Previous Article in Journal
Synergistic Catalysis for Algae Control: Integrating Sonocavitation and Chemical Catalysis
Previous Article in Special Issue
The Effect of Support and Reduction Methods on Catalyst Performance in the Selective Oxidation of 1,2-Propanediol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Boosting the Bifunctional Catalytic Activity of La0.85Y0.15Ni0.7Fe0.3O3 Perovskite Air Electrode with Facile Hybrid Strategy of Metallic Oxide for Rechargeable Zn–Air Batteries

College of Materials Science and Engineering, Nanjing Tech University, Nanjing 211816, China
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(8), 785; https://doi.org/10.3390/catal15080785 (registering DOI)
Submission received: 3 July 2025 / Revised: 12 August 2025 / Accepted: 15 August 2025 / Published: 17 August 2025
(This article belongs to the Special Issue Metal Oxide-Supported Catalysts)

Abstract

Developing cost-effective, sustainable, and high-performance air electrode catalysts for the oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) remains a significant challenge in the advancement of rechargeable zinc–air batteries (ZABs). Herein, we successfully construct a vacancy-rich heterogeneous perovskite La0.85Y0.15Ni0.7Fe0.3O3 (LYNF) hybridized with Co3O4 spinel nanoparticles using a simple chemical bath-assisted method. The Co3O4 composite LYNF material is systematically evaluated as the bifunctional catalyst for ZABs in the proportion of 25 wt%, 50w t%, and 75 wt% (denoted as LYNF-xCo3O4, x = 0.25, 0.5, 0.75). The results confirm an intimate coupling between the perovskite and spinel phases, along with a significant increase in oxygen vacancy concentration. Among the composites, LYNF-0.5Co3O4 exhibits the best performance, achieving an ORR onset potential of 0.813 V vs. RHE at −0.1 mA cm−2 and a lower OER overpotential of 441 mV at 10 mA cm−2. When applied as the air electrode catalyst in ZABs, LYNF-0.5Co3O4 displays the highest discharge voltage and a peak power density of 115 mW cm−2, representing a 20% improvement over pristine LYNF. The enhanced performance of the LYNF-0.5Co3O4 composite is attributed to the accumulation of Co3O4 nanoparticles within the LYNF matrix, which introduces numerous electrochemically active sites and facilitates the charge and mass transport during the catalytic process in ZABs.

