Surface electron reconfiguration of ceric dioxide artificial interface layer by cationic doping for dendrite-free zinc anode

Linlong Lu , Zheng Wang , Jingwen Cai , Zhengyu Bao , Yukai Lan , Yinze Zuo , Yidong Jiang , Wei Yan , Jiujun Zhang

Front. Energy ›› 2025, Vol. 19 ›› Issue (3) : 382 -394.

PDF (7545KB)
Front. Energy ›› 2025, Vol. 19 ›› Issue (3) : 382 -394. DOI: 10.1007/s11708-025-1002-8
RESEARCH ARTICLE

Surface electron reconfiguration of ceric dioxide artificial interface layer by cationic doping for dendrite-free zinc anode

Author information +
History +
PDF (7545KB)

Abstract

Aqueous zinc metal batteries (ZMBs) are regarded as strong contenders in secondary battery systems due to their high safety and abundant resources. However, the cycling performance of the Zn anode and the overall performance of the cells have often been hindered by the formation of Zn dendrites and the occurrence of parasitic side reactions. In this paper, a surface electron reconfiguration strategy is proposed to optimize the adsorption energy and migration energy of Zn2+ for a better Zn2+ deposition/stripping process by adjusting the electronic structure of ceric dioxide (CeO2) artificial interface layer with copper atoms (Cu) doped. Both experimental results and theoretical calculations demonstrate that the Cu2Ce7Ox interface facilitates rapid transport of Zn2+ due to the optimized electronic structure and appropriate electron density, leading to a highly reversible and stable Zn anode. Consequently, the Cu2Ce7Ox@Zn symmetric cell exhibits an overpotential of only 24 mV after stably cycling for over 1600 h at a current density of 1 mA/cm2 and a capacity of 1 mAh/cm2. Additionally, the cycle life of Cu/Zn asymmetric cells exceeds 2500 h, with an average Coulombic efficiency of 99.9%. This paper provides a novel approach to the artificial interface layer strategy, offering new insights for improving the performance of ZMBs.

Graphical abstract

Keywords

Zn metal batteries / interface layer / solvation structure / electronic structure / Cu2Ce7Ox

Cite this article

Download citation ▾
Linlong Lu, Zheng Wang, Jingwen Cai, Zhengyu Bao, Yukai Lan, Yinze Zuo, Yidong Jiang, Wei Yan, Jiujun Zhang. Surface electron reconfiguration of ceric dioxide artificial interface layer by cationic doping for dendrite-free zinc anode. Front. Energy, 2025, 19(3): 382-394 DOI:10.1007/s11708-025-1002-8

登录浏览全文

4963

注册一个新账户 忘记密码

1 Introduction

In the context of global energy structure transformation, the utilization and storage of renewable energy have become key scientific and technological issues [13]. Among various energy storage technologies, aqueous Zn metal batteries (ZMBs) have garnered significant attention due to their advantages, including safety, environmental friendliness, low cost, and abundant resources [47]. However, further application of ZMBs is hindered by several issues, such as the formation of Zn dendrite [812], side-reactions, and corrosion of Zn anode [1315], which are primarily caused by the sluggish kinetics of Zn2+ desolvation [16,17]. Therefore, optimizing the solvation shell of [Zn(H2O)6]2+ to accelerate the Zn2+ desolvation process is crucial for addressing the challenges faced by ZMBs [1821].

Several strategies have been explored to accelerate the Zn2+ desolvation process: 1) Electrolyte optimization: Certain electrolytes can provide more hydrogen bond donors and acceptors [2224], or have a higher affinity for Zn2+. These electrolytes interact with Zn2+ preferentially and enter its solvation shell, altering the solvation structure environment. For example, DMSO can occupy more positions in the solvation structure of Zn2+ and reduce direct contact between H2O molecules and Zn2+, due to the high number of hydrogen bond donors of dimethyl sulfoxide [25] (DMSO) compared to H2O. However, electrolyte additives may spontaneously precipitate during cell operation [26,27], thereby potentially blocking ion transport channels. 2) Structural design: The design of three-dimensional (3D) anodes [28,29] with abundant pores can provide more contact sites for solvent molecules and Zn2+, facilitating better Zn2+ diffusion. For example, fibrous Zn anodes [30] prepared using techniques such as electrospinning are intertwined with each other to form a 3D network structure. This structure [31] can shorten the diffusion path for Zn2+ and provide additional sites for solvation. However, the uneven distribution of current density in such structures, and the uneven deposition of desolvated Zn2+ may negatively affect the performance of the cell. 3) Surface modification: Interface layers [3234] that strongly interact with Zn2+ can facilitate the shedding of surrounding solvent molecules, reducing the nucleation energy barrier and accelerating the transmission of Zn2+ [35,36]. For example, Cao’s group [37] used a BaTiO3/P(VDF-TrFE) (BTO/PVT) interface layer to reduce the nucleation overpotential and charge transfer resistance (Rct). Zhou’s group [32] proposed the use of in situ-grown polyaniline to improve the redox kinetics of Zn2+ and redox kinetics. A zincophilic interface layer with strong chemical adsorption Zn2+ [3841], can improve the electrochemical performance of Zn anodes. Although enhancing the adsorption capacity for Zn2+ can reduce the nucleation overpotential and guide the uniform deposition of Zn2+ [42], excessive adsorption can hinder the migration of Zn2+ near the interface, the transfer of Zn2+ on the anode surface, negatively impacting the overall performance of the cell. The weak adsorption effect of the non-zincophilic interface layer on Zn2+ is often accompanied by a lower migration energy barrier, which can achieve rapid transfer of Zn2+ and a lower overpotential. However, it also leads to Zn2+ being too easy to move on the surface, hindering the uniform deposition of Zn2+. Therefore, the interlayer is equally important for the adsorption and migration of Zn2+ [4345].

