year, Ti–MnO2 with oxygen vacancies was reported by Mai’s group [60], indicating that the replacement of manganese with titanium and the introduction of oxygen vacancies could break through the manganese–oxygen octahedral walls, resulting in heterogeneous charge distribution. As revealed by the EIS spectrum (Figure 7c), Ti-doped MnO2 exhibited lower charge migration resistance, confirming that the unbalanced local electric field in the host structure could boost the mobility of ions/electrons. Furthermore, according to the DFT calculations (Figure 7d), the electron cloud of Ti substitution and its derived oxygen vacancies could balance the disordered interfacial electric field, allowing electron transit through the [MnO6] octahedral walls. Liang et al. proposed a K+-stabilized Mn-based cathode with rich oxygen defects (K0.8Mn8O16 with oxygen defects), which exhibited impressive stability over 1000 cycles with no obvious fading [68]. As described in Figure 4d,e, oxygen defects could accelerate H+ diffusion by opening the [MnO6] octahedral walls from the ab-plane. Moreover, the oxygen defects could reduce the energy for electron and charge transfer during the redox reaction, as illustrated in Figure 7e, KMO showed smaller overpotential gaps than pure MnO2 (1.399/1.614 V vs. 1.389/1.612 V). More recently, Zhang et al. designed a cathode (Ocu–Mn2O3) by replacing sites of trivalent manganese with divalent copper ions to create oxygen defects in Mn2O3 for better electronic conductivity [70]. Long et al. fabricated a low-bandgap cathode (NixMn3−xO4) via the replacement of Mn with Ni. The DOS indicated that Ni-doped Mn3O4 exhibited a narrower bandgap than pure Mn3O4 (1.20 eV), thereby significantly enhancing the electronic conductivity [51].

4.3. Pillar Effect to Stabilize the Host Structure

Manganese-based oxide cathode materials inevitably suffer from the dissolution of manganese caused by the Jahn–Teller effect, disproportionation, and irreversible phase transformation during the charge/discharge process [74,87,88], resulting in structural instability. This is essentially due to the unavoidable presence of some trivalent manganese during the charge/discharge cycle, and Mn3+ is extremely unstable because of its high-spin d4 (= t2g3eg1) electronic configuration in a symmetric octahedron [21,89,90], tending to transform into Mn4+ and Mn2+. Compared to the Mn4+–O bond length (1.93 Å), the Mn2+–O/Mn3+–O bonds are longer (2.23 Å/2.045 Å), which weakens the Mn–O bonding strength and increases the disorder degree of the crystal structure [91]. After the insertion of Zn2+, the [Mn3+O6] octahedrons or [Mn2+O4] tetrahedrons can be easily extruded by the strong electrostatic repulsion between the carrier ions and host framework, resulting in manganese dissolution or irreversible phase transition, which is further reflected in the rapid capacity decay [21]. Ion doping is considered as an effective strategy to prevent material frameworks from collapsing, whereby guest ions can behave as pillars to help support the host structure.
Zhi et al. pre-intercalated Na ions and water molecules in δ-MnO2 (δ-NMOH) to obtain a stable cathode [65]. As shown in Figure 8a, the layered structure of δ-NMOH consisted of two parts; one part was composed of manganese–oxygen octahedra connected by edge-sharing to form a layer, and the other part was composed of sodium ions and water molecules acting as pillars in the middle of the layer. Due to the stabilizing effect of Na+ and H2O molecules, when H+ and Zn2+ were inserted/extracted, the host structure could only expand or shrink the interlayer spacing, whereas no change in crystal structure would occur, resulting in better structural stability. An impressive cycle life of 10,000 cycles with 98% retention was measured (Figure 8b) [65]. Wang et al. constructed Zn2+ ion-stabilized MnO2 (ZMO) nanospheres for a cathode with a long lifespan. The ZMO electrode manifested a rate capacity of 124 mAh·g−1 at 3.0 A·g−1, attributed to its high surface area and unique mesoporous texture for sufficient active sites and electrolyte permeation. Furthermore, an outstanding cyclability over 2000 cycles was achieved by the layered-type structure filled with zinc ions for enhanced stability [57]. Zhang et al. reported a K+-doped α-MnO2 (α-K0.19MnO2) cathode with a high K content, which exhibited high rate performance (20 °C, 113 mAh·g−1) [34]. As shown in Figure 8c, K+ was still located in the α-MnO2 tunnel after the water-solvated H+ was inserted into the host framework; while the subsequent intercalation of Zn2+ caused a transformation of MnO2 from the tunnel to the layer, the pre-intercalation K+ remained in the interlayer to stabilize the structure and expand the channel for zinc-ion migration. In 2021, a multivalent cobalt-doped Mn3O4 (Co–Mn3O4) with an outstanding cyclability over 1000 cycles at 2000 mA·g−1 was reported by Wang et al. [50]. Co ions with different valence states played important roles in the phase change charge product (Co–MnO2) and discharge products (Co–ZnMn2O4 and Co–MnOOH). The Co2+ ions doped between the [MnO6] octahedrons on both sides of the δ-MnO2 layers acted as an interlayer pillar due to their stronger adsorption energy, as shown in Figure 8d.

5. Summary and Perspectives

AZIBs have great application prospects in the field of large-scale energy storage due to their safety, environmental friendliness, and low cost. Among various cathode materials, manganese-based compounds are most likely to be commercialized due to their high capacity and suitable potential. However, the low electrical conductivity and structural instability of manganese-based materials caused by the Jahn–Teller effect greatly limit their development. It is well known that the structure of a material determines its properties. Due to modification of the crystal structure, the ion doping optimization strategy is regarded as the most effective approach for the construction of high-performance manganese-based cathodes. In this review, on the basis of the redox mechanism, we divided the synthetic principles of ion doping strategies into four categories and summarized them in detail so as to inspire the design of synthetic routes. According to the summary of the different valence states of the doping ions, we believe that, even though both monovalent ion and multivalent ion doping can improve the performance of manganese-based materials, it has been reported that only multivalent ion doping can replace manganese sites or play a catalytic role in the process of zinc storage [47,52,60,72], which shows that multivalent ions present more possibilities than monovalent ions. We also focused on the effect of ion doping strategies on the electrochemical performance of manganese-based cathode materials and provided a detailed summary of their research progress. Regardless of the type of dopant ion, such as cation vs. anion or monovalent vs. multivalent ion, the interlayer spacing can be expanded, the ion transport rate can be improved, defects can be introduced for high electronic conductivity, and a pillar can be established to stabilize the structure, thereby enhancing the cycle performance and rate capability of the ZIBs. Despite many successful cases demonstrated in previous studies, challenges and opportunities for the ion doping modification strategy still exist, which deserve further in-depth research and discussion.