Graphical Abstract

1. Introduction

The global energy crisis, coupled with the escalating greenhouse effect, has presented formidable challenges to sustainable development worldwide. Given these critical issues, substantial research efforts have been devoted to advancing clean and renewable energy technologies as viable alternatives to conventional fossil fuels [1]. Among the emerging technologies, metal–air batteries [2,3], water splitting [4,5], and carbon dioxide reduction have garnered substantial attention. Metal–air batteries, in particular, stand out due to their fuel-cell-like configuration, in which atmospheric oxygen acts as the cathodic reactant and metals such as zinc, lithium, aluminum, or magnesium serve as anodes [6]. ZABs, owing to their high theoretical energy density, have emerged as focal points of research interest [7]. Nevertheless, the commercial viability of rechargeable ZABs faces challenges such as poor cycle stability, pronounced charge–discharge polarization, and suboptimal energy efficiency [8]. Mitigating these challenges necessitates enhancing the kinetic efficiency of the oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) at air electrodes. Accordingly, the development of efficient bifunctional oxygen electrocatalysts is vital for improving the overall performance of ZABs.
Perovskites are a class of crystalline materials with the general formula ABO3, whose unique crystal structure endows them with excellent physicochemical properties, thus attracting extensive attention in the fields of catalysis, energy storage, and conversion. Among them, rare-earth transition metal-based perovskites have emerged as a research hotspot in the field of electrocatalysis due to their favorable electrical conductivity, stable structure, and tunable electronic properties [9,10]. Notably, recent studies have elucidated the molecular orbital principles underlying the intrinsic OER and ORR activities in alkaline solutions, highlighting the potential role of LaNiO3 with an eg electron configuration in enhancing both ORR and OER activities Strategies to enhance perovskite catalytic activity include cationic and anionic doping [10], the construction of hybrid or composite structures [11,12], and the adoption of diverse preparation methods [13].
In the field of electrocatalysis, hybridization strategies that integrate multiple functional materials have proven effective in modulating active sites and tailoring electronic structures at the atomic level. Such approaches can yield synergistic effects that surpass the performance of individual components, offering a powerful route to design high-performance electrocatalysts [14,15]. Co3O4, with its typical spinel structure, is widely recognized as a highly suitable catalyst material for both ORR and OER due to its possession of dual valence states [16,17]. Quan et al. [18] demonstrated the synthesis of SrCo0.5Fe0.3Mo0.2O3-δ/Co3O4 via a simple mixture, achieving overpotential as low as 328 mV. The hybridization of Co3O4 with perovskite materials results in biphasic synergies, increasing the number of active sites. Zhang et al. [19] exhibited the fabrication of Co3O4-modified blood powder doped with mesoporous carbon, demonstrating impressive ORR and OER potentials of 0.83 V and 380 mV, respectively. In our previous study, Y and Fe co-doped LaNiO3 perovskite oxide (La0.85Y0.15Ni0.7Fe0.3O3, LYNF) showed a better OER performance with overpotential of 510 mV [20]. However, the effect of coupling LYNF with Co3O4 spinel has not been thoroughly investigated in ZABs.
This research identifies the high polarization arising from sluggish ORR and OER kinetics as the core bottleneck hindering the commercialization of rechargeable ZABs, leading to poor stability and low energy efficiency. To address this fundamental challenge, we developed a high-performance bifunctional catalyst, LYNF-xCo3O4. Utilizing a facile chemical bath-assisted composite strategy [21,22], we successfully constructed a LYNF-xCo3O4 heterostructure featuring abundant interfaces, a high concentration of oxygen vacancies, and strong synergistic effects. This design significantly enhanced intrinsic ORR and OER activity, increased the density of active sites, improved charge transport capability, and ultimately boosted catalyst stability [23,24]. Experimental results demonstrate that the optimal composite, LYNF-0.5Co3O4, achieves remarkably low ORR and OER overpotentials and Tafel slopes, exhibits improved stability, and enhances overall energy efficiency. These findings provide an effective material solution for overcoming the key challenges facing rechargeable ZABs. In this work, LYNF is hybridized with Co3O4 spinel nanoparticles using a simple chemical bath-assisted method. A series of LYNF-xCo3O4 composites (x = 0, 0.25, 0.5, 0.75) are synthesized and evaluated for their ORR/OER performance. Furthermore, their applicability as air electrode catalysts in ZABs is systematically investigated.