In this work, Cu atoms were successfully doped into Ce atoms to form Cu2Ce7Ox through a simple high-temperature calcination process [46]. Compared to ceric dioxide (CeO2), the electron density near the Ce atoms in Cu2Ce7Ox decreases, and electrons become more dispersed between Cu and O atoms (Fig.1(a)). Moreover, Cu2Ce7Ox generates an unusual s-p-d orbital hybridization, which triggers electron rearrangement. This alters the electrostatic force due to the alteration of the electron cloud distribution, enhancing the interaction between the O atoms in Ce-O and Zn, thereby improving the chemical affinity for Zn2+. Since Cu2Ce7Ox, as the Zn anode interface layer, has appropriate adsorption energy and migration energy, it not only improves the cycling ability of the Zn anode but also reduces the nucleation overpotential during the deposition and dissolution processes.

The Cu2Ce7Ox@Zn/Cu2Ce7Ox@Zn symmetric cell exhibits a low overpotential of 24 mV, with stable cycling for over 1600 h at a current density of 1 mA/cm2 and an electrical capacity of 1 mAh/cm2, which is more than 10 times that of a bare Zn anode. The Cu/Cu2Ce7Ox@Zn/Cu asymmetric cells show a long cycle life of over 2500 h, with an average Coulombic efficiency of 99.9%. Furthermore, when assembled into a Cu2Ce7Ox@Zn/MnO2 full cell, the battery maintains an excellent capacity retention of 88.9% over the first 800 cycles at 1 A/g. This work offers a novel strategy for constructing stable ZMBs, providing a significant step forward in their development for energy storage applications.

2 Results and discussion

The zincophilic interface layer constructed at the electrode/electrolyte interface of a Zn anode plays a significant role in promoting the desolubilization of [Zn(H2O)6]2+ and ensuring the even deposition of Zn2+ ions [47,48]. To explore the mechanism of CeO2 and Cu2Ce7Ox on interfacial solvation, the electron distribution in these materials was analyzed using theoretical calculation. The electron localization function (ELF) was used to analyze the distribution of electrons around atoms. As shown in Fig.1(b)–Fig.1(e), doping Cu atoms to replace a portion of the Ce atoms in CeO2 results in an unusual s-p-d orbital hybridization, which induces electron rearrangement, thereby enhancing the Ce-O bond. This doping of Cu atoms causes the electrons around Ce and O atoms to disperse. The decrease in electron density near O atoms weakens the chemical interaction between O atoms and Zn2+, which reduces the chemical affinity of O atoms for Zn2+. The behavior of Zn2+ at the electrode/electrolyte interface was verified by calculating the adsorption energy, migration energy for Zn2+ and the activation energy.

To further investigate the behavior of Zn2+ at the electrode/electrolyte interface, the adsorption energy, migration energy, and activation energy were calculated. The adsorption energies of CeO2 and Cu2Ce7Ox for Zn2+ were calculated to evaluate the chemical affinity of the interface layer to Zn2+. The adsorption energy of the CeO2 interface layer for Zn2+ was found to be −3.21 eV, which is significantly higher than that of bare Zn (−0.36 eV), suggesting a reduction in the dissolution energy barrier (Fig.1(f)). However, excessive adsorption energy can hinder the subsequent Zn2+ migration and nucleation and may cause deactivation of the CeO2 interface layer, which is harmful to the construction of highly stable ZMBs. In contrast, Cu2Ce7Ox exhibits a moderate adsorption energy of −1.88 eV, which is ideal for reducing the dissolution energy barrier while ensuring the long-term ZMBs operation.

Additionally, the migration energy of Zn2+ was assessed. As shown in Fig.1(g), on the bare Zn surface, the migration energy of Zn2+ is 0.35 eV, which is beneficial to two-dimensional (2D) planar diffusion of Zn2+. Moreover, on the Cu2Ce7Ox interface layer, the migration energy increases to 0.91 eV, which can inhibit excessive 2D planar diffusion, promoting more uniform Zn2+ deposition. However, the migration energy on the CeO2 interface layer is 1.83 eV, likely due to its high adsorption energy, which is not conducive to Zn2+ migration and nucleation. Moreover, the activation energy for the reaction kinetics was calculated by measuring the charge transfer resistance (Rct) to reaction kinetics (Fig. S1). The activation energy of the Cu2Ce7Ox@Zn electrode was found to be 27.85 kJ/mol, which is lower than that of the bare Zn electrode (38.83 kJ/mol), indicating fast reaction kinetics on the Cu2Ce7Ox@Zn surface (Fig.1(h)).

These findings suggest that the Cu2Ce7Ox interface layer optimally balances adsorption and migration energy for Zn2+, improving the performance of Zn metal anodes in aqueous zinc metal batteries (ZMBs) by enhancing both the dissolution and deposition processes.

To confirm the successful doping of Cu atoms into CeO2, a series of characterizations were conducted on the synthesized materials. As shown in Fig.2(a) with X-ray diffraction (XRD analysis), the diffraction peaks at 28.6°, 33.1°, 47.5°, and 56.3° correspond to the (111), (200), (220), and (311) crystal planes of CeO2, respectively, based on the JCPDS-43-1002 reference. When compared to pure CeO2, due to the change of the lattice of CeO2 by Cu doping, the XRD characteristic peak for Cu2Ce7Ox shift toward lower energy levels, showing that Cu doping has altered the lattice structure of CeO2. These XRD results suggest that Cu atoms are incorporated into the crystal structure of CeO2, replacing some Ce atoms.

SEM and high-resolution TEM images show that Cu2Ce7Ox is in a nanosheet-like morphology (Fig.2(b)). Moreover, elemental analysis based on the mass ratio of Cu and Ce indicates the presence of 8.34% Cu and 27.97% Ce, respectively, which approximates a 2:7 ratio. Energy dispersive X-ray spectroscopy (EDS) (Fig. S2) further confirms that Cu, Ce, and O are uniformly distributed throughout Cu2Ce7Ox, providing additional evidence that Cu atoms have been successfully doped into the CeO2 matrix.

In the TEM analysis, the (111) crystal plane of CeO2 was detected and measured, with a lattice spacing of 0.322 nm for Cu2Ce7Ox (Fig.2(c)). This value is larger than that of pure CeO2 (Fig. S3), suggesting that Cu atom doping has expanded the lattice structure of CeO2. According to the Bragg’s equation, the shift in the XRD peaks toward lower angles is consistent with an increase in lattice spacing, which further confirms that the doping of Cu atoms into CeO2 causes a lattice distortion. Notably, the lattice spacing corresponding to Cu atoms was not detected in the TEM, indicating that Cu atoms are fully integrated into the CeO2 substrate.