Firstly, the identification of the exact potion of the intercalated ions in the MnOx crystal structure is crucial in understanding the chemical state of Mn, as well as the effect of the intercalated ions on the zinc/proton intercalation into the host structure.
Secondly, a deep understanding of the influence of the pre-intercalated ions in the manganese-based oxides is required through in situ characterizations, including in situ XRD, in situ TEM, in situ Raman spectroscopy, and in situ FT-IR spectroscopy, to reveal the fundamental mechanism.
Thirdly, although ion doping greatly improves the electrochemical performance of cathode materials, the current status of manganese oxide cathodes is still in its infancy, operating at relative low mass loading (below 1 mg/cm2, corresponding to an areal capacity of 0.1–0.2 mAh/cm2); hence, we are far from the realizing the commercialization of zinc-ion batteries, requiring an areal capacity of over 2 mAh/cm2. In this sense, a mass loading above 10 mg/cm2 is highly needed to achieve a practical zinc-ion battery. Additionally, in terms of a high-mass-loading cathode, the utilization of a zinc metal anode is another typical challenge to achieving a high energy density of the zinc anode.
Lastly, most synthesis methods are too complex and expensive for large-scale industrial production; therefore, feasible and low-cost doping methods should be urgently developed on the basis of the actual conditions.

Author Contributions

Conceptualization, B.Z., L.Z. and K.Z.; validation, B.Z., W.S. and K.Z.; investigation, B.Z., J.C. and W.S.; resources, L.Z. and K.Z; data curation, B.Z., Y.S. and L.Z.; writing—original draft preparation, B.Z. and J.C.; writing—review and editing, L.Z. and K.Z.; supervision, L.Z. and K.Z.; project administration, K.Z.; funding acquisition, L.Z. and K.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Natural Science Foundation of China (21905169 and 22109123) and the Sanya Science and Education Innovation Park of Wuhan University of Technology (2020KF0022 and 2021KF0020).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chu, S.; Majumdar, A. Opportunities and challenges for a sustainable energy future. Nature 2012, 488, 294–303. [Google Scholar] [CrossRef] [PubMed]
  2. Hosenuzzaman, M.; Rahim, N.A.; Selvaraj, J.; Hasanuzzaman, M.; Malek, A.B.M.A.; Nahar, A. Global prospects, progress, policies, and environmental impact of solar photovoltaic power generation. Renew. Sustain. Energy Rev. 2015, 41, 284–297. [Google Scholar] [CrossRef]
  3. Meng, J.; Guo, H.; Niu, C.; Zhao, Y.; Xu, L.; Li, Q.; Mai, L. Advances in structure and property optimizations of battery electrode materials. Joule 2017, 1, 522–547. [Google Scholar] [CrossRef] [Green Version]
  4. Sterl, S.; Vanderkelen, I.; Chawanda, C.J.; Russo, D.; Brecha, R.J.; van Griensven, A.; van Lipzig, N.P.M.; Thiery, W. Smart renewable electricity portfolios in West Africa. Nat. Sustain. 2020, 3, 710–719. [Google Scholar] [CrossRef]
  5. Larcher, D.; Tarascon, J.M. Towards greener and more sustainable batteries for electrical energy storage. Nat. Chem. 2015, 7, 19–29. [Google Scholar] [CrossRef] [PubMed]
  6. Tang, B.; Shan, L.; Liang, S.; Zhou, J. Issues and opportunities facing aqueous zinc-ion batteries. Energy Environ. Sci. 2019, 12, 3288–3304. [Google Scholar] [CrossRef]
  7. Blanc, L.E.; Kundu, D.; Nazar, L.F. Scientific challenges for the implementation of Zn-ion batteries. Joule 2020, 4, 771–799. [Google Scholar] [CrossRef]
  8. Goodenough, J.B.; Park, K.S. The Li-ion rechargeable battery: A perspective. J. Am. Chem. Soc. 2013, 135, 1167–1176. [Google Scholar] [CrossRef]
  9. Cao, Y.; Li, M.; Lu, J.; Liu, J.; Amine, K. Bridging the academic and industrial metrics for next-generation practical batteries. Nat. Nanotechnol. 2019, 14, 200–207. [Google Scholar] [CrossRef]
  10. Liu, T.; Dai, A.; Lu, J.; Yuan, Y.; Xiao, Y.; Yu, L.; Li, M.; Gim, J.; Ma, L.; Liu, J.; et al. Correlation between manganese dissolution and dynamic phase stability in spinel-based lithium-ion battery. Nat. Commun. 2019, 10, 4721. [Google Scholar] [CrossRef] [Green Version]
  11. Zhao, K.; Sun, C.; Yu, Y.; Dong, Y.; Zhang, C.; Wang, C.; Voyles, P.M.; Mai, L.; Wang, X. Surface gradient Ti-doped MnO2 nanowires for high-rate and long-life lithium battery. ACS Appl. Mater. Interfaces 2018, 10, 44376–44384. [Google Scholar] [CrossRef] [PubMed]
  12. Evarts, E.C. Lithium batteries: To the limits of lithium. Nature 2015, 526, S93–S95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Chen, L.; An, Q.; Mai, L. Recent advances and prospects of cathode materials for rechargeable aqueous zinc-ion batteries. Adv. Mater. Interfaces 2019, 6, 1900387. [Google Scholar] [CrossRef]