2. Results and Discussion

2.1. Material Structure and Analysis

The samples of LYNF, Co3O4, and LYNF-xCo3O4 (x = 0.25, 0.5, 0.75) were analyzed by powder XRD. As shown in Figure 1, the diffraction peaks located at 31.3°, 36.8°, 38.6°, 44.8°, 55.6°, 59.4°, 65.2°, 74.1°, and 77.3° can be well detected as Co3O4 [25,26]. The XRD Rietveld image of LYNF-Co3O4 is shown in Figure S4. The XRD pattern reveals excellent agreement between the observed diffraction peaks (red “obs.”) and the calculated peaks (black “calc.”) derived from the crystal structures of LYNF and Co3O4. The characteristic peak positions for LYNF and Co3O4 (indicated by vertical markers) correspond well to the experimental pattern. The absence of any unidentified diffraction peaks confirms the high phase purity of the sample, with no detectable impurity phases. All perovskite phases exhibit a cubic structure with a space group of Pm-3m. Importantly, in all composite samples, the characteristic diffraction peaks of the perovskite phase match those of LYNF, and the characteristic peaks of the spinel phase align with those of Co3O4. This indicates that the chemical bath-assisted composite process does not alter the composition of the spinel and perovskite phases [27]. The increasing intensity of the Co3O4-related peaks with higher Co3O4 content further suggests a uniform and controllable incorporation of spinel nanoparticles into the LYNF matrix.
To investigate the internal microstructure and element distribution of perovskite catalysts, transmission electron microscopy (TEM) characterization was performed. The TEM images of LYNF-0.5Co3O4 in Figure 2a reveal a heterogeneous architecture, with darker regions assigned to Co3O4 nanoparticles and brighter regions corresponding to the perovskite matrix. High-resolution TEM (HRTEM), as seen in Figure 2b, shows clear lattice fringes with interplanar spacings of 0.245 nm and 0.293 nm, which can be indexed to the (311) plane of Co3O4 and the (110) plane of LYNF, respectively. The intimate contact at the interface suggests strong phase interaction without significant agglomeration or phase separation, which is expected to benefit charge transfer and interfacial electron transport. The SEM-EDS images in Figure 3a–g show the uniform distribution of Co, Fe, Y, La, Ni, and O elements in the LYNF-0.5Co3O4 sample, indicating homogeneity of the hybrid structure, which is beneficial for improving the OER performance. The SEM images in Figure S2 show microstructures of LYNF-Co3O4 composites with different Co3O4 doping levels. LYNF-0.25Co3O4 (a) has notable agglomeration and dense structure. LYNF-0.5Co3O4 (b) exhibits a loose structure, uniform and rich pores, and good particle dispersion, optimal for exposing active sites and mass transfer. LYNF-0.75Co3O4 (c) has enhanced agglomeration and squeezed pores. Thus, LYNF-0.5Co3O4 has the best morphology, benefiting catalytic reactions via structural advantages.
X-ray photoelectron spectroscopy (XPS) was conducted to further probe the electronic structures and surface chemistry of the samples. Figure 4a shows the Ni 2p spectra of LYNF and LYNF-0.5Co3O4. The peaks at 854.3 and 872.1 eV are assigned to Ni3+ 2p3/2 and Ni2+ 2p1/2, respectively, confirming the coexistence of Ni2+ and Ni3+ oxidation states [28,29,30]. A positive shift of 0.2 eV in the Ni 2p3/2 binding energy in the composite indicates an increase in Ni3+ content, suggesting enhanced oxidation states due to interaction with Co3O4. Figure 4b shows that the peaks at 710.1 and 723.8 eV correspond to Fe 2p3/2 and Fe 2p1/2, respectively [31,32]. Compared to LYNF, the Fe 2p3/2 peaks of LYNF-0.5Co3O4 shift from 710.4 eV to 711.1 eV.
In the O 1s spectra in Figure 4c, four distinct oxygen species are deconvoluted. The peak at 528.5 eV (O1) is attributed to lattice oxygen (O2−), while the peaks at 530.4 eV (O2), 531.7 eV (O3), and 533.0 eV (O4) are assigned to oxygen vacancies, surface-adsorbed oxygen species (e.g., −OH or O2−), and weakly bound surface species (such as carbonates or adsorbed H2O), respectively [33,34]. Notably, the proportion of the O2 component in LYNF-0.5Co3O4 reaches 31.5%, which is significantly higher than that in pristine LYNF (16.8%). This substantial increase in oxygen vacancy content suggests that the introduction of Co3O4 effectively promotes the generation of surface defects. XPS shows a 0.2 eV positive shift in Ni 2p3/2 binding energy for LYNF-0.5Co3O4, indicating increased Ni3+ content-critical active sites for ORR and OER, enhancing adsorption and activation of oxygen intermediates. Higher Fe oxidation states strengthen affinity for oxygen-containing intermediates, accelerating rate-determining steps in both reactions (e.g., * OOH formation in OER, O = O cleavage in ORR). Oxygen vacancies, significantly increased with Co3O4, act as O2 adsorption sites, reducing dissociation energy for ORR and facilitating lattice oxygen participation in OER via redox mechanisms [35,36].
Such structures not only facilitate the adsorption and activation of oxygen intermediates but also enhance the electronic conductivity of the composite, thereby contributing to improved electrocatalytic performance. The synergistic effect between the spinel Co3O4 and the perovskite LYNF is essential for optimizing the surface electronic environment and boosting the overall catalytic activity.

