Next Article in Journal
Data-Driven Detection and Prediction of Refrigeration Equipment Failures Using Rough Sets Theory and the Internet of Things
Previous Article in Journal
Multidimensional Evaluation and Research of Energy Storage Technologies for Nuclear Power Frequency Regulation Scenarios
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Co-Doping Inducing d-Electron Delocalization in α-MnO2 for High-Performance Zinc-Ion Batteries

1
College of Chemistry and Chemical Engineering, North University of China, Taiyuan 030051, China
2
School of Marine Science and Technology, National Industry-Education Platform for Energy Storage, Tianjin University, Tianjin 300072, China
*
Authors to whom correspondence should be addressed.
Processes 2025, 13(11), 3617; https://doi.org/10.3390/pr13113617 (registering DOI)
Submission received: 14 October 2025 / Revised: 4 November 2025 / Accepted: 6 November 2025 / Published: 8 November 2025
(This article belongs to the Special Issue Advanced Technologies for Energy Storage)

Abstract

Element doping technology is widely recognized as an effective strategy for high-performance MnO2-based cathode materials. While this approach improves the electronic and ionic conductivity of MnO2, it is often accompanied by the introduction of oxygen vacancies. This synergistic effect poses challenges for precisely investigating the effect of doping elements on the d-electron configuration of the Mn site and establishing atomic-level structure-activity relationships for high-energy aqueous zinc-MnO2 batteries. In this paper, the rational design of d-electron configurations in the Mn site has been achieved through simple Co doping in α-MnO2 (CMO). Experimental results confirm that the introduction of Co can delocalize the d-electrons of the Mn site and increase the ratio of eg (dz2 and dx2−y2) occupancy. Consequently, the charge transfer resistance, electrode polarization, and Zn2+ diffusion coefficient of the CMO-2 cathode have been greatly optimized. Thus, the as-prepared electrode delivers a high specific capacity of 287.4 mAh g−1 at 1 A g−1, with a capacity retention rate of 92.8% and a corresponding remaining capacity of 199.7 mAh g−1 after 700 cycles. This study paves the road for the designation and construction of high-energy MnO2 cathodes with optimized electronic structures for advanced aqueous zinc ion batteries.

1. Introduction

The urgent need for high-performance and scalable energy storage has triggered a great deal of research into battery technologies beyond lithium-ion systems. Aqueous zinc-ion batteries (AZIBs) have gained considerable attention due to the advantages of non-flammable aqueous electrolytes, minimal environmental impact, and significantly lower material costs [1,2,3,4]. MnO2 is a promising cathode candidate owing to its high theoretical capacity (≈308 mAh g−1), multiple accessible oxidation states, natural abundance, and low cost [5,6,7]. The electrochemical viability of MnO2 for AZIBs applications was first conclusively demonstrated by Kang and colleagues in 2011, which established the foundation for subsequent research efforts in this field [8]. However, the application of MnO2 cathodes is still hindered by the critical limitations of unsatisfactory electronic and ion conductivity. Worse still, the redox processes of MnO2 involve complex phase transformations and dissolution issues in which Mn(IV) reduction produces unstable Mn(III) that disproportionates into Mn(IV) and Mn(II), resulting in permanent structural degradation and material loss [9,10]. Plenty of works have been reported focusing on nanostructuring and composite engineering strategies, while the capacity and cycling stability of MnO2 are still insufficient [11,12]. The key to developing high-performance MnO2-based materials should be emphasized on the rational design of their intrinsic electronic configuration, which is essential for fundamentally improving electronic and ionic transport kinetics, as well as establishing a clear atomic-level understanding of the structure–performance relationship [13,14,15].
The electronic configuration of MnO2 is intrinsically related to the delocalization degree of d-electrons and the occupancy ratio of the eg (dz2 and dx2−y2) and t2g orbitals (dxy, dxz, and dyz) of the Mn site [16,17,18]. Element doping to achieve more delocalized electrons and a high occupancy ratio of the eg orbitals is regarded as an appealing strategy for high electrochemical performance of MnO2 [19,20,21]. It should be noted that the abundant oxygen vacancies and expanded lattice spacing can be widely observed during the doping process, and the enhanced electrochemical performances are always attributed to the synergistic effect between dopants and oxygen vacancies, which is, however, an important obstacle to the precise design of the d-electron configuration of MnO2 and a better understanding of the mechanisms for enhanced electrochemical performance [22,23,24,25]. The research, conducted by Professor Gong Yun’s team [26] from Chongqing University, prepared vanadium-doped manganese dioxide K0.41V0.26MnO2−x·0.5H2O (VMnO). It was found that V doping effectively suppressed the Jahn-Teller effect through orbital hybridization and electron transfer, stabilizing the low-valent Mn. Meanwhile, oxygen vacancies disrupted the local symmetry of the Mn central coordination environment, facilitating electron transfer along the path of Mn-O-V, thereby enhancing electronic conductivity. Recently, Feng’s group [27] proposed a Mg-doped α-MnO2 (MMO) with deficient oxygen vacancy by precisely controlled hydrothermal reaction conditions to reveal the mechanisms of Mg doping in boosting MMO’s electrochemical performance. The electron/ion transfer kinetics for Zn2+ storage had been improved, and the as-prepared electrode can deliver a capacity of 311 mAh g−1 at 0.6 A g−1. Nevertheless, the d-electron configuration of MnO2 remains inadequately characterized, and the work mechanism of the dopants on the specific capacity and the kinetics of MnO2 should be deeply discussed from an electronic perspective.
In this work, a Co-doped α-MnO2 (CMO) with deficient oxygen vacancy is prepared to optimize the d-electron configuration of MnO2 with delocalized d-electrons spin states and higher eg occupancy. The selection of Co2+ (0.65 Å) as a dopant was based on its near-identical ionic radius to both Mn3+ (0.645 Å) and Mn4+ (0.53 Å), ensuring less impact on the crystal structure of α-MnO2. The X-ray photoelectron spectroscopy (XPS), electron paramagnetic resonance (EPR), and density functional theory (DFT) calculations have been conducted to comprehensively investigate the d-electron configurations of Mn sites. The structural characterization and EPR data suggested that CMO-2 featured stable crystal structures and a deficient oxygen vacancy. XPS and theoretical calculations demonstrated that CMO-2 had an optimized eg occupancy ratio of 44.15%, higher than that of MnO2 (40.09%). Band structure analysis indicates that CMO-2 exhibits a narrower bandgap and a significantly elevated density of states (DOS) near the Fermi level. This transformation in electronic structure fundamentally enhances its conductivity, thereby optimizing electron transport efficiency within the electrode. Leveraging these advantages, the CMO-2 electrode achieved a high specific capacity of 287.4 mAh g−1 at a current density of 1 A g−1. Its energy density and power density reached 183.9 Wh kg−1 and 659.7 W kg−1, respectively, with overall performance surpassing the vast majority of reported MnO2-based AZIBs materials. In principle, an effective strategy of engineering the d-electron configurations has been presented in this work to construct a lattice-stable and kinetically optimized cathode towards advanced AZIBs.