  14. Huang, J.; Guo, Z.; Ma, Y.; Bin, D.; Wang, Y.; Xia, Y. Recent progress of rechargeable batteries using mild aqueous electrolytes. Small Methods 2019, 3, 1800272. [Google Scholar] [CrossRef] [Green Version]
  15. Chao, D.; Zhou, W.; Xie, F.; Ye, C.; Li, H.; Jaroniec, M.; Qiao, S.Z. Roadmap for advanced aqueous batteries: From design of materials to applications. Sci. Adv. 2020, 6, eaba4098. [Google Scholar] [CrossRef]
  16. Xiong, F.; Jiang, Y.; Cheng, L.; Yu, R.; Tan, S.; Tang, C.; Zuo, C.; An, Q.; Zhao, Y.; Gaumet, J.J. Low-strain TiP2O7 with three-dimensional ion channels as long-life and high-rate anode material for Mg-ion batteries. Interdiscip. Mater. 2022, 1, 140–147. [Google Scholar] [CrossRef]
  17. Li, C.; Xie, X.; Liang, S.; Zhou, J. Issues and future perspective on zinc metal anode for rechargeable aqueous zinc-ion batteries. Energy Environ. Mater. 2020, 3, 146–159. [Google Scholar] [CrossRef]
  18. Shin, J.; Lee, J.; Park, Y.; Choi, J.W. Aqueous zinc ion batteries: Focus on zinc metal anodes. Chem. Sci. 2020, 11, 2028–2044. [Google Scholar] [CrossRef] [Green Version]
  19. Huang, S.; Zhu, J.; Tian, J.; Niu, Z. Recent progress in the electrolytes of aqueous zinc-ion batteries. Chem. Eur. J. 2019, 25, 14480–14494. [Google Scholar] [CrossRef]
  20. Shi, W.; Lee, W.S.V.; Xue, J. Recent development of Mn-based oxides as zinc-ion battery cathode. ChemSusChem 2021, 14, 1634–1658. [Google Scholar] [CrossRef]
  21. Zhao, Q.; Song, A.; Ding, S.; Qin, R.; Cui, Y.; Li, S.; Pan, F. Preintercalation strategy in manganese oxides for electrochemical energy storage: Review and prospects. Adv. Mater. 2020, 32, e2002450. [Google Scholar] [CrossRef] [PubMed]
  22. Ming, J.; Guo, J.; Xia, C.; Wang, W.; Alshareef, H.N. Zinc-ion batteries: Materials, mechanisms, and applications. Mater. Sci. Eng. R Rep. 2019, 135, 58–84. [Google Scholar] [CrossRef]
  23. Yu, P.; Zeng, Y.; Zhang, H.; Yu, M.; Tong, Y.; Lu, X. Flexible Zn-ion batteries: Recent progresses and challenges. Small 2019, 15, e1804760. [Google Scholar] [CrossRef] [PubMed]
  24. Zhou, L.; Zhuang, Z.; Zhao, H.; Lin, M.; Zhao, D.; Mai, L. Intricate Hollow Structures: Controlled Synthesis and Applications in Energy Storage and Conversion. Adv. Mater. 2017, 29, 1602914. [Google Scholar] [CrossRef] [PubMed]
  25. Guo, C.; Liu, H.; Li, J.; Hou, Z.; Liang, J.; Zhou, J.; Zhu, Y.; Qian, Y. Ultrathin δ-MnO2 nanosheets as cathode for aqueous rechargeable zinc ion battery. Electrochim. Acta 2019, 304, 370–377. [Google Scholar] [CrossRef]
  26. Wang, J.; Wang, J.G.; Qin, X.; Wang, Y.; You, Z.; Liu, H.; Shao, M. Superfine MnO2 nanowires with rich defects toward boosted zinc ion storage performance. ACS Appl Mater. Interfaces 2020, 12, 34949–34958. [Google Scholar] [CrossRef]
  27. Baral, A.; Satish, L.; Zhang, G.; Ju, S.; Ghosh, M.K. A review of recent progress on nano MnO2: Synthesis, surface modification and applications. J. Inorg. Organomet. Polym Mater. 2020, 31, 899–922. [Google Scholar] [CrossRef]
  28. Xiong, T.; Zhang, Y.; Lee, W.S.V.; Xue, J. Defect engineering in manganese-based oxides for aqueous rechargeable zinc-ion batteries: A review. Adv. Energy Mater. 2020, 10, 2001769. [Google Scholar] [CrossRef]
  29. Xu, W.W.; Sun, C.L.; Wang, N.; Liao, X.B.; Zhao, K.N.; Yao, G.; Sun, Q.C.; Cheng, H.W.; Wang, Y.; Lu, X.G. Sn stabilized pyrovanadate structure rearrangement for zinc ion battery. Nano Energy 2021, 81, 105584. [Google Scholar] [CrossRef]
  30. Yao, X.; Zhao, Y.; Castro, F.A.; Mai, L. Rational design of preintercalated electrodes for rechargeable batteries. ACS Energy Lett. 2019, 4, 771–778. [Google Scholar] [CrossRef]
  31. Kitchaev, D.A.; Dacek, S.T.; Sun, W.; Ceder, G. Thermodynamics of phase selection in MnO2 framework structures through alkali intercalation and hydration. J. Am. Chem. Soc. 2017, 139, 2672–2681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Hu, X.; Kitchaev, D.A.; Wu, L.; Zhang, B.; Meng, Q.; Poyraz, A.S.; Marschilok, A.C.; Takeuchi, E.S.; Takeuchi, K.J.; Ceder, G. Revealing and rationalizing the rich polytypism of todorokite MnO2. J. Am. Chem. Soc. 2018, 140, 6961–6968. [Google Scholar] [CrossRef] [PubMed]
  33. Peng, X.; Peng, H.; Zhao, K.; Zhang, Y.; Xia, F.; Lyu, J.; Van Tendeloo, G.; Sun, C.; Wu, J. Direct visualization of atomic-scale heterogeneous structure dynamics in MnO2 nanowires. ACS Appl. Mater. Interfaces 2021, 13, 33644–33651. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, G.; Huang, H.; Bi, R.; Xiao, X.; Ma, T.; Zhang, L. K+ pre-intercalated manganese dioxide with enhanced Zn2+ diffusion for high rate and durable aqueous zinc-ion batteries. J. Mater. Chem. A 2019, 7, 20806–20812. [Google Scholar] [CrossRef]
  35. Zhao, Q.H.; Chen, X.; Yang, L.Y.; Chen, H.B.; Pan, F. Unravelling H+/Zn2+ synergistic intercalation in a novel phase of manganese oxide for high-performance aqueous rechargeable battery. Small 2019, 15, 1904545. [Google Scholar] [CrossRef]