X-ray photoelectron spectroscopy (XPS) was also employed to examine the chemical state of CeO2 and Cu2Ce7Ox [49]. The Ce 3d spectrum revealed two distinct peaks for Ce3+ and Ce4+ in both CeO2 and Cu2Ce7Ox. In the case of CeO2, the binding energies of Ce3+ and Ce4+ for the Cu2Ce7Ox are higher than those of CeO2. Additionally, the intensity of the Ce4+ peak at 901.4 eV in the Cu2Ce7Ox is higher than in CeO2, while the intensity of the nearby Ce3+ peak is diminished (Fig.2(d) and Fig.2(f)). This observation shows that the doped Cu atoms reduce the electron density around Ce atoms.

The O 1s XPS spectrum of CeO2 and Cu2Ce7Ox shows three distinct oxygen peaks: Oα, Oβ, and Oγ at 528.7, 530.9, and 532.2 eV, corresponding to oxygen atoms in the lattice structure, oxygen vacancy in the materials, and adsorbed oxygen species, respectively. It is worth noting that the proportion of oxygen vacancies (Oβ) in Cu2Ce7Ox has increased compared to CeO2, which further modifies the electronic structure of Cu2Ce7Ox, promoting Cu2Ce7Ox surface electron rearrangement (Fig.2(e) and Fig.2(g)).

The Cu2Ce7Ox material was then uniformly coated on a Zn foil using a simple doctor blade interface layer process, forming a 300 nm thick interface layer (Figs. S5 and S6). SEM images reveal that the nanosheet-like structure of Cu2Ce7Ox is still retained (Fig.2(h)), and the Cu2Ce7Ox interface layer remains smooth and dense, which is beneficial in reducing material shedding due to volume changes on the surface during cycling.

To investigate the interaction between Cu2Ce7Ox and the electrolyte, the contact angle of electrolyte on bare Zn and Cu2Ce7Ox@Zn in a 2 mol/L ZnSO4 solution was measured. The contact angles for bare Zn and Cu2Ce7Ox@Zn were found to be 92.8° and 109.6°, respectively (Fig.2(i)). The higher contact angle of Cu2Ce7Ox@Zn demonstrate better hydrophobicity compared to bare Zn, which can reduce the access of H2O molecules to the electrode/electrolyte interface, thereby improving the overall performance of the Zn anode in zinc metal batteries.

In summary, these characterizations confirm that Cu atoms have been successfully doped into CeO2, and the resulting Cu2Ce7Ox material exhibits significant changes in its crystal structure, electronic properties, and interaction with the electrolyte, all of which contribute to enhanced performance in zinc metal batteries.

The electrode/electrolyte interface plays a crucial role in determining the stability and performance of zinc metal batteries (ZMBs). In aqueous ZMBs, the stability of the Zn anode is even more crucial. Therefore, it is essential to conduct a series of analyses of the electrode/electrolyte interface to understand the impact of the interface on cell performance. One key factor influencing this is the solvation structure of Zn2+ ions, which can exist in several ion pair forms [50]: solvent-separated ion pair (SSIP), contact ion pair (CIP), and aggregate ion pair (AGG). Raman spectroscopy can effectively detect SSIP and CIP, and it can be used to evaluate the interfacial desolvation kinetics of Zn2+ ions based on the relative ratio of these ion pairs (Fig.3(a)).

Raman spectroscopy [51,52] was used to investigate the interfacial solvation structure on bare Zn/electrolyte, CeO2@Zn/electrolyte, and Cu2Ce7Ox@Zn/electrolyte interfaces. The results reveal that the proportion of CIP on Cu2Ce7Ox@Zn is 58.4%, which is higher than that on CeO2@Zn (48.9%) and bare Zn (42.4%). This indicates that more SO42– participate in the solvation structure and bound H2O content is reduced. A higher proportion of CIP can promote the desolvation of [Zn(H2O)6]2+ and reduce the adsorption and decomposition of H2O molecules at the interface, improving the overall performance (Fig.3(b)).

To further explore the solvation structure of the ZnSO4 electrolyte at the electrode/electrolyte interface, Fourier transform infrared spectroscopy (FTIR) was used to analyze the interactions between three kinds of Zn electrode [53,54]. In the FTIR spectra, characteristic peaks at 1025–1210, 1600–1800, and 3200–3700 cm−1 corresponded to the interactions of Zn2+ with SO42–, and the O–H bond in H2O molecules, and hydrogen bonds respectively (Fig.3(c)–Fig.3(e) and S7). Compared to the blank electrolyte, the characteristic peak of SO42– in ZnSO4 electrolyte on the Cu2Ce7Ox surface exhibits a redshift, indicating that SO42– participates more actively in the solvation structure. This increases the proportion of CIP and decreases the desolvation energy barrier. Additionally, the O–H bond of H2O molecules in ZnSO4 electrolyte on the Cu2Ce7Ox surface undergoes a blueshift, indicating that Cu2Ce7Ox weakens the interaction between H2O molecules, thereby increasing the proportion of free H2O. The peak for hydrogen bonding undergoes a redshift, indicating that Cu2Ce7Ox enhances the hydrogen bond network [55] and increases the stability of the electrolyte.

To quantify the influence of artificial interface on the solvation structure, the desolvation energy required for the transition from [Zn(H2O)6]2+ to Zn2+ was calculated (Fig.3(f) and S9). The results show that Cu2Ce7Ox@Zn/electrolyte interface has lower desolvation energy due to the enhanced hydrogen bond network, which reduces the binding energy between H2O molecules and Zn2+ in [Zn(H2O)6]2+, thereby accelerating the redox reaction kinetics.

Molecular dynamics (MD) simulations were employed to further investigate the solvation structure at the electrode/electrolyte interface. On the CeO2@Zn interface, Zn2+ is strongly adsorbed to the interior of CeO2, resulting in the loss of interaction between Zn2+ and SO42– (Fig. S8). MD simulations also provided the proportions on SSIP, CIP, and AGG for the ZnSO4 electrolyte on bare Zn, CeO2@Zn, and Cu2Ce7Ox@Zn interfaces (Fig.3(g)). The proportion of CIP near the Cu2Ce7Ox@Zn interface is the highest, reaching 40%, which is higher than that of CeO2@Zn (38%) and bare Zn (27%). This observation is consistent with the results from MD simulations, FTIR, and Raman spectroscopy.