2. Results and Discussion

The structure of the prepared CMO is shown in Figure 1a. In a typical preparation process, an appropriate amount of KMnO4 (Shentai Chemical Reagent Co., Ltd., Tianjin, China) and MnSO4 (GFCR Technology Reagent Co., Ltd., Tianjin, China) were added to deionized water to form a uniform solution. By gradually increasing the concentration of Co(NO3)2·6H2O (GFCR Technology Reagent Co., Ltd., Tianjin, China), CMO-1, CMO-2, and CMO-3 were successfully prepared through a one-step hydrothermal method. As shown in Figure 1b,c, CMO and bare MnO2 showed a nanorod structure, and the morphologies are not changed by Co doping. (Figure S1a,c). The samples are further analyzed by a microscope (TEM). As shown in Figure 1d and Figure S1b,d, the surface of the CMO-1, CMO-2, and CMO-3 is smooth, and no obvious defects or cracks are found, which is basically consistent with α-MnO2 (Figure 1f). Figure 1e,g (HRTEM images) demonstrates clear streaks with the crystal lattice fringe spacing of 0.258 and 0.253 nm, corresponding to the (211) crystal face of CMO-2 and α-MnO2, indicating that the nanorods have good crystallinity inside. In addition, the uniform distribution of Co, Mn, and O elements in the sample confirmed the successful preparation of CMO, as shown in Figure S1e.
XRD is further utilized to investigate the crystal structure of CMO and α-MnO2. As shown in Figure 2a, the diffraction peaks of all materials are well matched to those of JCPDS No. 44−0141 MnO2, demonstrating that Co doping does not produce a new crystal structure. Peaks at 12.42°, 17.84°, 28.64°, and 37.2° belong to the (110), (200), (310), and (211) faces of MnO2, respectively. It can be observed that the XRD patterns of CMO-1, CMO-2, and CMO-3 exhibit a slight left shift in peak positions and minor variation in peak sharpness compared to that of α-MnO2, which can be ascribed to the embedding of Co2+ [28]. However, more attention should be focused on the fact that there is no significant peak deviation that can be detected for all these samples, and the characteristic peaks of (110), (200), and (310) facets are still located at similar positions, demonstrating that Co doping negligibly affects the crystal structure of MnO2. To further verify the Co-doped crystal structure, the Raman spectra are carried out for the as-prepared materials, and the results are displayed in Figure 2b. The α-MnO2 featured a typical 2 × 2 tunnel structure consisting of [MnO6] octahedra, and the positions of the characteristic peaks of the four samples remained basically unchanged. The bands centered at ~579 and 633 cm−1 are related to the symmetric Mn-O vibrations, while the peak at ~383 cm−1 can be indexed to the Mn-O bending vibrations [29,30]. In addition, a slight decrease in peak intensities at ~633 cm−1 is detected, which can be attributed to a decreased average Mn oxidation state and a strong interaction between Co and oxygen in the [MnO6] octahedra. Figure 2c illustrates the crystalline structure of CMO, which basically consists of the [MnO6] octahedral of α-MnO2, while partial Mn was substituted by Co. The above results indicate that Co2+ is successfully incorporated into the MnO2, and the tunnel structure consisting of [MnO6] octahedral is not destroyed by Co doping.
The content of dopants also has a significant influence on the electronic configuration of the materials. Previously, efforts have reported that a high ratio of Mn/O in MnO2-based materials would lead to an increased average valence of Mn, suggesting a higher eg occupancy and more delocalized d-electrons [31,32,33]. To accurately determine the element content of these materials, XPS is conducted, and the as-calculated results are displayed in Figure 2d. The Co content of CMO-1, CMO-2, and CMO-3 is 3.81, 4.22, and 4.24%, respectively. Notably, CMO-2 showed the highest Mn/O ratio of 55.82/39.96, while the Mn/O content of CMO-1 and CMO-3 were 54.63/41.56 and 54.39/41.37%, respectively, which is consistent with the results obtained from inductively coupled plasma optical emission spectrometry (ICP-OES), demonstrated in Table S1. Compared with CMO-2, a further increase in the precursor concentration shows a limited impact on the Co content, but reduces the ratio of Mn/O in CMO-3, which can be ascribed to the excessive amount of precursors. All these results indicated that the CMO-2 can be regarded as the optimized candidate for Zn2+ storage, and further detailed characterizations and electrochemical performance will be focused on the sample.
In manganese-based cathode materials, the introduction of oxygen vacancies represents a core strategy for optimizing their electrochemical performance. This process is typically achieved through elemental doping or reduction. Nevertheless, in order to more accurately study the intrinsic effect of the doping element on the electronic structure of α-MnO2, the oxygen vacancies in the material should be minimized as much as possible. Our EPR measurements (Figure 3a) reveal no characteristic signals of oxygen vacancies, demonstrating that bare MnO2 and CMO-2 both featured deficient oxygen vacancy characteristics, indicating that CMO-2 can serve as an excellent candidate material for investigating the influence of Co doping on the d-electron configuration of the Mn site.
To further investigate the electron configurations of the cathodes, XPS spectra of CMO-2 were analyzed. Figure 3b is the Mn 2p spectrum consisting of Mn 2p3/2 at 642.1 eV and Mn 2p1/2 at 653.8 eV, with an energy difference of 11.7 eV. Compared with bare MnO2, the Mn 2p position of CMO-2 slightly shifts to the lower binding energy, which can be attributed to the doping by other metal ions. The Mn 2p3/2 orbital has been fitted into two characteristic peaks at ~642.41 eV and ~644.8 eV, corresponding to Mn3+ and Mn4+, respectively. Meanwhile, the Mn 2p1/2 orbital can be well fitted into peaks located at 653.6 eV (Mn4+) and 655.8 eV (Satellite peak) [16]. XPS analysis indicates that the Mn3+/Mn4+ ratio in CMO-2 is 3.7, higher than the 3.3 observed in bare MnO2. This result confirms that the introduction of cobalt induces a shift in manganese sites from Mn4+ to Mn3+, which suggests that the enhanced delocalization of electrons and increased d-electron cloud density of the Mn site have been successfully achieved in CMO-2. The Co 2p spectrum in Figure S2 presents typical peak positions of Co3+, corresponding to the Co 2p3/2 and Co 2p1/2 signals with a spin-energy separation of 15.1 eV [34]. As for the O 1s spectra, as shown in Figure 3c, there are three distinct peaks at~529.7 eV, ~531.3 eV, and 535.5 eV, attributable to the Mn-O-Mn, Mn-O-H, and H-O-H bonds, respectively [35]. The major O 1s peak for the sample shifts to a lower position, indicating a change in the coordination configuration of Metal-O (Mn, Co), probably due to the formation of the Co-O bonds [36]. By comparing the integrated areas of the peaks of α-MnO2 and CMO-2, the percentage of Mn-O-H is increased from 0.24 to 0.29 after the modification of the Co atom, which can be attributed to the combination of Co with O within the MnO2 lattice [37,38].
The density functional theory (DFT, by VASP5.4.4 (VASP Software GmbH., Vienna, Austria)) calculations are employed to elucidate the influence of Co doping on the d-electron configurations of the Mn site. As revealed in Figure 3d, the charge density difference (CDD) results indicate that the distortion degree in CMO-2 is greater than in the bare MnO2, in which a strengthened response of electrons can be observed. Notably, the interfacial charge state of the Mn atoms connected to the Co atom shifts from a slight charge accumulation to a pronounced charge depletion, indicating that the doping of Co significantly enhances the d-electron delocalization of Mn. Figure 3e presents the contribution of different d orbitals of the Mn atom to DOS (pDOS), listed in Table S2. Notably, the CMO-2 possesses a ratio of eg (dz2 and dx2−y2) occupancy around 44.15%, higher than the 40.09% for α-MnO2, indicating clearly an optimized d-electron configuration and an enlarged d-electron cloud density in CMO-2, which agrees well with the XPS and CDD results. Impressively, CMO-2 presents a decreased bandgap near the Fermi level from ~0.62 eV to ~0.41 eV, attributable to the more delocalized electrons [39]. All the above results demonstrate that the precise control of experimental conditions enables the incorporation of Co without introducing additional oxygen vacancies, while effectively optimizing the d-electron configuration of Mn. CMO-2 possesses appealing structural advantages at both the crystallographic and electronic levels, including negligible oxygen vacancies, higher Mn content, enhanced d-electron delocalization, higher eg orbital occupancy, and reduced bandgap, suggesting a promising potential for electrochemical Zn2+ storage.