  36. Ding, S.; Liu, L.; Qin, R.; Chen, X.; Song, A.; Li, J.; Li, S.; Zhao, Q.; Pan, F. Progressive “layer to hybrid spinel/layer” phase evolution with proton and Zn2+ co-intercalation to enable high performance of MnO2-based aqueous batteries. ACS Appl. Mater. Interfaces 2021, 13, 22466–22474. [Google Scholar] [CrossRef]
  37. Wang, D.; Zhang, S.; Li, C.; Chen, X.; Wang, W.; Han, Y.; Lin, H.; Shi, Z.; Feng, S. Engineering the interplanar spacing of K-birnessite for ultra-long cycle Zn-ion battery through “hydrothermal potassium insertion” strategy. Chem. Eng. J. 2022, 435, 134754. [Google Scholar] [CrossRef]
  38. Peng, H.; Fan, H.; Yang, C.; Tian, Y.; Wang, C.; Sui, J. Ultrathin δ-MnO2 nanoflakes with Na+ intercalation as a high-capacity cathode for aqueous zinc-ion batteries. RSC Adv. 2020, 10, 17702–17712. [Google Scholar] [CrossRef]
  39. Sun, T.; Nian, Q.; Zheng, S.; Shi, J.; Tao, Z. Layered Ca0. 28MnO2·0.5H2O as a high performance cathode for aqueous zinc-Ion battery. Small 2020, 16, 2000597. [Google Scholar] [CrossRef]
  40. Zhao, Q.; Song, A.; Zhao, W.; Qin, R.; Ding, S.; Chen, X.; Song, Y.; Yang, L.; Lin, H.; Li, S.; et al. Boosting the energy density of aqueous batteries via facile grotthuss proton transport. Angew. Chem. Int. Ed. Engl. 2021, 60, 4169–4174. [Google Scholar] [CrossRef]
  41. Wang, J.; Sun, X.; Zhao, H.; Xu, L.; Xia, J.; Luo, M.; Yang, Y.; Du, Y. Superior-performance aqueous zinc ion battery based on structural transformation of MnO2 by rare earth doping. J. Mater. Chem. C 2019, 123, 22735–22741. [Google Scholar] [CrossRef]
  42. Ma, K.; Li, Q.; Hong, C.; Yang, G.; Wang, C. Bi doping-enhanced reversible-phase transition of alpha-MnO2 raising the cycle capability of aqueous Zn-Mn batteries. ACS Appl. Mater. Interfaces 2021, 13, 55208–55217. [Google Scholar] [CrossRef] [PubMed]
  43. Chen, C.; Shi, M.; Zhao, Y.; Yang, C.; Zhao, L.; Yan, C. Al-intercalated MnO2 cathode with reversible phase transition for aqueous Zn-ion batteries. Chem. Eng. J. 2021, 422, 130375. [Google Scholar] [CrossRef]
  44. Xu, J.; Hu, X.; Alam, M.A.; Muhammad, G.; Lv, Y.; Wang, M.; Zhu, C.; Xiong, W. Al-doped α-MnO2 coated by lignin for high-performance rechargeable aqueous zinc-ion batteries. RSC Adv. 2021, 11, 35280–35286. [Google Scholar] [CrossRef] [PubMed]
  45. Wu, Y.; Zhang, K.; Chen, S.; Liu, Y.; Tao, Y.; Zhang, X.; Ding, Y.; Dai, S. Proton inserted manganese dioxides as a reversible cathode for aqueous Zn-ion batteries. ACS Appl. Energy Mater. 2019, 3, 319–327. [Google Scholar] [CrossRef]
  46. Yadav, G.G.; Gallaway, J.W.; Turney, D.E.; Nyce, M.; Huang, J.; Wei, X.; Banerjee, S. Regenerable Cu-intercalated MnO2 layered cathode for highly cyclable energy dense batteries. Nat. Commun. 2017, 8, 14424. [Google Scholar] [CrossRef] [Green Version]
  47. Zhong, Y.; Xu, X.; Veder, J.P.; Shao, Z. Self-recovery chemistry and cobalt-catalyzed electrochemical deposition of cathode for boosting performance of aqueous zinc-Ion batteries. iScience 2020, 23, 100943. [Google Scholar] [CrossRef] [Green Version]
  48. Wu, Y.; Wang, M.; Tao, Y.; Zhang, K.; Cai, M.; Ding, Y.; Liu, X.; Hayat, T.; Alsaedi, A.; Dai, S. Electrochemically derived graphene-like carbon film as a superb substrate for high-performance aqueous Zn-ion batteries. Adv. Funct. Mater. 2020, 30, 1907120. [Google Scholar] [CrossRef]
  49. Kataoka, F.; Ishida, T.; Nagita, K.; Kumbhar, V.; Yamabuki, K.; Nakayama, M. Cobalt-doped layered MnO2 thin film electrochemically grown on nitrogen-doped carbon cloth for aqueous zinc-ion batteries. ACS Appl. Energy Mater. 2020, 3, 4720–4726. [Google Scholar] [CrossRef]
  50. Ji, J.; Wan, H.; Zhang, B.; Wang, C.; Gan, Y.; Tan, Q.; Wang, N.; Yao, J.; Zheng, Z.; Liang, P.; et al. Co2+/3+/4+-regulated electron state of Mn-O for superb aqueous zinc-manganese oxide batteries. Adv. Energy Mater. 2020, 11, 2003203. [Google Scholar] [CrossRef]
  51. Long, J.; Gu, J.; Yang, Z.; Mao, J.; Hao, J.; Chen, Z.; Guo, Z. Highly porous, low band-gap NixMn3−xO4 (0.55 ≤ x ≤ 1.2) spinel nanoparticles with in situ coated carbon as advanced cathode materials for zinc-ion batteries. J. Mater. Chem. A 2019, 7, 17854–17866. [Google Scholar] [CrossRef]
  52. Zhang, D.; Cao, J.; Zhang, X.; Insin, N.; Wang, S.; Han, J.; Zhao, Y.; Qin, J.; Huang, Y. Inhibition of manganese dissolution in Mn2O3 cathode with controllable Ni2+ incorporation for high-performance zinc ion battery. Adv. Funct. Mater. 2021, 31, 2009412. [Google Scholar] [CrossRef]
  53. Fu, Y.; Wei, Q.; Zhang, G.; Wang, X.; Zhang, J.; Hu, Y.; Wang, D.; Zuin, L.; Zhou, T.; Wu, Y. High-performance reversible aqueous Zn-ion battery based on porous MnOx nanorods coated by MOF-derived N-doped carbon. Adv. Energy Mater. 2018, 8, 1801445. [Google Scholar] [CrossRef]