Moreover, to investigate the stability of these interfaces, MD simulations were also used to calculate the content of interfacial free H2O (Fig.3(h)). The content of free H2O near the bare Zn interface is 84.6%, which is relatively low and could hinder the Zn2+ desolvation process. In contrast, the interfaces of Cu2Ce7Ox@Zn (86.3%) and CeO2@Zn (86.5%) have a higher content of free H2O, indicating that these interfaces enhance the desolvation process of Zn2+.

Finally, the radial distribution function (RDF) and coordination number of Zn2+ with H2O molecules and SO42– at these interfaces were also calculated using MD simulations (Fig.3(i) and Fig.3(j)). The coordination number [21] for Zn-O (SO42–) of ZnSO4 electrolyte at the Cu2Ce7Ox interface is 0.61, which is higher than that at the CeO2 (0.23) and bare Zn interfaces (0.50). In particular, the coordination number of Zn-O (SO42–) of ZnSO4 electrolyte on CeO2 interface is 0.23, which is lower than that on the Cu2Ce7Ox interface and the bare Zn interface due to the excessive adsorption of CeO2 for Zn2+.

In conclusion, this analysis demonstrates that an appropriate adsorption capacity for Zn2+ is crucial for enhancing the desolvation process and improving the redox kinetics of Zn2+, ultimately contributing to the improved performance of ZMBs.

To demonstrate the protective effect of the Cu2Ce7Ox interlayer for Zn anode, as well as its role in inhibiting side reactions and promoting uniform deposition of Zn2+, the immersion experiment of the anode without or with the Cu2Ce7Ox@Zn interface was performed, respectively (Fig.4(a) and Fig.4(b)). After soaking for seven days, the bare Zn surface showed the formation of flake-like Zn4SO4(OH)6·3H2O (ZHS), a product of the corrosion process. Nevertheless, thanks to the protective effect of the Cu2Ce7Ox artificial interface layer, no significant formation of ZHS was observed on the Cu2Ce7Ox@Zn surface (Fig.4(c), Fig.4(d) and S10), indicating that the Cu2Ce7Ox interlayer has an excellent anti-corrosion ability. This protective layer prevents the anode interface from being eroded by H2O molecule, hereby inhibiting the generation of ZHS.

To understand the impact of Cu2Ce7Ox on the Zn deposition behavior, in situ optical microscopic was performed on the electrode without or with Cu2Ce7Ox interface in ZnSO4 electrolyte at a current density of 5 mA/cm2 (Fig.4(e) and Fig.4(f)). After 10 min of deposition, protrusions began to form on the surface of the bare Zn electrode. As time increased to 30 min, these protrusions grow in size, causing the bare Zn surface to become very uneven. These protrusions would eventually continue to grow under the action of an electric field and finally form dendrites. In contrast, after 30 min of deposition under the same current conditions, the Cu2Ce7Ox@Zn surface remained flat with no protrusions, demonstrating the protective effect of the Cu2Ce7Ox interface layer on Zn electrode.

Further tests using a symmetric cell with bare Zn electrode or Cu2Ce7Ox@Zn electrode after several cycles were conducted and analyzed by SEM (Fig. S11). After 20 cycles, the bare Zn surface exhibited visible ZHS, indicating uneven Zn deposition. After 50 cycles, the bare Zn surface began to crack due to volume changes caused by uneven deposition and stripping, as well as dendrite growth. In contrast, no obvious ZHS were observed on the Cu2Ce7Ox@Zn surface, where Zn2+ was evenly deposited on the Cu2Ce7Ox@Zn electrode surface in a sheet-like form after 50 cycles, indicating stable and uniform deposition behavior on the Cu2Ce7Ox@Zn electrode.

In conclusion, the Cu2Ce7Ox interlayer effectively promotes the smooth and uniform deposition of Zn2+ on the anode surface. The schematic diagram of the mechanism (Fig.4(g) and Fig.4(h)) shows that the Cu2Ce7Ox@Zn interlayer inhibits the formation of ZHS and inhibits the hydrogen evolution reaction (HER) [56,57], which enable Zn2+ to be uniformly deposited on the anode surface, inhibiting the appearance of Zn dendrites. This ultimately enhances the stability and lifespan of the Zn anode in zinc metal batteries.

To evaluate the impact of Cu2Ce7Ox on regulating the Zn deposition behavior and overall cell performance, symmetric cells with different electrode configurations, i.e., bare Zn/bare Zn, CeO2@Zn/CeO2@Zn, and Cu2Ce7Ox@Zn/Cu2Ce7Ox@Zn, were assembled and subjected to a series of electrochemical tests.

Nucleation behavior: At an overpotential of −150 mV, chronoamperometry was performed to analyze the respective nucleation modes for nucleation modes (Fig.5(a)). The current density of the bare Zn keeps increasing for up to 300 s, a behavior attributed to preferential nucleation occurring in local regions on the electrode surface. Zn2+ preferentially accumulates in these regions for nucleation, causing the local current density to increase continuously, indicative of an irregular 2D deposition nucleation mode. This irregular nucleation causes Zn2+ to easily move freely on the anode surface, when it reaches the Zn anode surface. Moreover, the Zn2+ will continuously deposit randomly in places with lower nucleation energy, eventually leading to a homogeneous surface of the Zn anode which is prone to dendrite formation. In contrast, the symmetric cells with CeO2@Zn and Cu2Ce7Ox@Zn electrodes showed a different pattern. After 50 s, the current density remains basically stable. This is because CeO2 and Cu2Ce7Ox provide nucleation sites for the electrodes, restricting the migration of Zn2+ on the electrode surface. As a result, Zn2+ can nucleate and grow uniformly under the action of the electric field. They show a three-dimensional (3D) progressive nucleation mode in the subsequent deposition process [58,59]. This nucleation mode can limit the movement of Zn2+ on the CeO2@Zn and Cu2Ce7Ox@Zn electrode surface. Moreover, nucleation can occur simultaneously in multiple directions on the CeO2@Zn and Cu2Ce7Ox@Zn electrode surface, which is conducive to increasing of nucleation sites and improving the deposition efficiency, thereby inhibiting the formation of dendrites. More importantly, because of the strong adsorption energy of CeO2@Zn electrode for Zn2+, it is not conducive to the homogeneous deposition of Zn2+ in the subsequent deposition process on the CeO2@Zn electrode surface.