The electrochemical performance of α-MnO2, CMO-1, CMO-2, and CMO-3 was investigated as a cathode with zinc foil as an anode and 2 M ZnSO4/0.2 M MnSO4 aqueous solution as electrolyte. The corresponding galvanostatic charge/discharge (GCD) curves at 1 A g−1 between 0.8 and 1.9 V are demonstrated in Figure S3. A comparative analysis of these cathodes reveals that Co doping can significantly enhance the specific capacitance and reduce polarization voltage simultaneously, and CMO-2 demonstrates the highest specific capacity of 217.4 mAh g−1 and the lowest polarization potential of ~173 mV at 1 A g−1, which is consistent with the previous characterizations and discussions about the electronic and crystal structure.
To investigate the zinc storage behavior of CMO-2, cyclic voltammetry (CV) was employed across the 0.8–1.9 voltage range at a scan rate of 0.3 mV s−1. Figure 4a reveals that CMO-2 exhibits redox characteristics similar to those of α-MnO2. The peaks observed near 1.32 V and 1.21 V correspond to the insertion/extraction reactions of H+ and Zn2+, respectively. The redox peaks of the CMO-2 displayed a larger integrated area, enhanced current response, and decreased peak potential separation compared with the α-MnO2 electrode, indicating an enhanced charge storage capacity and improved reaction kinetics and reversibility.
The rate performance of CMO-2 and α-MnO2 has been systematically evaluated. As shown in Figure 4b,c, the specific capacity of CMO-2 is 236, 233, 219, 209 mAh g−1 from 0.3 to 1 A g−1, respectively, while that of bare MnO2 is only 187, 124, 112, 98 mAh g−1. Even at 8 A g−1, CMO-2 can still achieve a capacity of 65 mAh g−1, significantly superior to the 13 mAh g−1 of bare MnO2 (Figure 4d). When the current density returns to 1 A g−1 again, the capacity of α-MnO2 is about 45 mAh g−1, while the capacity of CMO-2 can still reach 200 mAh g−1. The enhancement of H+ intercalation kinetics through cobalt doping enables the CMO-2 cathode to maintain a longer voltage plateau than bare MnO2, even at high current densities, specifically within the voltage range above 1.3 V [27]. The long-cycle stability of the prepared samples is tested at 1 A g−1. As shown in Figure 4e, the CMO-2 can deliver the highest capacity of 287.4 mAh g−1 with 92.8% capacity retention after 700 cycles, whereas the specific capacity of bare MnO2 is only around 54.49 mAh g-1. When the current density is raised to 4 A g−1, the CMO-2 electrode delivered an average capacity of 107 mAh g−1 in Figure S4, and the capacity can be maintained above 100 mAh g−1 after 500 cycles, which is much better than that of α-MnO2 (25/15 mAh g−1). As shown in Figure S5, the CMO-2 cell can deliver an energy density of 183.9 Wh kg−1 and a power density of 659.7 W kg−1, which are superior to most of the previously reported works. The commonly used electrolyte for aqueous zinc-ion batteries is ZnSO4 and MnSO4, and when manganese-based materials are used as the cathode for aqueous zinc-ion batteries, the voltage window is generally within the range of 0.8–2.0 V. As can be seen from the Figure S5, the specific capacity range of these materials at 1A g−1 is 60–180 mAh g−1, which is significantly lower than that of CMO-2. CMO-2 maintains a capacity of around 200 mAh g−1 after 700 cycles at 1A g−1, and the highest specific capacity during cycling can reach 287.4 mAh g−1.
For a better understanding of energy storage mechanisms, ex situ XRD is carried out at 0.2 A g−1 to investigate the crystalline composition cycling from P1 to P8. As shown in Figure 5a–c, the diffraction peaks of zinc hydroxide sulfate (ZSH) exhibit high reversibility during charge–discharge cycles. The signal initially appears at point P1 and intensifies as discharge progresses to P2; during the charging phase, starting from P3, it gradually weakens until it completely disappears at point P5 (1.9 V). Furthermore, the brief single plateau in the initial discharge curve, coupled with the coexistence of MnOOH and zinc manganese oxide (ZMO) diffraction peaks at point P2, collectively reveals the mechanism of synergistic H+ and Zn2+ co-intercalation into the crystal lattice during the first discharge cycle. During the second cycle (P5 to P8), a continuous weakening of the ZMO diffraction peak signal is observable [22]. Meanwhile, the slight peak shift can be observed in the XRD patterns of ZSH and ZMO during cycling. The bulk crystal structure of α-MnO2 remains stable throughout the cycling process, with its diffraction peaks consistently present, while simultaneously confirming that the Zn2+ insertion/extraction reaction is only partially reversible [40,41]. The morphologies of the CMO-2 cathode during the test are displayed in Figure S6. Numerous layered sheets can be clearly observed when discharged to 0.8 V, then disappear after fully charging to 1.9 V, attributing to the reversible forming of ZSH species. The fibrous morphology of the α-MnO2 has been well maintained, however, revealing the optimized structural stability.
A systematic analysis of the relationship between scan rate and peak current has laid the foundation for a deeper understanding of its energy storage kinetics mechanism [38]. As shown in Figure 5d and Figure S7, the cyclic voltametric curves of CMO-2 and α-MnO2 within the range of 0.1 to 1 mV s−1 exhibit excellent peak stability. Although peak shifts occur with increasing scan rates due to enhanced polarization, the fundamental shapes of the redox peaks remain preserved. As shown in Figure 5e and Figure S8, the calculated b-values for CMO-2 (0.60–0.62) are higher than those of α-MnO2 (0.58–0.60), indicating that they have similar kinetic behaviors, while CMO-2 shows an increase in the proportion of capacitive contribution. High capacitive contribution of 87.34% for CMO-2 electrode at 1.0 mV s−1 can be obtained (the inset of Figure 5e), which is superior to the 76.15% for bare MnO2 (Figure S9). The proportion of capacitive contributions progressively increased during the test for both cathodes (Figure 5f and Figure S10); however, the capacitive percentage of CMO-2 was persistently higher than that of α-MnO2, demonstrating a capacitive-dominated charge storage mechanism at high scan rates and thus endowing the CMO-2 with an outperformed electrochemical performance. The electrochemical impedance spectroscopy (EIS) is utilized to comparatively analyze the key impedance parameters of CMO-2 and α-MnO2 cathodes, as displayed in Figure 5g. The CMO-2 cathode shows a smaller diameter of the semicircle in the high-frequency region and possesses a lower charge transfer resistance of 122 than that of MnO2 (181.5), indicating that Co2+ doping has a positive effect on the electrochemical kinetics of the system. Furthermore, the galvanostatic intermittent titration technique (GITT) is carried out to investigate the ion diffusion in the as-designed cathodes [42]. As shown in Figure 5h and Figure S11, CMO-2 displayed a higher ion diffusion coefficient, ranging from 10−13.42 to 10−10.21 cm2 S−1, than that of bare MnO2 (10−14.28 to 10−11.14 cm2 S−1), indicating the H+ and Zn2+ insertion processes were simultaneously boosted by Co doping [43,44,45]. Such improved Zn2+ transport kinetics of CMO-2 may also be related to its decreased adsorption energy of Zn2+, as depicted in Figure 5i.