2.2. Electrochemical Performance

The electrochemical performance of these catalysts was assessed via LSV polarization curves in an O2-saturated 0.1 M KOH electrolyte using a three-electrode system. As shown in Figure 5a, LYNF-0.5Co3O4 exhibits the highest limiting current density and the most positive onset potential (0.813 V) among all samples, outperforming LYNF-0.75Co3O4 (0.789 V), LYNF-0.25Co3O4 (0.775 V), and pristine LYNF (0.751 V). Moreover, its half-wave potential (0.630 V) surpasses that of LYNF-0.75Co3O4 (0.612 V), LYNF-0.25Co3O4 (0.608 V), and pristine LYNF (0.584 V). The initial potential of LYNF-0.5Co3O4 is also higher (0.813 V) than that of LYNF-0.25Co3O4 (0.775 V) and LYNF (0.751 V), indicating a superior ORR performance. Figure S7 is for OER and ORR performance comparison chart. These enhancements can be attributed to the synergistic coupling between the spinel and perovskite phases, which improves electronic conductivity and facilitates oxygen adsorption or activation.
The corresponding Tafel plots in Figure 5b further demonstrate the improved ORR kinetics of LYNF-0.5Co3O4, which exhibits the lowest Tafel slope of 91.4 mV dec−1. In comparison, the slopes of LYNF-0.75 Co3O4, LYNF-0.25 Co3O4, and pristine LYNF are 92.2, 114, and 141 mV dec−1, respectively. A smaller Tafel slope reflects more favorable electron transfer kinetics during the reaction.
As depicted in Figure 5c, the overpotential of LYNF-0.5Co3O4 at the current density of 10 mA cm−2 is 441 mV, which is lower than that of LYNF-0.25Co3O4 (455 mV), LYNF-0.75Co3O4 (454 mV), and LYNF (510 mV). The OER kinetics were evaluated by calculating the Tafel slope, as shown in Figure 5d. LYNF-0.5Co3O4 exhibits a Tafel slope of 57 mV dec−1, which is lower than those of LYNF-0.75Co3O4 (64.1 mV dec−1), LYNF-0.25Co3O4 (64.5 mV dec−1), and LYNF (96.9 mV dec−1). Figure S8 is overpotentials to afford 10 mA cm−2 for different electrocatalysts [35,36]. It indicates that LYNF-0.5Co3O4 has the lowest overpotential. These results indicate that the integration of LYNF with Co3O4 enhances the OER kinetics. This improvement can be attributed to the heterostructures formed upon recombination, which typically exhibit enhanced electron and ion transport pathways. These features reduce charge transfer resistance and thereby promote superior OER performance. The Koutecky–Levich (K-L) plots (Figure S6) of LYNF and LYNF-0.5Co3O4 derived from the RRDE measurements showed excellent linearity between J−1 and ω−1/2, indicating first-order kinetics of ORR for the two materials. Based on the K-L equation, the electron-transfer number (n) of LYNF-0.5Co3O4 was calculated to be about 3.95, confirming its favorable four-electron reaction path toward catalyzing ORR.
For the electrochemical impedance spectroscopy (EIS) measurements, the Nyquist plots were analyzed using ZView software (Version 3.5c) to fit the equivalent circuit. The equivalent circuit employed was a typical Rs (RctCPE) model, where Rs represents the series resistance, Rct is the charge transfer resistance, and CPE denotes the constant phase element. The Nyquist plots for LYNF and LYNF-xCo3O4 (x = 0, 0.25, 0.5, 0.75) samples are shown in Figure 5e. All plots exhibit typical semicircular profiles in the high-to-mid frequency region, which correspond to the charge transfer resistance (Rct) at the electrode/electrolyte interface [37,38]. Among the tested samples, LYNF-0.5Co3O4 exhibits the smallest semicircle diameter, indicating the lowest Rct value. In contrast, pristine LYNF shows the largest Rct, suggesting sluggish electron transfer during the OER process. The trend in Rct values (LYNF-0.5Co3O4 < LYNF-0.25Co3O4 ≈ LYNF-0.75Co3O4 < LYNF) is consistent with the observed electrocatalytic activity trends from LSV and Tafel analyses. The reduced charge transfer resistance of LYNF-0.5Co3O4 can be attributed to the synergistic effect between LYNF and Co3O4, which improves the electrical conductivity, enhances electron mobility, and facilitates faster charge transfer at the catalyst surface.
Figure 6a–d presents the CV curves of the various samples recorded within a non-Faradaic region. LYNF-0.5Co3O4 exhibits the largest Cdl value of 2.54 mF cm−2, which is significantly higher than that of pristine LYNF. The LYNF-based composites show slightly lower Cdl values: 1.91 mF cm−2 for LYNF-0.25Co3O4, 1.83 mF cm−2 for LYNF-0.75Co3O4, and 1.59 mF cm−2 for pristine LYNF. A higher Cdl indicates a larger electrochemically active surface area, which provides more accessible active sites for catalytic reactions. As shown in Figure S5, the BET surface areas of LYNF and LYNF-0.5Co3O4 are calculated to be 2.8513 m2 g−1 and 7.8153 m2 g−1, respectively. This indicates that the introduction of Co3O4 significantly increases the BET surface area of the catalyst. To link this with electrochemical performance, the electrochemical surface area (derived from Cdl) of LYNF-0.5Co3O4 is larger than that of LYNF, which is consistent with the trend of BET surface area. This consistency supports that the increased BET surface area provides more active sites, contributing to the enhanced electrochemical performance.
The long-term electrochemical stability of the catalysts is also investigated by accelerated aging tests. LYNF-0.5Co3O4 displays excellent stability over 2000 CV cycles. As depicted in Figure 7a–d, it is observed that the potential for ORR and OER LYNF-0.5Co3O4 decreases by 1 mV and 40 mV at 10 mA cm−2, respectively. Meanwhile, the values for LYNF are 8 mV and 140 mV. These results further demonstrate that LYNF-0.5Co3O4 exhibits promising catalytic stability in alkaline electrolyte, which is essential for practical rechargeable ZAB applications.