  54. Zhang, Y.; Deng, S.; Luo, M.; Pan, G.; Zeng, Y.; Lu, X.; Ai, C.; Liu, Q.; Xiong, Q.; Wang, X.; et al. Defect promoted capacity and durability of N-MnO2-x branch arrays via low-temperature NH3 treatment for advanced aqueous zinc ion batteries. Small 2019, 15, e1905452. [Google Scholar] [CrossRef]
  55. Cheng, X.; Xiao, J.; Ye, M.; Zhang, Y.; Yang, Y.; Li, C.C. Achieving stable zinc-ion storage performance of manganese oxides by synergistic engineering of the interlayer structure and interface. ACS Appl. Mater. Interfaces 2022, 14, 10489–10497. [Google Scholar] [CrossRef]
  56. Zhao, Y.; Zhang, P.; Liang, J.; Xia, X.; Ren, L.; Song, L.; Liu, W.; Sun, X. Uncovering sulfur doping effect in MnO2 nanosheets as an efficient cathode for aqueous zinc ion battery. Energy Storage Mater. 2022, 47, 424–433. [Google Scholar] [CrossRef]
  57. Wang, J.; Wang, J.-G.; Liu, H.; Wei, C.; Kang, F. Zinc ion stabilized MnO2 nanospheres for high capacity and long lifespan aqueous zinc-ion batteries. J. Mater. Chem. A 2019, 7, 13727–13735. [Google Scholar] [CrossRef]
  58. Zhang, M.; Wu, W.; Luo, J.; Zhang, H.; Liu, J.; Liu, X.; Yang, Y.; Lu, X. A high-energy-density aqueous zinc–manganese battery with a La–Ca co-doped ε-MnO2 cathode. J. Mater. Chem. A 2020, 8, 11642–11648. [Google Scholar] [CrossRef]
  59. Sun, T.; Nian, Q.; Zheng, S.; Yuan, X.; Tao, Z. Water cointercalation for high-energy-density aqueous zinc-ion battery based potassium manganite cathode. J. Power Sources 2020, 478, 228758. [Google Scholar] [CrossRef]
  60. Lian, S.; Sun, C.; Xu, W.; Huo, W.; Luo, Y.; Zhao, K.; Yao, G.; Xu, W.; Zhang, Y.; Li, Z.; et al. Built-in oriented electric field facilitating durable Zn MnO2 battery. Nano Energy 2019, 62, 79–84. [Google Scholar] [CrossRef]
  61. Fenta, F.W.; Olbasa, B.W.; Tsai, M.-C.; Weret, M.A.; Zegeye, T.A.; Huang, C.-J.; Huang, W.-H.; Zeleke, T.S.; Sahalie, N.A.; Pao, C.-W. Electrochemical transformation reaction of Cu–MnO in aqueous rechargeable zinc-ion batteries for high performance and long cycle life. J. Mater. Chem. A 2020, 8, 17595–17607. [Google Scholar] [CrossRef]
  62. Sada, K.; Senthilkumar, B.; Barpanda, P. Cryptomelane K1.33Mn8O16 as a cathode for rechargeable aqueous zinc-ion batteries. J. Mater. Chem. A 2019, 7, 23981–23988. [Google Scholar] [CrossRef]
  63. Sun, M.; Li, D.S.; Wang, Y.F.; Liu, W.L.; Ren, M.M.; Kong, F.G.; Wang, S.J.; Guo, Y.Z.; Liu, Y.M. Mn3O4@NC composite nanorods as a cathode for rechargeable aqueous Zn-ion batteries. ChemElectroChem 2019, 6, 2510–2516. [Google Scholar] [CrossRef]
  64. Qiu, N.; Chen, H.; Yang, Z.; Sun, S.; Wang, Y. Low-cost birnessite as a promising cathode for high-performance aqueous rechargeable batteries. Electrochim. Acta 2018, 272, 154–160. [Google Scholar] [CrossRef]
  65. Wang, D.; Wang, L.; Liang, G.; Li, H.; Liu, Z.; Tang, Z.; Liang, J.; Zhi, C. A superior delta-MnO2 cathode and a self-healing Zn-delta-MnO2 battery. ACS Nano 2019, 13, 10643–10652. [Google Scholar] [CrossRef]
  66. Hou, Z.; Dong, M.; Xiong, Y.; Zhang, X.; Ao, H.; Liu, M.; Zhu, Y.; Qian, Y. A high-energy and long-life aqueous Zn/birnessite battery via reversible water and Zn2+ coinsertion. Small 2020, 16, e2001228. [Google Scholar] [CrossRef] [PubMed]
  67. Li, J.; Li, L.; Shi, H.; Zhong, Z.; Niu, X.; Zeng, P.; Long, Z.; Chen, X.; Peng, J.; Luo, Z. Electrochemical energy storage behavior of Na0. 44MnO2 in aqueous zinc-ion battery. ACS Sustain. Chem. Eng. 2020, 8, 10673–10681. [Google Scholar]
  68. Fang, G.; Zhu, C.; Chen, M.; Zhou, J.; Tang, B.; Cao, X.; Zheng, X.; Pan, A.; Liang, S. Suppressing manganese dissolution in potassium manganate with rich oxygen defects engaged high-energy-density and durable aqueous zinc-ion battery. Adv. Funct. Mater. 2019, 29, 1808375. [Google Scholar] [CrossRef]
  69. Wang, L.; Wu, Q.; Abraham, A.; West, P.J.; Housel, L.M.; Singh, G.; Sadique, N.; Quilty, C.D.; Wu, D.; Takeuchi, E.S. Silver-containing α-MnO2 nanorods: Electrochemistry in rechargeable aqueous Zn-MnO2 batteries. J. Electrochem. Soc. 2019, 166, A3575. [Google Scholar] [CrossRef]
  70. Liu, N.; Wu, X.; Yin, Y.; Chen, A.; Zhao, C.; Guo, Z.; Fan, L.; Zhang, N. Constructing the efficient ion diffusion pathway by introducing oxygen defects in Mn2O3 for high-performance aqueous zinc-ion batteries. ACS Appl Mater. Interfaces 2020, 12, 28199–28205. [Google Scholar] [CrossRef]
  71. Zhang, H.; Liu, Q.; Wang, J.; Chen, K.; Xue, D.; Liu, J.; Lu, X. Boosting the Zn-ion storage capability of birnessite manganese oxide nanoflorets by La3+ intercalation. J. Mater. Chem. A 2019, 7, 22079–22083. [Google Scholar] [CrossRef]
  72. Wang, D.; Zhang, S.; Li, C.; Chen, X.; Zhang, W.; Ge, X.; Lin, H.; Shi, Z.; Feng, S. High-performance aqueous zinc-ion battery based on an Al0.35Mn2.52O4 cathode: A design Strategy from defect engineering and atomic composition tuning. Small 2022, 18, 2105970. [Google Scholar] [CrossRef] [PubMed]