3. Conclusions

In summary, we have proposed an oxygen-vacancy-free Co-doping strategy, utilizing a one-step hydrothermal method to synthesize bare MnO2 and MnO2 with different cobalt doping ratios. This approach aims to rationally design the d-electron configuration of manganese (Mn) centers. The introduction of Co endows the CMO-2 with better lattice stability, higher eg occupancy, and stronger electron delocalization, which not only guarantees a high capacitive performance but also benefits the improvement of the electron/ion transport kinetics of active materials. All the above merits jointly boosted the energy and power densities of the devices. As anticipated, the composite electrodes exhibit an excellent specific capacity of up to 287.4 mAh g−1 at 1 A g−1, simultaneously with a satisfactory rate and cyclic performance. The synergistic effect of Mn and Co in the CMO-2 provides a high energy density (183.9 Wh kg−1) and power density (659.7 W kg−1). We believe that the strategy presented here displays an important research direction for the high-performance AZIBs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pr13113617/s1, Table S1. The ICP-OES result of the CMO; Table S2. The integral area of Mn atom orbitals and delocalized d-electron proportion in α-MnO2 and CMO-2; Figure S1. SEM and TEM images of (a,b) CMO-1 and (c,d) CMO-3; (e) TEM images of CMO-2 and its corresponding mapping images; Figure S2. Co 2p XPS spectra of the CMO-2; Figure S3. The GCD curves of α-MnO2, CMO-1, CMO-2 and CMO-3 at 1 A g−1; Figure S4. Cycle performances of α-MnO2 and CMO-2 at 4 A g−1; Figure S5. Comparison of capacitive performances of CMO-2 and other Mn-based cathodes [14,46,47,48,49,50,51,52,53,54]. Figure S6. The morphologies of CMO-2 at different discharge and charge states; Figure S7. CV curves of α-MnO2 at different scan rate; Figure S8. Fitted linear plots of log(i) and log(v) for different peaks of α-MnO2; Figure S9. CV curve with a capacitive contribution of α-MnO2 at 1 mV s−1; Figure S10. The capacitive and diffusion contributions of α-MnO2 at various scan rates; Figure S11. Zn2+ diffusion coefficient of α-MnO2 calculated from GITT.