2.3. Zn–Air Battery Performance

To further evaluate the practical applicability of LYNF-0.5Co3O4, aqueous ZABs were assembled using it as the air cathode catalyst. The battery configuration is schematically illustrated in Figure 8a. The electrochemical performance of the ZABs was evaluated through galvanodynamic discharge and cycling tests. As shown in Figure 8b, the LYNF-0.5Co3O4-based ZAB achieved a peak power density of 115 mW cm−2, which is 20% higher than that of the pristine LYNF-based ZAB (93 mW cm−2). The enhanced output reflects the superior ORR and OER bifunctional catalytic activity of the composite catalyst, which facilitates rapid oxygen-related ORR processes and effective charge transfer.
In addition to power output, long-term cycling stability was examined under a constant current density of 5 mA cm−2 in Figure 8c. The LYNF-0.5Co3O4-based ZAB exhibits excellent cycle stability, with a relatively small voltage gap between charge and discharge even after prolonged operation. In contrast, the battery using pristine LYNF shows a pronounced increase in overpotential during cycling, indicating more severe degradation. These results demonstrate that hybridizing LYNF with Co3O4 not only enhances catalytic activity but also substantially improves the stability of the air electrode, making LYNF-0.5Co3O4 a highly promising candidate for practical rechargeable ZAB applications.