  73. Xu, J.W.; Gao, Q.L.; Xia, Y.M.; Lin, X.S.; Liu, W.L.; Ren, M.M.; Kong, F.G.; Wang, S.J.; Lin, C. High-performance reversible aqueous zinc-ion battery based on iron-doped alpha-manganese dioxide coated by polypyrrole. J. Colloid Interface Sci. 2021, 598, 419–429. [Google Scholar] [CrossRef] [PubMed]
  74. Alfaruqi, M.H.; Islam, S.; Mathew, V.; Song, J.; Kim, S.; Tung, D.P.; Jo, J.; Kim, S.; Baboo, J.P.; Xiu, Z.; et al. Ambient redox synthesis of vanadium-doped manganese dioxide nanoparticles and their enhanced zinc storage properties. Appl. Surf. Sci. 2017, 404, 435–442. [Google Scholar] [CrossRef]
  75. Tao, Y.; Li, Z.; Tang, L.; Pu, X.; Cao, T.; Cheng, D.; Xu, Q.; Liu, H.; Wang, Y.; Xia, Y. Nickel and cobalt co-substituted spinel ZnMn2O4@N-rGO for increased capacity and stability as a cathode material for rechargeable aqueous zinc-ion battery. Electrochim. Acta 2020, 331, 135296. [Google Scholar] [CrossRef]
  76. Qiu, W.; Li, Y.; You, A.; Zhang, Z.; Li, G.; Lu, X.; Tong, Y. High-performance flexible quasi-solid-state Zn–MnO2 battery based on MnO2 nanorod arrays coated 3D porous nitrogen-doped carbon cloth. J. Mater. Chem. A 2017, 5, 14838–14846. [Google Scholar] [CrossRef]
  77. Chen, L.; Yang, Z.; Qin, H.; Zeng, X.; Meng, J. Advanced electrochemical performance of ZnMn2O4/N-doped graphene hybrid as cathode material for zinc ion battery. J. Power Sources 2019, 425, 162–169. [Google Scholar] [CrossRef]
  78. Chen, L.; Yang, Z.; Cui, F.; Meng, J.; Jiang, Y.; Long, J.; Zeng, X. Ultrathin MnO2 nanoflakes grown on N-doped hollow carbon spheres for high-performance aqueous zinc ion batteries. Mater. Chem. Front. 2020, 4, 213–221. [Google Scholar] [CrossRef]
  79. Zhang, Y.; Deng, S.; Pan, G.; Zhang, H.; Liu, B.; Wang, X.L.; Zheng, X.; Liu, Q.; Wang, X.; Xia, X.; et al. Introducing oxygen defects into phosphate ions intercalated manganese dioxide/vertical multilayer graphene arrays to boost flexible zinc ion storage. Small Methods 2020, 4, 1900828. [Google Scholar] [CrossRef]
  80. Li, J.; McColl, K.; Lu, X.; Sathasivam, S.; Dong, H.; Kang, L.; Li, Z.; Zhao, S.; Kafizas, A.G.; Wang, R.; et al. Multi-scale investigations of δ-Ni0.25V2O5·nH2O cathode materials in aqueous zinc-ion batteries. Adv. Energy Mater. 2020, 10, 2000058. [Google Scholar] [CrossRef]
  81. Kundu, D.; Hosseini Vajargah, S.; Wan, L.; Adams, B.; Prendergast, D.; Nazar, L.F. Aqueous vs. nonaqueous Zn-ion batteries: Consequences of the desolvation penalty at the interface. Energy Environ. Sci. 2018, 11, 881–892. [Google Scholar] [CrossRef]
  82. Liu, Y.; He, G.; Jiang, H.; Parkin, I.P.; Shearing, P.R.; Brett, D.J.L. Cathode design for aqueous rechargeable multivalent ion batteries: Challenges and opportunities. Adv. Funct. Mater. 2021, 31, 2010445. [Google Scholar] [CrossRef]
  83. Yan, M.; He, P.; Chen, Y.; Wang, S.; Wei, Q.; Zhao, K.; Xu, X.; An, Q.; Shuang, Y.; Shao, Y. Water-lubricated intercalation in V2O5·nH2O for high-capacity and high-rate aqueous rechargeable zinc batteries. Adv. Mater. 2018, 30, 1703725. [Google Scholar] [CrossRef] [PubMed]
  84. Zheng, J.; Ye, Y.; Pan, F. ‘Structure units’ as material genes in cathode materials for lithium-ion batteries. Natl. Sci. Rev. 2020, 7, 242–245. [Google Scholar] [CrossRef] [PubMed]
  85. Young, M.J.; Holder, A.M.; George, S.M.; Musgrave, C.B. Charge storage in cation incorporated α-MnO2. Chem. Mater. 2015, 27, 1172–1180. [Google Scholar] [CrossRef] [Green Version]
  86. 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]
  87. Zhang, N.; Cheng, F.; Liu, J.; Wang, L.; Long, X.; Liu, X.; Li, F.; Chen, J. Rechargeable aqueous zinc-manganese dioxide batteries with high energy and power densities. Nat. Commun. 2017, 8, 405. [Google Scholar] [CrossRef] [Green Version]
  88. Alfaruqi, M.H.; Mathew, V.; Gim, J.; Kim, S.; Song, J.; Baboo, J.P.; Choi, S.H.; Kim, J. Electrochemically induced structural transformation in a γ-MnO2 cathode of a high capacity zinc-ion battery system. Chem. Mater. 2015, 27, 3609–3620. [Google Scholar] [CrossRef]
  89. Kobayashi, S.; Kottegoda, I.R.; Uchimoto, Y.; Wakihara, M. XANES and EXAFS analysis of nano-size manganese dioxide as a cathode material for lithium-ion batteries. J. Mater. Chem. 2004, 14, 1843–1848. [Google Scholar] [CrossRef]
  90. Berg, H.; Göransson, K.; Noläng, B.; Thomas, J.O. Electronic structure and stability of the LixMn2O4 (0 < x < 2) system. J. Mater. Chem. 1999, 9, 2813–2820. [Google Scholar]
  91. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. Sect. A. 1976, 32, 751–767. [Google Scholar] [CrossRef]
Figure 1. (a) Crystal structures of various manganese-based oxides [31]. (b) The formation free energies for AxMnO2·H2O when doping with Li+, Na+, K+, Mg2+, and Ca2+ [31].