Author Contributions

Formal analysis, X.Z., R.Z. and J.S.; investigation, J.L. (Jiachen Liang); data curation, C.Z. and J.L. (Jinli Lv); writing—original draft, J.L. (Jinli Lv); writing—review and editing, J.L. (Jiachen Liang); supervision, J.L. (Jiachen Liang); project administration, J.S.; funding acquisition, J.L. (Jiachen Liang). All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No. 52102053), the Fundamental Research Program of Shanxi Province (No. 20210302124655).

Data Availability Statement

Data available on request from the authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Dunn, B.; Kamath, H.; Tarascon, J.M. Electrical Energy Storage for the Grid: A Battery of Choices. Science 2011, 334, 928–935. [Google Scholar] [CrossRef]
  2. Li, S.L.; Shang, J.; Li, M.L.; Xu, M.W.; Zeng, F.B.; Yin, H.; Tang, Y.; Han, C.; Cheng, H.-M. Design and Synthesis of a π-Conjugated N-Heteroaromatic Material for Aqueous Zinc-Organic Batteries with Ultrahigh Rate and Extremely Long Life. Adv. Mater. 2023, 35, 2207115. [Google Scholar] [CrossRef]
  3. Li, B.Y.; Ma, Y.T.; Ma, J.B.; Chen, L.K.; Zhao, Y.; Tang, M.C. Challenges and Opportunities Facing Zinc Anodes for Aqueous Zinc-Ion Battery. Energy Mater. Devices 2024, 2, 9370044. [Google Scholar] [CrossRef]
  4. Shi, X.; Xie, J.; Yang, F.; Wang, F.; Zheng, D.; Cao, X.; Yu, Y.; Liu, Q.; Lu, X. Compacting Electric Double Layer Enables Carbon Electrode with Ultrahigh Zn Ion Storage Capability. Angew. Chem.-Int. Ed. 2022, 61, e202214773. [Google Scholar] [CrossRef]
  5. Chuai, M.; Yang, J.; Wang, M.; Yuan, Y.; Liu, Z.; Xu, Y.; Yin, Y.; Sun, J.; Zheng, X.; Chen, N.; et al. High-Performance Zn Battery with Transition Metal Ions Co-Regulated Electrolytic MnO2. eScience 2021, 1, 178–185. [Google Scholar] [CrossRef]
  6. 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, 202002450. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, Y.; Zhang, Y.-Z.; Gao, Y.-Q.; Sheng, G.; Elshof, J.E.T. Defect Engineering of MnO2 Nanosheets by Substitutional Doping for Printable Solid-State Micro-Supercapacitors. Nano Energy 2020, 68, 104306. [Google Scholar] [CrossRef]
  8. Xu, C.; Li, B.; Du, H.; Kang, F. Energetic Zinc Ion Chemistry: The Rechargeable Zinc Ion Battery. Angew. Chem.-Int. Ed. 2012, 51, 933–935. [Google Scholar] [CrossRef]
  9. Xu, Y.; Zhang, G.; Liu, J.; Zhang, J.; Wang, X.; Pu, X.; Wang, J.; Yan, C.; Cao, Y.; Yang, H.; et al. Recent Advances on Challenges and Strategies of Manganese Dioxide Cathodes for Aqueous Zinc-Ion Batteries. Energy Environ. Mater. 2023, 6, 12575. [Google Scholar] [CrossRef]
  10. Ye, X.; Han, D.; Jiang, G.; Cui, C.; Guo, Y.; Wang, Y.; Zhang, Z.; Weng, Z.; Yang, Q.-H. Unraveling the deposition/dissolution chemistry of MnO2 for high-energy aqueous batteries. Energy Environ. Sci. 2023, 16, 1016–1023. [Google Scholar] [CrossRef]
  11. Zhang, N.; Cheng, F.; Liu, Y.; Zhao, Q.; Lei, K.; Chen, C.; Liu, X.; Chen, J. Cation-Deficient Spinel ZnMn2O4 Cathode in Zn(CF3SO3)2 Electrolyte for Rechargeable Aqueous Zn-Ion Battery. J. Am. Chem. Soc. 2016, 138, 12894–12901. [Google Scholar] [CrossRef]
  12. Lin, M.; Shao, F.; Weng, S.; Xiong, S.; Liu, S.; Jiang, S.; Xu, Y.; Jiao, Y.; Chen, J. Boosted charge transfer in oxygen vacancy-rich K+ birnessite MnO2 for water oxidation and zinc-ion batteries. Electrochim. Acta 2021, 378, 138147. [Google Scholar] [CrossRef]
  13. Yao, S.; Wang, S.; Liu, R.; Liu, X.; Fu, Z.; Wang, D.; Hao, H.; Yang, Z.; Yan, Y.-M. Delocalizing the d-electrons spin states of Mn site in MnO2 for anion-intercalation energy storage. Nano Energy 2022, 99, 107391. [Google Scholar] [CrossRef]
  14. Chen, X.; Xu, Z.; Guo, S.; Sun, B.; Li, Q.; Meng, Q.; Wei, F.; Qi, J.; Sui, Y.; Cao, P. Oxygen Vacancy-Rich Cobalt-Doped MnO2 Nanorods for Zn Ion Batteries. ACS Appl. Mater. Interfaces 2025, 17, 12074–12084. [Google Scholar] [CrossRef] [PubMed]
  15. Zhao, W.; Fee, J.; Khanna, H.; March, S.; Nisly, N.; Rubio, S.J.B.; Cui, C.; Li, Z.; Suib, S.L. A two-electron transfer mechanism of the Zn-doped δ-MnO2 cathode toward aqueous Zn-ion batteries with ultrahigh capacity. J. Mater. Chem. A 2022, 10, 6762–6771. [Google Scholar] [CrossRef]
  16. Huang, Y.; Peng, Y.; Ouyang, Q.; Feng, Q.; Wang, H.; Zheng, D.; Wang, F.; Lu, X.; Liu, Q. Boosting Zn2+ Intercalation in Manganese Oxides for Aqueous Zinc Ion Batteries via Delocalizing the d-Electrons Spin States of Mn Site. Energy Storage Mater. 2024, 70, 103476. [Google Scholar] [CrossRef]
  17. Wang, Y.; Fan, Y.; Liao, D.; Wu, Y.; Yu, Y.; Hu, C. Highly Zn2+-Conductive and Robust Modified Montmorillonite Protective Layer of Electrodes Toward High-Performance Rechargeable Zinc-Ion Batteries. Energy Storage Mater. 2022, 51, 212–222. [Google Scholar] [CrossRef]
  18. Zhang, B.; Cheng, G.; Lan, B.; Zheng, X.; Sun, M.; Ye, F.; Yu, L.; Cheng, X. Crystallization Design of MnO2 via Acid Towards Better Oxygen Reduction Activity. Crystengcomm 2016, 18, 6895–6902. [Google Scholar] [CrossRef]
  19. Kong, X.; Ke, J.; Wang, Z.; Liu, Y.; Wang, Y.; Zhou, W.; Yang, Z.; Yan, W.; Geng, Z.; Zeng, J. Co-based Molecular Catalysts for Efficient CO2 Reduction via Regulating Spin States. Appl. Catal. B-Environ. Energy 2021, 290, 120067. [Google Scholar] [CrossRef]
  20. Liu, Y.; Xiao, C.; Huang, P.; Cheng, M.; Xie, Y. Regulating the Charge and Spin Ordering of Two-Dimensional Ultrathin Solids for Electrocatalytic Water Splitting. Chem 2018, 4, 1263–1283. [Google Scholar] [CrossRef]
  21. Wu, Z.-Z.; Gao, F.-Y.; Gao, M.-R. Regulating the Oxidation State of Nanomaterials for Electrocatalytic CO2 reduction. Energy Environ. Sci. 2021, 14, 1121–1139. [Google Scholar] [CrossRef]