Corrosion resistance and hydrogen evolution: To further investigate of the kinetic processes of the three kinds of Zn electrodes, Tafel curves [60,61] were obtained to characterize the protective effect of artificial layer on Zn anode (Fig.5(b)), and the corrosion potentials and corrosion currents of the electrodes were measured. The Cu2Ce7Ox@Zn electrode has the highest corrosion potential and the lowest corrosion current, indicating that it has superior corrosion resistance both thermodynamically and kinetically. Furthermore, LSV was used to examine the hydrogen evolution potentials of bare Zn/bare Zn, CeO2@Zn/CeO2@Zn, and Cu2Ce7Ox@Zn/Cu2Ce7Ox@Zn symmetric cells, respectively (Fig.5(c)). Obviously, the hydrogen evolution potential of Cu2Ce7Ox@Zn/Cu2Ce7Ox@Zn symmetric cell is more negative than others. Cu2Ce7Ox@Zn shows outstanding ability in inhibiting the corrosion resistance and hydrogen evolution potential, which is crucial for achieving uniform deposition of Zn2+.

Cyclic stability: To evaluate the long-term stability of the Zn anode, symmetric cells with bare Zn/bare Zn, CeO2@Zn/CeO2@Zn, and Cu2Ce7Ox@Zn/Cu2Ce7Ox@Zn electrodes were cycled for a prolonged period (Fig.5(d)). The bare Zn/bare Zn symmetric cell short-circuited and failed after approximately 90 hours while the CeO2@Zn/CeO2@Zn symmetric cell lasted for 700 h. Although CeO2@Zn electrode can improve cell cycle performance, its overly strong adsorption effect on Zn2+ leads to unfavorable Zn desorption/adsorption during deposition. The bare Zn/bare Zn and CeO2@Zn/CeO2@Zn symmetric cell display a cyclic stability of 147 and 700 h and an overpotential of 53 and 36 mV. In contrast, the Cu2Ce7Ox@Zn/Cu2Ce7Ox@Zn symmetric cell has exceptional stability, cycling for more than 1600 h with an overpotential of only 24 mV. Moreover, the Rct for the Cu2Ce7Ox@Zn/Cu2Ce7Ox@Zn symmetric cell is only 150 Ω at room temperature, which is lower than the Rct of the bare Zn/bare Zn (332 Ω) and CeO2@Zn/CeO2@Zn (202 Ω) symmetric cells (Fig. S12). These results show that Cu atoms doping changes the physicochemical properties of the CeO2 substrate, balancing the adsorption energy and migration energy of Zn2+, enhancing the performance of the ZMBs.

High current performance: Even at a high current density of 5 mA/cm2, the Cu2Ce7Ox@Zn/Cu2Ce7Ox@Zn symmetric cell can still cycle for nearly 1600 h, indicating excellent stability under demanding conditions (Fig. S13). Furthermore, rate performance tests reveals that the overpotential for Cu2Ce7Ox@Zn is the smallest, indicating its excellent interface stability (Fig.5(e)).

Coulombic efficiency: In the long-cycle test using Cu/Zn asymmetric cells at a current density of 2 mA/cm2 and a capacity of 1 mAh/cm2 (Fig. S15), the Cu2Ce7Ox@Zn electrode can significantly optimize the Zn deposition process and improve the cycle life, showing a high average Coulombic efficiency [62] of 99.9%. The potential capacity curves demonstrate that Cu2Ce7Ox@Zn/Cu asymmetric cells have a better reversibility and the lowest nucleation overpotential (Fig. S14).

Impedance analysis: Finally, in situ impedance measurements were conducted on symmetric cells during charge/discharge process, at a current density of 0.5 mA/cm2 and a capacity of 0.5 mAh/cm2 (Fig.5 (f) and 5(g)). Cu2Ce7Ox@Zn/Cu2Ce7Ox@Zn symmetric cell has a smaller Rct during cycling compared to the bare Zn/bare Zn symmetric cell, indicating that the Cu2Ce7Ox@Zn interface remains more stable throughout the cycle and significantly improves the performance of the Zn anode.

In summary, the incorporation of Cu2Ce7Ox into the Zn anode interface significantly improves the electrochemical performance, stability, and lifespan of the ZMB. The Cu2Ce7Ox@Zn electrode provides superior corrosion resistance, enhances Zn deposition uniformity, reduces dendrite formation, and promotes stable cycling performance even under high current densities. These findings demonstrate that Cu2Ce7Ox is a promising material for improving the performance of Zn-based batteries, providing better interface stability and a more efficient Zn deposition process.

The study of full cells [63] with MnO2 cathodes, synthesized with manganese dioxide (MnO2) cathode, aimed to explore the improvement of the overall performance of the cell by constructing the Cu2Ce7Ox interlayer.

Cyclic voltammetry (CV) analysis: The CV curves of the bare Zn/MnO2, CeO2@Zn/MnO2, and Cu2Ce7Ox@Zn/MnO2 full cells in the third cycle, at a scanning rate of 0.5 mV/s (Fig. S16) show that the Cu2Ce7Ox@Zn/MnO2 full cell has faster redox kinetics and lower polarization overpotential, compared to the other two configurations. The Cu2Ce7Ox interlayer helps balance the adsorption/desorption process and migration of Zn2+, thus improving the overall performance of the cell.

Rate performance: To simulate the overall stability of the cells during operation, the rate performance of the three full cells was tested. The Cu2Ce7Ox@Zn/MnO2 full cell operates stably with a capacity of 234.5 mAh/g at a current density of 0.2 A/g. Even at a large current density of 2 A/g, the Cu2Ce7Ox@Zn/MnO2 full cell still maintained a capacity of 138.9 mAh/g (Fig.6(a)). This demonstrates the ability of the Cu2Ce7Ox interlayer to maintain stable performance at high current densities, ensuring long-term stability.