Figure 1. (a) Crystal structures of various manganese-based oxides [31]. (b) The formation free energies for AxMnO2·H2O when doping with Li+, Na+, K+, Mg2+, and Ca2+ [31].
Energies 15 04698 g001
Figure 2. Synthesis methods of ion-doped manganese-based oxides: (a) hydrothermal synthesis of Al-doped α-MnO2 coated with lignin [44]; (b) ion penetration method for cobalt-modified δ-MnO2 [47]; (c) electrodeposition method for Na+-doped δ-MnO2@GCF [48]; (d) calcination treatment for Ni-doped Mn2O3 [52]; (e) schematic illustration of the synthesis of Mn3O4@NC cathode [63].
Figure 2. Synthesis methods of ion-doped manganese-based oxides: (a) hydrothermal synthesis of Al-doped α-MnO2 coated with lignin [44]; (b) ion penetration method for cobalt-modified δ-MnO2 [47]; (c) electrodeposition method for Na+-doped δ-MnO2@GCF [48]; (d) calcination treatment for Ni-doped Mn2O3 [52]; (e) schematic illustration of the synthesis of Mn3O4@NC cathode [63].
Energies 15 04698 g002
Figure 3. Na+-doped manganese oxides: (a) explanation of the self-healing process of CPU [65]; (b) schematic diagram of the self-healing mechanism for CPU [65]; (c) binding energy (Eb) of Zn2+ with MnO2/Na–MnO2 [48]; (d) difference in charge density of Zn0.1MnO2/Zn0.1Na0.1MnO2 [48].
Figure 3. Na+-doped manganese oxides: (a) explanation of the self-healing process of CPU [65]; (b) schematic diagram of the self-healing mechanism for CPU [65]; (c) binding energy (Eb) of Zn2+ with MnO2/Na–MnO2 [48]; (d) difference in charge density of Zn0.1MnO2/Zn0.1Na0.1MnO2 [48].
Energies 15 04698 g003
Figure 4. K+-doped manganese oxides: (a) rate performance of α-K0.19MnO2 [34]; (b) cycle performance of α-K0.19MnO2 with/without K-salt additive [34]; (c) Mn content analysis in the electrolyte of KMO and α-MnO2 [68]; (d,e) schematic illustration of H+ diffusion in KMO with/without oxygen defects [68].
Figure 4. K+-doped manganese oxides: (a) rate performance of α-K0.19MnO2 [34]; (b) cycle performance of α-K0.19MnO2 with/without K-salt additive [34]; (c) Mn content analysis in the electrolyte of KMO and α-MnO2 [68]; (d,e) schematic illustration of H+ diffusion in KMO with/without oxygen defects [68].
Energies 15 04698 g004
Figure 5. (a) Schematic diagram of electrochemical reaction for Cu2+-doped Bi–birnessite [46]. (b) Mechanistic model of how Ni doping improves the performance of α-MnO2 [40]. (c) Schematic illustration of structural evolution from Mn3O4 to AMO and D-AMO [72].
Figure 5. (a) Schematic diagram of electrochemical reaction for Cu2+-doped Bi–birnessite [46]. (b) Mechanistic model of how Ni doping improves the performance of α-MnO2 [40]. (c) Schematic illustration of structural evolution from Mn3O4 to AMO and D-AMO [72].
Energies 15 04698 g005
Figure 6. Enlarged interlayer spacing and faster ion diffusion kinetics: (a) XRD patterns of V-doped MnO2 [74]; (b) structure schematic of KMO [37]; (c) XRD patterns of KMO and MnO2 [37]; (d) calculation results of H+ and Zn2+ diffusion energy barriers for KMO and MnO2 electrodes using DFT [37]; (e) GITT image of KMO and MnO2 electrodes [37]; (f) results of ion diffusion coefficients for KMO and MnO2 electrodes [37].
Figure 6. Enlarged interlayer spacing and faster ion diffusion kinetics: (a) XRD patterns of V-doped MnO2 [74]; (b) structure schematic of KMO [37]; (c) XRD patterns of KMO and MnO2 [37]; (d) calculation results of H+ and Zn2+ diffusion energy barriers for KMO and MnO2 electrodes using DFT [37]; (e) GITT image of KMO and MnO2 electrodes [37]; (f) results of ion diffusion coefficients for KMO and MnO2 electrodes [37].