  22. Zarshad, N.; Rahman, A.U.; Wu, J.; Ali, A.; Raziq, F.; Han, L.; Wang, P.; Li, G.; Ni, H. Enhanced Energy Density and Wide Potential Window for K Incorporated MnO2@Carbon Cloth Supercapacitor. Chem. Eng. J. 2021, 415, 128967. [Google Scholar] [CrossRef]
  23. Wang, Z.; Han, K.; Wan, Q.; Fang, Y.; Qu, X.; Li, P. Mo-Pre-Intercalated MnO2 Cathode with Highly Stable Layered Structure and Expanded Interlayer Spacing for Aqueous Zn-Ion Batteries. ACS Appl. Mater. Interfaces 2023, 15, 859–869. [Google Scholar] [CrossRef] [PubMed]
  24. Xie, M.; Lin, M.; Feng, C.; Liu, Z.; Xu, Y.; Wang, N.; Zhang, X.; Jiao, Y.; Chen, J. Coupling Zn2+ Doping and Rich Oxygen Vacancies in MnO2 Nanowire Toward Advanced Aqueous Zinc-Ion Batteries. J. Colloid Interface Sci. 2023, 645, 400–409. [Google Scholar] [CrossRef] [PubMed]
  25. Gallaway, J.W.; Yadav, G.G.; Turney, D.E.; Nyce, M.; Huang, J.; Chen-Wiegart Y-C, K.; Williams, G.; Thieme, J.; Okasinski, J.S.; Wei, X.; et al. An Operando Study of the Initial Discharge of Bi and Bi/Cu Modified MnO2. J. Electrochem. Soc. 2018, 165, A2935–A2947. [Google Scholar] [CrossRef]
  26. Alfaruqi, M.H.; Islam, S.; Mathew, V.; Song, J. 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]
  27. Li, Q.; Wang, C.; Zhu, Y.; Du, W.; Liu, W.; Yao, M.; Wang, Y.; Qian, Y.; Feng, S. Unlocking the Critical Role of Mg Doping in α-MnO2 Cathode for Aqueous Zinc Ion Batteries. Chem. Eng. J. 2024, 485, 150077. [Google Scholar] [CrossRef]
  28. Zhong, R.; Xu, M.; Fu, N.; Liu, R.; Zhou, A.; Wang, X.; Yang, Z. A flexible High-Performance Symmetric Quasi-Solid Supercapacitor Based on Ni-doped MnO2 Nano-Array@Carbon Cloth. Electrochim. Acta 2020, 348, 136209. [Google Scholar] [CrossRef]
  29. Shao, W.; Xia, Y.; Luo, X.; Bai, L.; Zhang, J.; Sun, G.; Xie, C.; Zhang, X.; Yan, W.; Xie, Y. Structurally Distorted Wolframite-Type CoxFe1−xWO4 Solid Solution for Eenhanced Oxygen Evolution Reaction. Nano Energy 2018, 50, 717–722. [Google Scholar] [CrossRef]
  30. Zhang, J.; Li, W.; Wang, J.; Pu, X.; Zhang, G.; Wang, S.; Wang, N.; Li, X. Engineering p-Band Center of Oxygen Boosting H+ Intercalation in δ-MnO2 for Aqueous Zinc Ion Batteries. Angew. Chem.-Int. Ed. 2023, 62, e202215654. [Google Scholar] [CrossRef]
  31. Yan, L.; Zhu, C.; Hao, J.; Liang, X.; Bai, Y.; Hu, Q.; Tan, B.; Liu, B.; Zou, X.; Xiang, B. A Universal Voltage Design for Triggering Manganese Dioxide Defects Construction to Significantly Boost the Pseudocapacitance. Adv. Funct. Mater. 2021, 31, 2102693. [Google Scholar] [CrossRef]
  32. 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]
  33. Zhai, T.; Lu, X.; Ling, Y.; Yu, M.; Wang, G.; Liu, T.; Liang, C.; Tong, Y.; Li, Y. A New Benchmark Capacitance for Supercapacitor Anodes by Mixed-Valence Sulfur-Doped V6O13−x. Adv. Mater. 2014, 26, 5869–5875. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, P.-W.; Li, K.; Yu, Y.-X.; Zhang, W.-D. Cobalt-Doped Graphitic Carbon Nitride Photocatalysts with High Activity for Hydrogen Evolution. Appl. Surf. Sci. 2017, 392, 608–615. [Google Scholar] [CrossRef]
  35. Li, Y.; Li, X.; Duan, H.; Xie, S.; Dai, R.; Rong, J.; Kang, F.; Dong, L. Aerogel-Structured MnO2 Cathode Assembled by Defect-Rich Ultrathin Nanosheets for Zinc-Ion Batteries. Chem. Eng. J. 2022, 441, 136008. [Google Scholar] [CrossRef]
  36. Zhao, Y.; Zhang, J.; Wu, W.; Guo, X.; Xiong, P.; Liu, H.; Wang, G. Cobalt-Doped MnO2 Ultrathin Nanosheets with Abundant Oxygen Vacancies Supported on functionalized Carbon Nanofibers for Efficient Oxygen Evolution. Nano Energy 2018, 54, 129–137. [Google Scholar] [CrossRef]
  37. Davoglio, R.A.; Cabello, G.; Marco, J.F.; Biaggio, S.R. Synthesis and Characterization of α-MnO2 Nanoneedles for Electrochemical Supercapacitors. Electrochim. Acta 2018, 261, 428–435. [Google Scholar] [CrossRef]
  38. Miao, J.; Lin, H.; Mao, Z.; He, S.; Xu, M.; Li, Q. Electrochemical Performance of Sn-doped δ-MnO2 Hollow Nanoparticles for Supercapacitors. J. Mater. Sci.-Mater. Electron. 2018, 29, 2689–2697. [Google Scholar] [CrossRef]
  39. Lin, C.; Zhang, H.; Zhang, X.; Liu, Y.; Zhang, Y. Kinetics-Driven MnO2 Nanoflowers Supported by Interconnected Porous Hollow Carbon Spheres for Zinc-Ion Batteries. ACS Appl. Mater. Interfaces 2023, 15, 14388–14398. [Google Scholar] [CrossRef]
  40. Wang, C.; Yang, H.; Wang, B.; Ding, P.; Wan, Y.; Bao, W.; Li, Y.; Ma, S.; Liu, Y.; Lu, Y.; et al. Dual Cation Doping Enabling Simultaneously Boosted Capacity and Rate Capability of MnO2 Cathodes for Zn//MnO2 Batteries. Nano Res. 2023, 16, 9488–9495. [Google Scholar] [CrossRef]
  41. Xia, J.; Zhou, Y.; Zhang, J.; Lu, T.; Gong, W.; Zhang, D.; Wang, X.; Di, J. Triggering High Capacity and Superior Reversibility of Manganese Oxides Cathode via Magnesium Modulation for Zn//MnO2 Batteries. Small 2023, 19, 2301906. [Google Scholar] [CrossRef]
  42. Guo, L.; Zhang, C.; Song, J.F.; Liang, J.C. Building g-C3N4/Fe2O3 Heterojunctions on Carbon Nanotubes for Enhanced Electron Conductivity and Pseudocapacitive Performances. J. Mater. Sci. Mater. Electron. 2023, 34, 2324. [Google Scholar] [CrossRef]
  43. Zhang, Y.; Liu, Y.; Liu, Z.; Wu, X.; Wen, Y.; Chen, H.; Ni, X.; Liu, G.; Huang, J.; Peng, S. MnO2 Cathode Materials with the Improved Stability via Nitrogen Doping for Aqueous Zinc-Ion Batteries. J. Energy Chem. 2022, 64, 23–32. [Google Scholar] [CrossRef]
  44. Gou, L.; Li, J.; Liang, K.; Zhao, S.; Li, D.; Fan, X. Bi-MOF Modulating MnO2 Deposition Enables Ultra-Stable Cathode-Free Aqueous Zinc-Ion Batteries. Small 2023, 19, 2208233. [Google Scholar] [CrossRef] [PubMed]