Galvanostatic charge-discharge (GCD) curves: The GCD curves (Fig.6(b) and S17) of the Cu2Ce7Ox@Zn/MnO2 full cell shows the highest capacity with the smallest electrode polarization, which is consistent with the CV results. These results further indicate that the Cu2Ce7Ox@Zn interface effectively reduces polarization and enhances the efficiency of the Zn deposition process.

Self-discharge performance: To assess the self-discharge characteristics of the cells, the CeO2@Zn/MnO2 and Cu2Ce7Ox@Zn/MnO2 were left standing for 24 h to measure the impact of self-discharge on these full cells after one complete cycle (Fig.6(c)). The Cu2Ce7Ox@Zn/MnO2 full cell retained 98.9% of its initial capacity, which is significantly higher than that of the CeO2@Zn/MnO2 full cell (90.4%). This highlights the Cu2Ce7Ox@Zn/MnO2 full cell’s superior ability to maintain capacity during resting periods, likely due to its enhanced interface stability.

Cycle performance: The cycle performance of these full cells was evaluated at current densities of 0.5 and 1 A/g, respectively (Fig.6(d) and Fig.6(e)). The CeO2@Zn/MnO2 and Cu2Ce7Ox@Zn/MnO2 show an upward curve for a certain period, which is the electrochemical activation phenomenon of the cells in the initial stage. The Cu2Ce7Ox/MnO2 full cell exhibits an initial capacity of 190.1 mAh/g at 0.5 A/g, and retains 145 mAh/g after more than 500 cycles. At a higher current density of 1 A/g, the Cu2Ce7Ox/MnO2 full cell displays an initial capacity of 180.9 mAh/g and a remaining specific capacity of 160.9 mAh/g after 800 cycles. These results indicate that the Cu2Ce7Ox artificial interface layer enhances the long-term cycling stability of the Zn-MnO2 full cells by improving Zn2+ migration and deposition, reducing dendrite formation, and maintaining uniform Zn deposition.

Low-temperature performance: To assess the performance of the Cu2Ce7Ox interlayer at low temperatures, the cyclic performance of the full cells with bare Zn/MnO2 and Cu2Ce7Ox@Zn/MnO2 was tested at 0 °C. Impressively, the Cu2Ce7Ox/MnO2 full cell has a maximum capacity of 138.2 mAh/g at a current density of 0.2 A/g, and the reversible capacity is 92.9 mAh/g after 500 cycles, showcasing its outstanding initial capacity even at subzero temperatures (Fig.6(f)), which highlights the great potential of the Cu2Ce7Ox interlayer to improve the low-temperature stability of ZMBs.

In summary, the Cu2Ce7Ox interlayer significantly enhances the performance of Zn/MnO2 full cells by improving the redox kinetics, reducing polarization, maintaining stable Zn2+ migration, and enhancing cycle life. The Cu2Ce7Ox@Zn/MnO2 full cell outperforms the bare Zn/MnO2 and CeO2@Zn/MnO2 cells in terms of rate performance, self-discharge retention, and long-term cycling stability. Furthermore, it shows excellent low-temperature performance, demonstrating its great potential for use in practical applications, such as in high-power, long-cycle life, and low-temperature environments. The incorporation of Cu into the CeO2 matrix effectively balances Zn2+ adsorption and migration, making Cu2Ce7Ox a promising material for improving the overall performance of Zn-based batteries.

3 Conclusions

In conclusion, this study presents a novel strategy of adjusting the d-band center of CeO2 and changing the electronic structure by doping Cu metal elements to balance the kinetics of the adsorption/desorption and migration processes of Zn by CeO2 as interlayers. The findings indicate that Cu2Ce7Ox as the interlayer for Zn anode, provides appropriate adsorption energy and migration energy for Zn2+, which benefits from the regulation of the d-band center.

The Cu2Ce7Ox interlayer significantly improves the stability and lifespan of Zn anodes by reducing the overpotential of the Zn2+ deposition/stripping process. Electrochemical testing shows that the Cu2Ce7Ox@Zn electrode maintains a high cycle life of over 1600 h with excellent reversibility and a high average Coulombic efficiency of 99.9% during the deposition/stripping process. Additionally, the Cu2Ce7Ox@Zn electrode has the smallest initial nucleation overpotential at a current density of 2 mA/cm2 and a capacity of 1 mAh/cm2. The Cu2Ce7Ox@Zn/MnO2 full cell exhibits remarkable cycling performance, with a capacity retention rate of 92.2% after 800 cycles at a current density 1 A/g, further highlighting the benefits of the Cu2Ce7Ox interlayer in enhancing the stability and efficiency of Zn-based batteries.

Overall, this work offers a promising approach to improving Zn anode protection and provides valuable insights for developing advanced strategies to enhance the performance of Zn-MnO2 full cells and other Zn-based batteries.

References

[1]

Gan L, Liu F, Yuan X. . Alumina modified sodium vanadate cathode for aqueous zinc-ion batteries. Frontiers in Energy, 2023, 17(6): 775–781

[2]

Liu S, Yu X, Yan Y. . Dendrite-free lithium deposition enabled by interfacial regulation via dipole-dipole interaction in anode-free lithium metal batteries. Energy Storage Materials, 2023, 62: 102959

[3]

Zhong Y, Xu X, Veder J P. . Self-recovery chemistry and cobalt-catalyzed electrochemical deposition of cathode for boosting performance of aqueous zinc-ion batteries. iScience, 2020, 23(3): 100943

[4]

Liu J, Xu C, Chen Z. . Progress in aqueous rechargeable batteries. Green Energy & Environment, 2018, 3(1): 20–41

[5]

Liu Y, Lu X, Lai F. . Rechargeable aqueous Zn-based energy storage devices. Joule, 2021, 5(11): 2845–2903

[6]

Ruan P, Liang S, Lu B. . Design strategies for high-energy-density aqueous zinc batteries. Angewandte Chemie International Edition, 2022, 61(17): e202200598

[7]