3. Experimental Section

3.1. Synthesis of Samples

LYNF was synthesized utilizing the conventional self-propagating combustion method according to our previous study. Initially, La(NO3)3·6H2O, Y(NO3)3·6H2O, Ni(NO3)2·6H2O, and Fe(NO3)3·9H2O in stoichiometric amounts were dissolved in 50 mL of distilled water. Subsequently, ethylenediaminetetraacetic acid (EDTA) and citric acid (CA) were incorporated into the aqueous solution, followed by pH adjustment to 8 using ammonia. The solution was then stirred at 80 °C until gelation occurred. Upon gel formation, the precursor gel was transferred into an electric furnace for spontaneous combustion. The resultant product was subjected to calcination at 800 °C for 5 h. Co3O4 was synthesized via a high-temperature calcination method by heating Co(NO3)2·6H2O at 400 °C for 3 h. Subsequently, Co3O4 and LYNF were mixed at mass ratios of 75 wt%, 50 wt%, and 25 wt%, respectively, and dispersed in 60 mL of deionized water. The mixtures were stirred and heated in a water bath at 90 °C. Afterward, the resulting suspensions were subjected to calcination at 400 °C for 3 h. The obtained products were denoted as LYNF-0.75Co3O4, LYNF-0.5Co3O4, and LYNF-0.25Co3O4, respectively. The synthesis procedure for LYNF-0.5Co3O4 electrocatalysts is schematically illustrated in the Supplementary Materials (Figure S1).

3.2. Characterization

X-ray diffraction (XRD) patterns were acquired utilizing a powder X-ray diffractometer, which employed a Cu Kβ source for irradiation, operating at a scan rate of 10° per minute. High-resolution transmission electron microscopy (TEM) images were acquired via a JEOL-Model-JEM-2100F field-emission electron microscope at an accelerating voltage of 200 kV. To investigate the surface chemistry and ascertain the atomic composition, X-ray photoelectron spectroscopy (XPS) was employed.

3.3. Electrochemical Test

The electrochemical characterization was conducted using a three-electrode system, incorporating a Pt wire as the counter electrode, an Ag/AgCl electrode (3.5 M KCl solution) as the reference electrode, and a 0.1 M KOH solution serving as the electrolyte. Initially, a mixture comprising 5 mg of catalyst, 5 mg of Super P Li conductive carbon, 50 µL of Nafion solution (5 wt%), 300 µL of isopropanol, and 700 µL of water was prepared. The suspension was homogenized using ultrasound treatment for 1 h. The glassy carbon electrode was subsequently polished with 0.4 mm and 30 nm Al2O3 particles in succession, followed by thorough rinsing with deionized water and ethanol. A precise volume of 5 µL from the homogeneous catalyst suspension was then applied onto a glassy carbon disc electrode, possessing a diameter of 4 mm and a surface area of 0.1256 cm2, using a pipette. The ORR/OER polarization curves were recorded via linear sweep voltammetry (LSV) at a sweep rate of 5 mV s−1, with potential windows set to 0 V to 1 V (vs. Ag/AgCl) for OER, respectively. The electrochemical impedance spectroscopy (EIS) potentials were tested at a potential of 0.76 V vs. Ag/AgCl, with frequencies ranging from 100,000 to 0.1 Hz and amplitudes of 5 mV. The capacitive currents were determined using cyclic voltammetry (CV) across varying scan rates of 20, 40, 60, 80, 100, and 120 mV/s to estimate the electric double-layer capacitance (Cdl).
Notably, a lower overpotential signifies superior OER catalytic efficiency and an enhanced rate of mass transfer. The overpotential (η) was derived using the following equation [39]:
η = E j = 10 mA / c m 2 1.23   V   vs .   RHE
where 1.23 V represents the equilibrium potential of O2/H2O.
Within the non-Faradaic potential regime, the double-layer capacitance (Cdl) was assessed via CV, facilitating the estimation of the electrochemical surface area (ECSA). ECSA serves as an indicator of the catalyst’s effective specific surface area pertinent to the oxygen evolution reaction (OER) and is derived using the following formula [40]:
ECSA = C dl C s
where Cs represents the specific capacitance of the electrode material in a unit area. Given that ECSA is directly proportional to Cdl under identical conditions, Cdl indirectly reflects the extent of the catalyst’s surface area involved in OER activities.