  45. Luo, S.; Xu, J.; Yuan, B.; Xu, L.; Zheng, R.; Wang, Y.; Zhang, M.; Lu, Y.; Luo, Y. Few-Layer δ-MnO2 Nanosheets Grown on Three-Dimensional N-doped Hierarchically Porous Carbon Networks for Long-Life Aqueous Zinc Ion Batteries. Carbon 2023, 203, 326–336. [Google Scholar] [CrossRef]
  46. Hu, X.; Liao, Y.; Wu, M.; Zheng, W.; Long, M.; Chen, L. Mesoporous copper-doped δ-MnO2 superstructures to enable high-performance aqueous zinc-ion batteries. J. Colloid Interface Sci. 2024, 674, 297–305. [Google Scholar] [CrossRef]
  47. Ni, Z.; Liang, X.; Zhao, L.; Zhao, H.; Ge, B.; Li, W. Tin doping manganese dioxide cathode materials with the improved stability for aqueous zinc-ion batteries. Mater. Chem. Phys. 2022, 287, 126238. [Google Scholar] [CrossRef]
  48. 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]
  49. Xu, P.; Yi, H.; Shi, G.; Xiong, Z.; Hu, Y.; Wang, R.; Zhang, H.; Wang, B. Mg ion pre-intercalated MnO2 nanospheres as high-performance cathode materials for aqueous Zn-ion batteries. Dalton Trans. 2022, 51, 4695–4703. [Google Scholar] [CrossRef]
  50. Liang, X.; Liu, X.; Wang, P.; Guo, Z.; Chen, X.; Yao, J.; Li, J.; Gan, Y.; Lv, L.; Tao, L.; et al. Ion-exchange induced Ni doping of α-MnO2 cathode with structural modification for aqueous zinc ion batteries. J. Power Sources 2025, 635, 236518. [Google Scholar] [CrossRef]
  51. Li, X.; Liu, W.; Sun, H.; Chi, Y.; Gengzang, D.; Zhang, G.; Jiao, H.; Chen, Q.; Wang, P.; Deng, X.; et al. Developing high-performance Zn2+-pre-inserted MnO2 via in situ electrodeposition for aqueous Zn-ion batteries. Dalton Trans. 2025, 54, 11694–11701. [Google Scholar] [CrossRef]
  52. Wang, C.; Liu, H.; Han, L.; Zhang, J.; Liu, R.; Kong, F. Reduced graphene oxide modified ɛ-MnO2 nanoflowers cathode with oxygen vacancy for advanced rechargeable aqueous zinc-ion batteries. Materials Research Bulletin 2025, 191, 113543. [Google Scholar] [CrossRef]
  53. Zhu, Z.X.; Lin, Z.W.; Sun, Z.W.; Zhang, P.X.; Li, C.P.; Dong, R.; Mi, H.W. Deciphering H+/Zn2+ co-intercalation mechanism of MOF-derived 2D MnO/C cathode for long cycle life aqueous zinc-ion batteries. Rare Met. 2022, 41, 3729–3739. [Google Scholar] [CrossRef]
  54. Luo, S.; Xu, J.; Yuan, B.; Chen, L.; Xu, L.; Zheng, R.; Wang, Y.; Zhang, M.; Lu, Y.; Luo, Y. Honeycomb-like δ-MnO2/NCP integrated cathode for advanced aqueous zinc-ion batteries. Electrochim. Acta 2023, 468, 143192. [Google Scholar] [CrossRef]
Figure 1. (a) The synthesis diagram; the SEM images of (b) α-MnO2 and (c) CMO-2; parts (d,e) are TEM and HRTEM images of CMO-2; parts (f,g) are TEM and HRTEM images of α-MnO2.
Figure 1. (a) The synthesis diagram; the SEM images of (b) α-MnO2 and (c) CMO-2; parts (d,e) are TEM and HRTEM images of CMO-2; parts (f,g) are TEM and HRTEM images of α-MnO2.
Processes 13 03617 g001
Figure 2. (a) XRD patterns and (b) Raman spectra of α-MnO2, CMO-1, CMO-2, and CMO-3; (c) Schematic illustration of the crystalline structure for CMO-2; (d) The Co, Mn, and O contents of CMO-1, CMO-2, and MMO-3 calculated from XPS results, respectively.
Figure 2. (a) XRD patterns and (b) Raman spectra of α-MnO2, CMO-1, CMO-2, and CMO-3; (c) Schematic illustration of the crystalline structure for CMO-2; (d) The Co, Mn, and O contents of CMO-1, CMO-2, and MMO-3 calculated from XPS results, respectively.
Processes 13 03617 g002
Figure 3. EPR spectra (a) and XPS (b) Mn2p, (c) O1s of α-MnO2 and CMO-2. (d) Charge density difference in Zn-inserted MnO2 for α-MnO2 and CMO-2 (the yellow and blue colors indicate electron accumulation and depletion, respectively). (e) The pDOS of different Mn atomic orbitals in α-MnO2 and CMO-2 models.
Figure 3. EPR spectra (a) and XPS (b) Mn2p, (c) O1s of α-MnO2 and CMO-2. (d) Charge density difference in Zn-inserted MnO2 for α-MnO2 and CMO-2 (the yellow and blue colors indicate electron accumulation and depletion, respectively). (e) The pDOS of different Mn atomic orbitals in α-MnO2 and CMO-2 models.
Processes 13 03617 g003
Figure 4. (a) Comparison of CV curves of α-MnO2 and CMO-2 at 0.3 mV s−1; GCD curves of (b) α-MnO2 and (c) CMO-2 at different current densities. (d) Rate performance of α-MnO2 and CMO-2 at different current densities. (e) Long cycle performance of α-MnO2 and CMO-2 at 1 A g−1.
Figure 4. (a) Comparison of CV curves of α-MnO2 and CMO-2 at 0.3 mV s−1; GCD curves of (b) α-MnO2 and (c) CMO-2 at different current densities. (d) Rate performance of α-MnO2 and CMO-2 at different current densities. (e) Long cycle performance of α-MnO2 and CMO-2 at 1 A g−1.
Processes 13 03617 g004
Figure 5. (a) GCD curve of CMO-2 at 0.2 A g−1. (b,c) Ex situ XRD patterns of CMO-2; (d) CV curves of CMO-2 at different scan rates. (e) The fitting profiles of log(i) vs. log(v) corresponding to the redox peaks in CMO-2 CV curves. The inset shows the pseudocapacitive contribution for CMO-2 at 1.0 mV/s. (f) Comparative analysis of diffusion-controlled and capacitive contributions to CMO-2 at varying scan rates. (g) EIS curves and the inserted equivalent circuit of α-MnO2 and CMO-2; (h) Zn2+ diffusion coefficient of CMO-2 calculated from GITT. (i) Zn2+ adsorption energy of α-MnO2 and CMO-2.
Figure 5. (a) GCD curve of CMO-2 at 0.2 A g−1. (b,c) Ex situ XRD patterns of CMO-2; (d) CV curves of CMO-2 at different scan rates. (e) The fitting profiles of log(i) vs. log(v) corresponding to the redox peaks in CMO-2 CV curves. The inset shows the pseudocapacitive contribution for CMO-2 at 1.0 mV/s. (f) Comparative analysis of diffusion-controlled and capacitive contributions to CMO-2 at varying scan rates. (g) EIS curves and the inserted equivalent circuit of α-MnO2 and CMO-2; (h) Zn2+ diffusion coefficient of CMO-2 calculated from GITT. (i) Zn2+ adsorption energy of α-MnO2 and CMO-2.
Processes 13 03617 g005
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