Li Z, Beyene T T, Zhu K. . Realizing fast plating/stripping of high-performance Zn metal anode with a low Zn loading. Journal of Metals, Materials and Minerals, 2024, 34(2): 2009

[8]

Lenhart B, Kathan D, Hiemer V. . Statistical approach to design Zn particle size, shape, and crystallinity for alkaline batteries. Frontiers in Energy, 2024, 18(5): 650–664

[9]

Liu W, Liu X, Ning F. . Fabrication of a heterovalent dual-cation pre-embedded hydrated vanadium oxide cathode for high-performance zinc ion storage. Journal of Materials Chemistry. A, Materials for Energy and Sustainability, 2024, 12(20): 11883–11894

[10]

Wu K, Liu X, Ning F. . Engineering of charge density at the anode/electrolyte interface for long-life Zn anode in aqueous zinc ion battery. ChemSusChem, 2025, 18(1): e202401251

[11]

Yan Y, Shu C, Zheng R. . Modulating sand’s time by ion-transport-enhancement toward dendrite-free lithium metal anode. Nano Research, 2022, 15(4): 3150–3160

[12]

Wang C, Liu S, Xu H. . Adjusting Li+ solvation structures via dipole–dipole interaction to construct inorganic-rich interphase for high-performance Li metal batteries. Small, 2024, 20(24): 2308995

[13]

Zhang X, Hu J P, Fu N. . Comprehensive review on zinc-ion battery anode: Challenges and strategies. InfoMat, 2022, 4(7): e12306

[14]

Yu X, Li Z, Wu X. . Ten concerns of Zn metal anode for rechargeable aqueous zinc batteries. Joule, 2023, 7(6): 1145–1175

[15]

Yang J, Yin B, Sun Y. . Zinc anode for mild aqueous zinc-ion batteries: Challenges, strategies, and perspectives. Nano-Micro Letters, 2022, 14(1): 42

[16]

Xu Y, Guo Z, Song M. . Electrolyte stabilizes Zn2+ reduction reaction process: Solvation, interface and kinetics. Batteries & Supercaps, 2024, 7(11): e202400237

[17]

Yang H, Chen D, Zhao R. . Reunderstanding aqueous Zn electrochemistry from interfacial specific adsorption of solvation structures. Energy & Environmental Science, 2023, 16(7): 2910–2923

[18]

Li B, Liu S, Geng Y. . Achieving stable zinc metal anode via polyaniline interface regulation of Zn ion flux and desolvation. Advanced Functional Materials, 2024, 34(5): 2214033

[19]

Song Y, Huang S, Li C. . Regulating solvation structure and enhancing anion-derived solid electrolyte interphase with N, N′-dimethylpropyleneurea co-solvent for long-term and dendrite-free Zn metal anodes. Energy Storage Materials, 2024, 71: 103629

[20]

Wang M, Ma J, Meng Y. . High-capacity zinc anode with 96 % utilization rate enabled by solvation structure design. Angewandte Chemie International Edition, 2023, 62(3): e202214966

[21]

Zhang H, Ning F, Guo Y. . Unraveling the mechanisms of aqueous zinc ion batteries via first-principles calculations. ACS Energy Letters, 2024, 9(10): 4761–4784

[22]

You C, Wu R, Yuan X. . An inexpensive electrolyte with double-site hydrogen bonding and a regulated Zn2+ solvation structure for aqueous Zn-ion batteries capable of high-rate and ultra-long low-temperature operation. Energy & Environmental Science, 2023, 16(11): 5096–5107

[23]

Meng Q, Bai Q, Zhao R. . Attenuating water activity through impeded proton transfer resulting from hydrogen bond enhancement effect for fast and ultra-stable Zn metal anode. Advanced Energy Materials, 2023, 13(44): 2302828

[24]

Zhan S, Guo Y, Wu K. . Regulating the interfacial charge density by constructing a novel Zn anode-electrolyte interface for highly reversible Zn anode. Chemistry, 2024, 30(3): e202303211

[25]

Feng D, Cao F, Hou L. . Immunizing aqueous Zn batteries against dendrite formation and side reactions at various temperatures via electrolyte additives. Small, 2021, 17(42): 2103195

[26]

Xie C, Li Y, Wang Q. . Issues and solutions toward zinc anode in aqueous zinc-ion batteries: A mini review. Carbon Energy, 2020, 2(4): 540–560

[27]

Shin J, Lee J, Park Y. . Aqueous zinc ion batteries: Focus on zinc metal anodes. Chemical Science, 2020, 11(8): 2028–2044

[28]

Guo N, Huo W, Dong X. . A review on 3D zinc anodes for zinc ion batteries. Small Methods, 2022, 6(9): 2200597

[29]

Li B, Ruan P, Xu X. . Covalent organic framework with 3D ordered channel and multi-functional groups endows Zn anode with superior stability. Nano-Micro Letters, 2024, 16(1): 76

[30]

Cha J S, Park S, Hwang Y. . Stable zinc metal battery development: Using fibrous zirconia for rapid surface conduction of zinc ions with modified water solvation structure. Small, 2025, 21(1): 2406481

[31]

Wang J, Zhang H, Yang L. . In situ implanting 3D carbon network reinforced zinc composite by powder metallurgy for highly reversible Zn-based battery anodes. Angewandte Chemie International Edition, 2024, 63(10): e202318149

[32]

Li B, Liu S, Geng Y. . Achieving stable zinc metal anode via polyaniline interface regulation of Zn-ion flux and desolvation. Advanced Functional Materials, 2024, 34(5): 2214033

[33]

Han W, Tan Y, Ni L. . Sn penetrated zincophilic interface design in porous Zn substrate for high performance Zn-ion battery. Small Methods, 2024, 2401499

[34]

Luan X, Qi L, Zheng Z. . Step by step induced growth of zinc-metal interface on graphdiyne for aqueous zinc-ion batteries. Angewandte Chemie International Edition, 2023, 62(8): e202215968

[35]

Zheng S, Zhao W, Chen J. . 2D materials boost advanced Zn anodes: Principles, advances, and challenges. Nano-Micro Letters, 2023, 15(1): 46

[36]