Energies 15 04698 g006
Figure 7. Defect engineering for improved electrical conductivity: (a) XANES and FT spectra of N–MnO2−x and MnO2 [54]; (b) density of states results of MnO2 and N–MnO2−x [54]; (c) schematic illustration of charge density differences and charge transfer behavior in Ti–MnO2 and MnO2 electrodes [60]; (d) GITT profile of Ti–MnO2 and MnO2 electrodes; (e) CV curves of KMO and α-MnO2 [68].
Figure 7. Defect engineering for improved electrical conductivity: (a) XANES and FT spectra of N–MnO2−x and MnO2 [54]; (b) density of states results of MnO2 and N–MnO2−x [54]; (c) schematic illustration of charge density differences and charge transfer behavior in Ti–MnO2 and MnO2 electrodes [60]; (d) GITT profile of Ti–MnO2 and MnO2 electrodes; (e) CV curves of KMO and α-MnO2 [68].
Energies 15 04698 g007
Figure 8. Pillar effect for stabilized cyclability: (a) schematic illustration of δ-NMOH [65]; (b) cycling performance of δ-NMOH [65]; (c) structural illustration of the α-K0.19MnO2 [34]; (d) interpretation of the energy storage mechanism of Co–Mn3O4 [50].
Figure 8. Pillar effect for stabilized cyclability: (a) schematic illustration of δ-NMOH [65]; (b) cycling performance of δ-NMOH [65]; (c) structural illustration of the α-K0.19MnO2 [34]; (d) interpretation of the energy storage mechanism of Co–Mn3O4 [50].
Energies 15 04698 g008
Table 1. Summary of the synthesis strategies and electrochemical performance of various dopants for manganese-based zinc-ion battery cathodes.
Table 1. Summary of the synthesis strategies and electrochemical performance of various dopants for manganese-based zinc-ion battery cathodes.
Cathode MaterialSynthesis MethodCapacity
(mAh·g−1/mA·g−1)
Rate Performance
(mAh·g−1/mA·g−1)
Cycle Performance
(X %/cycles)
Ref.
MnO2–birnessiteOther266/100134/200083.7%/2000[64]
δ-Na0.44Mn2O4·1.5H2OOther278/300106/600098%/10,000[65]
Na:MnO2/GCFElectrodeposition381.8/100175/200075.2%/1000[48]
Na+–δ-MnO2Hydrothermal335/500194/400093%/1000[38]
Na0.1MnO2·0.5H2OOther270/30100/300098%/5000[66]
Na0.44MnO2Other301.3/10080/100069.3%/800[67]
N–Na2Mn3O7Calcination treatment300/200100/10,00078.9%/550[55]
α-K0.19MnO2Hydrothermal266/300113/600090%/400[34]
K1.33Mn8O16Other312/3080/150080%/650[62]
K0.8Mn8O16Hydrothermal300/100154/1000100%/1000[68]
K0.36H0.26MnO2·0.28H2OHydrothermal329.8/30100.1/300093.4%/3000[36]
K0.29MnO2·0.67H2OHydrothermal300/200158/200098%/12,000[37]
Ag1.5Mn8O16Other240/50--[69]
MnO2H0.16(H2O)0.27Hydrothermal275.6/30115.1/300079%/2000[35]
H1.57Mn2O4Ion exchange281/100133.4/100093.5%/1000[45]
Cu–MnOOther320/150156/90070%/1000[61]
Ocu–Mn2O3Other246/50112/100088%/600[70]
NixMn3−xO4Calcination treatment139.7/5098.5/120093.5%/800[51]
Ni-doped Mn2O3Calcination treatment252/100132/100085.6%/2500[52]
Ni0.052K0.119Mn0.948O2·0.208H2OHydrothermal303/30175/120071.4%/2000[40]
Zn2+-δ-MnO2Other358/300124/300080%/2000[57]
Co-Mn3O4Electrodeposition362/10090/400080%/1100[50]
Ca0.28MnO2·0.5H2OHydrothermal298/175124.5/3500100%/5000[39]
Ce-doped MnO2Hydrothermal270/150134/150070%/100[41]
La3+-δ-MnO2Other278.5/100121.8/160071%/200[71]
Bi-α-MnO2Hydrothermal325/300150/500090.9%/2000[42]
Al0.35Mn2.52O4Hydrothermal302/100147/150095%/1000[72]
Al-intercalated MnO2Hydrothermal401/100229/400094.5%/2000[43]
Fe/α-MnO2@PPyOther270/10096/80075%/1000[73]
Ti–MnO2Other259/100179/100080%/4000[60]
V-doped MnO2Other266/6667/1064-[74]
La–Ca co-doped ε-MnO2Other297.3/200161/160076.8%/200[58]
ZnNixCoyMn2−x−yO4@N-rGOOther200.5/1093.5/150079%/900[75]
N–CC@MnO2Other353/500201/600093.6%/1000[76]
N–C@MnOxCalcination treatment305/500100/2000100%/1600[53]
Mn3O4@NCOther280/10097/1000100%/700[63]
ZnMn2O4/NOther221/100110/100097.4%/2500[77]
MnO2–NHCSOther349/100100/200078.7%/2000[78]
N–MnO2–xCalcination treatment285/200155.2/200085.7%/1000[54]
P–MnO2−x@VMGOther302.8/500150.1/10,000Above 90%/1000[79]
S–MnO2Calcination treatment324/200205/200095%/1000[56]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, B.; Chen, J.; Sun, W.; Shao, Y.; Zhang, L.; Zhao, K. Challenges and Perspectives for Doping Strategy for Manganese-Based Zinc-ion Battery Cathode. Energies 2022, 15, 4698. https://doi.org/10.3390/en15134698

AMA Style

Zhang B, Chen J, Sun W, Shao Y, Zhang L, Zhao K. Challenges and Perspectives for Doping Strategy for Manganese-Based Zinc-ion Battery Cathode. Energies. 2022; 15(13):4698. https://doi.org/10.3390/en15134698

Chicago/Turabian Style

Zhang, Bomian, Jinghui Chen, Weiyi Sun, Yubo Shao, Lei Zhang, and Kangning Zhao. 2022. "Challenges and Perspectives for Doping Strategy for Manganese-Based Zinc-ion Battery Cathode" Energies 15, no. 13: 4698. https://doi.org/10.3390/en15134698

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