4. Challenges and Limitations

4.1. Contributions of LYNF-xCo3O4

The LYNF-0.5Co3O4 catalyst achieves a significant reduction in overpotential for both ORR and OER through optimized electronic structure and enhanced interfacial contacts. Specifically, it exhibits an elevated oxygen vacancy concentration of 31.5%, an ORR onset potential of 0.813 V, and an OER overpotential of 441 mV at 10 mA cm−2, accompanied by reduced Tafel slopes for both reactions. These features directly enhance charge transfer kinetics, thereby contributing to mitigated charge–discharge polarization and improved energy efficiency. Accelerated aging tests further demonstrate the catalyst’s favorable stability: after 2000 cyclic voltammetry (CV) cycles, the decay in ORR and OER potentials is merely 1 mV and 40 mV, respectively. Moreover, the assembled ZABs with this catalyst shows only a slight increase in the charge–discharge voltage gap during cycling at 5 mA cm−2, indicating a distinct contribution to short-term cycling stability.

4.2. Limitations of LYNF-xCo3O4

With respect to long-term stability, despite the demonstrated stability in 2000 cyclic voltammetry (CV) cycles and approximately 100 h of battery cycling tests, there remains a gap compared to the requirement of thousands of cycles in practical applications. Additionally, the mechanisms underlying active site loss, oxygen vacancy reconstruction, or interactions with electrolytes (e.g., carbonation) during long-term cycling of the catalyst have not been explored. In terms of adaptability to practical operating conditions, the performance of the catalyst was only evaluated in 0.1 M KOH electrolyte, whereas practical ZABs typically utilize higher-concentration electrolytes (e.g., 6 M KOH) and may encounter issues such as zinc dendrite growth and electrolyte drying [41,42]. The stability and catalytic activity of the catalyst under these complex conditions remain unvalidated. Regarding challenges in cost and large-scale preparation, although the chemical bath-assisted method is described as “facile”, the introduction of Co3O4 may increase material costs, and the feasibility of this method in large-scale production (e.g., uniformity control, yield) has not been discussed, which limits the evaluation of its commercialization potential. Concerning limitations in mechanistic research, although oxygen vacancies and interfacial electron coupling are hypothesized to be key to performance enhancement, there is a lack of in situ characterization (e.g., in situ XPS, in situ Raman) to directly confirm the dynamic role of these factors in catalytic reactions, and the precise identification of active sites remains unclear. These limitations will be the focus of our future research to further enhance the practical application value of the catalyst [43].

5. Conclusions

In conclusion, this study presents a simple hybrid strategy to further optimize the Y and Fe co-doped perovskite material. Based on LYNF, hybrid samples with different composite ratios were obtained by adjusting only the stoichiometric coefficient of the Co element without changing the composition of the matrix perovskite. Among them, LYNF-0.5Co3O4 with a fine heterostructure has a better activity than the original LYNF. It proved that there are a large number of charge transfers and strong electronic coupling in LYNF-0.5Co3O4. Meanwhile, the dispersed and compact interface between Co3O4 and perovskite generates a synergistic effect, which can provide more active sites and effectively adapt to local environmental changes during the ORR and OER process. This study indicates that the one-step mixing method is an effective approach to preparing high-performance catalysts, providing a new direction for the search of bifunctional electrocatalysts with practical application value.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/catal15080785/s1. References [44,45,46,47] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, G.Z. and J.B.; formal analysis, J.Y.; writing—original draft preparation, X.Y.; writing—review and editing, X.Y., G.Z., J.B., and J.Y.; project administration, Y.Z.; funding acquisition, Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Priority Academic Program Development (PAPD) of Jiangsu Higher Education Institutions and Top-notch Academic Programs Project of Jiangsu Higher Education Institutions (TAPP).