Li C, Xie X, Liang S. . Issues and future perspective on zinc metal anode for rechargeable aqueous zinc-ion batteries. Energy & Environmental Materials, 2020, 3(2): 146–159

[37]

Zong Q, Lv B, Liu C. . Dendrite-free and highly stable Zn metal anode with BaTiO3/P(VDF-TrFE) coating. ACS Energy Letters, 2023, 8(7): 2886–2896

[38]

Zhao C, Zhang Y, Gao J. . Zincophilic design and the electrode/electrolyte interface for aqueous zinc-ion batteries: A review. Batteries & Supercaps, 2023, 6(5): e202200478

[39]

LeeB, SonM G, SongS A, et al. Insights into the design of zincophilic artificial protective layers enabling uniform nucleation and deposition for stable dendrite-free Zn anodes. Journal of Colloid and Interface Science, 2024, 680(Pt B): 640–650

[40]

Liu J, Ye C, Wu H. . 2D mesoporous zincophilic sieve for high-rate sulfur-based aqueous zinc batteries. Journal of the American Chemical Society, 2023, 145(9): 5384–5392

[41]

Mawintorn T, Lolupiman K, Kiatwisarnkij N. . Fabrication and characterization of zinc anode on nickel conductive cloth for high-performance zinc ion battery applications. Journal of Metals, Materials and Minerals, 2024, 34(3): 2083

[42]

Jin S, Yin J, Gao X. . Production of fast-charge Zn-based aqueous batteries via interfacial adsorption of ion-oligomer complexes. Nature Communications, 2022, 13(1): 2283

[43]

Meng Y, Wang M, Xu J. . Balancing interfacial reactions through regulating p-band centers by an indium tin oxide protective layer for stable Zn metal anodes. Angewandte Chemie International Edition, 2023, 62(40): e202308454

[44]

Zong Q, Li R, Wang J. . Tailoring the whole deposition process from hydrated Zn2+ to Zn0 for stable and reversible Zn anode. Angewandte Chemie International Edition, 2024, 63(41): e202409957

[45]

Wang C, Ouyang T, Wang X. . Accelerating lithium ion transport via increasing the entropy of the electrolyte for stable lithium metal batteries. Journal of Energy Chemistry, 2024, 99: 384–392

[46]

Li Y, Bai X, Yuan D. . Cu-based high-entropy two-dimensional oxide as stable and active photothermal catalyst. Nature Communications, 2023, 14(1): 3171

[47]

Li Y, Yu Z, Huang J. . Constructing solid electrolyte interphase for aqueous zinc batteries. Angewandte Chemie International Edition, 2023, 62(47): e202309957

[48]

Yang J, Zhao R, Wang Y. . Insights on artificial interphases of Zn and electrolyte: Protection mechanisms, constructing techniques, applicability, and prospective. Advanced Functional Materials, 2023, 33(14): 2213510

[49]

Gao F, Tang Y, Liu J. . Nickel foam supported cuce mixed metal oxide as monolith catalyst for no removal. Chemical Engineering Journal, 2023, 474: 145713

[50]

Hou T, Fong K D, Wang J. . The solvation structure, transport properties and reduction behavior of carbonate-based electrolytes of lithium-ion batteries. Chemical Science, 2021, 12(44): 14740–14751

[51]

Wang T, Wang P, Pan L. . Stabling zinc metal anode with polydopamine regulation through dual effects of fast desolvation and ion confinement. Advanced Energy Materials, 2023, 13(5): 2203523

[52]

Yang H, Chang Z, Qiao Y. . Constructing a super-saturated electrolyte front surface for stable rechargeable aqueous zinc batteries. Angewandte Chemie International Edition, 2020, 59(24): 9377–9381

[53]

Yu X, Chen M, Li Z. . Unlocking dynamic solvation chemistry and hydrogen evolution mechanism in aqueous zinc batteries. Journal of the American Chemical Society, 2024, 146(25): 17103–17113

[54]

Dong D, Wang T, Sun Y. . Hydrotropic solubilization of zinc acetates for sustainable aqueous battery electrolytes. Nature Sustainability, 2023, 6(11): 1474–1484

[55]

Wang C, Liu S, Wang X. . Energy level regulation of anions via hydrogen bond effects to construct a stable solid electrolyte interface for a high-stability lithium metal anode. Chemical Communications, 2024, 60(55): 7045–7048

[56]

Liu X, Guo Y, Ning F. . Fundamental understanding of hydrogen evolution reaction on zinc anode surface: A first-principles study. Nano-Micro Letters, 2024, 16(1): 111

[57]

Yang C, Woottapanit P, Geng S. . Highly reversible Zn anode design through oriented ZnO (002) facets. Advanced Materials, 2024, 36(49): 2408908

[58]

Xie W, Zhu K, Jiang W. . Highly 002-oriented dendrite-free anode achieved by enhanced interfacial electrostatic adsorption for aqueous zinc-ion batteries. ACS Nano, 2024, 18(32): 21184–21197

[59]

Mu Y, Li Z, Wu B. . 3D hierarchical graphene matrices enable stable Zn anodes for aqueous Zn batteries. Nature Communications, 2023, 14(1): 4205

[60]

Li C, Zhang X, Qu G. . Highly reversible Zn metal anode securing by functional electrolyte modulation. Advanced Energy Materials, 2024, 14(34): 2400872

[61]

Guo Q, Mo W, Huang J. . Reconfiguring the coordination structure in deep eutectic electrolytes for enabling stable operation of zinc-ion batteries. Nano Letters, 2024, 24(47): 14965–14972

[62]

Wu Z, Wang Y, Zhi C. Zinc-anode reversibility and capacity inflection as an evaluation criterion. Joule, 2024, 8(9): 2442–2448

[63]

Lolupima K, Cao J, Zhang D. . A review on the development of metals-doped vanadium oxides for zinc-ion battery. Journal of Metals, Materials and Minerals, 2024, 34(3): 2084

RIGHTS & PERMISSIONS

Higher Education Press 2025

AI Summary AI Mindmap
PDF (7545KB)

Supplementary files

FEP-25030-OF-LL_suppl_1

3807

Accesses

0

Citation

Detail

Sections
Recommended

AI思维导图

/