Liang, J.; Zhang, C.; Lv, J.; Zheng, X.; Zhou, R.; Song, J. Co-Doping Inducing d-Electron Delocalization in α-MnO2 for High-Performance Zinc-Ion Batteries. Processes 2025, 13, 3617. https://doi.org/10.3390/pr13113617

AMA Style

Liang J, Zhang C, Lv J, Zheng X, Zhou R, Song J. Co-Doping Inducing d-Electron Delocalization in α-MnO2 for High-Performance Zinc-Ion Batteries. Processes. 2025; 13(11):3617. https://doi.org/10.3390/pr13113617

Chicago/Turabian Style

Liang, Jiachen, Chen Zhang, Jinli Lv, Xiaoqing Zheng, Ruisha Zhou, and Jiangfeng Song. 2025. "Co-Doping Inducing d-Electron Delocalization in α-MnO2 for High-Performance Zinc-Ion Batteries" Processes 13, no. 11: 3617. https://doi.org/10.3390/pr13113617

APA Style

Liang, J., Zhang, C., Lv, J., Zheng, X., Zhou, R., & Song, J. (2025). Co-Doping Inducing d-Electron Delocalization in α-MnO2 for High-Performance Zinc-Ion Batteries. Processes, 13(11), 3617. https://doi.org/10.3390/pr13113617

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

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