Informed Consent Statement

All authors have read and agreed to the published version of the manuscript.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

This work was supported by the Priority Academic Program Development (PAPD) of Jiangsu Higher Education Institutions and Top-notch Academic Programs Project of Jiangsu Higher Education Institutions (TAPP). We acknowledge the Jushang Scientific Research Service Platform (www.jsceshi.cn accessed on 4 July 2024.) for providing SEM measurement.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bhoyate, S.D.; Kim, J.; de Souza, F.M.; Lin, J.; Lee, E.; Kumar, A.; Gupta, R.K. Science and engineering for non-noble-metal-based electrocatalysts to boost their ORR performance: A critical review. Coord. Chem. Rev. 2023, 474, 214854. [Google Scholar] [CrossRef]
  2. Wang, J.; Gao, S.; Hu, S. Zinc-air battery-H2O2 generation system: Current progress, key challenges, optimization strategies and future developments. Chin. Chem. Lett. 2025, 111000, in press. [Google Scholar] [CrossRef]
  3. Zhu, Z.; Zhu, J.; Chen, Y.; Liu, X.; Zhang, M.; Yang, M.; Liu, M.; Wu, J.; Li, S.; Huo, F. Leather waste as precursor to prepare bifunctional catalyst for alkaline and neutral zinc-air batteries. Chin. Chem. Lett. 2023, 34, 107756. [Google Scholar] [CrossRef]
  4. Cui, X.; Zhao, Q.; Huang, Z.; Xiao, Y.; Wan, Y.; Li, S.; Lee, C.-S. Water-Splitting Based and Related Therapeutic Effects: Evolving Concepts, Progress, and Perspectives. Small 2020, 16, 2004551. [Google Scholar] [CrossRef] [PubMed]
  5. Tang, J.; Liu, T.; Miao, S.; Cho, Y. Emerging Energy Harvesting Technology for Electro/Photo-Catalytic Water Splitting Application. Catalysts 2021, 11, 142. [Google Scholar] [CrossRef]
  6. Akram, S.; Hussain, S.; Arif, M.; Ali, M.H.; Tariq, M.; Rauf, A.; Munawar, K.S.; Alkahtani, H.M.; Alhaj Zen, A.; Ali Shah, S.A. Psidium guajava mediated green synthesized cobalt oxide nanoparticles dispersed on reduced graphene oxide for electrocatalytic water splitting. RSC Adv. 2025, 15, 13786–13798. [Google Scholar] [CrossRef]
  7. Chen, R.; Hung, S.-F.; Zhou, D.; Gao, J.; Yang, C.; Tao, H.; Yang, H.B.; Zhang, L.; Zhang, L.; Xiong, Q.; et al. Layered Structure Causes Bulk NiFe Layered Double Hydroxide Unstable in Alkaline Oxygen Evolution Reaction. Adv. Mater. 2019, 31, e1903909. [Google Scholar] [CrossRef]
  8. Tan, D.; Xiong, H.; Zhang, T.; Fan, X.; Wang, J.; Xu, F. Recent progress in noble-metal-free electrocatalysts for alkaline oxygen evolution reaction. Front. Chem. 2022, 10, 1071274. [Google Scholar] [CrossRef]
  9. Hou, A.; Ouyang, R. Origin of the Different Trends of Experimental Activity on Perovskite Catalysts between OER and ORR. ACS Appl. Energy Mater. 2025, 8, 3481–3490. [Google Scholar] [CrossRef]
  10. Ren, X.; Zhai, Y.; Yang, N.; Wang, B.; Liu, S. Lattice Oxygen Redox Dynamics in Zeolite-Encapsulated CsPbBr3 Perovskite OER Electrocatalysts. Adv. Sci. 2025, 12, e2412679. [Google Scholar] [CrossRef]
  11. Du, X.; Ai, H.; Chen, M.; Liu, D.; Chen, S.; Wang, X.; Lo, K.H.; Pan, H. PLD-fabricated perovskite oxide nanofilm as efficient electrocatalyst with highly enhanced water oxidation performance. Appl. Catal. B Environ. 2020, 272, 119046. [Google Scholar] [CrossRef]
  12. Li, S.-F.; Zheng, J.; Hu, L.; Ma, Y.; Yan, D. Facile surface defect engineering on perovskite oxides for enhanced OER performance. Dalton Trans. 2023, 52, 4207–4213. [Google Scholar] [CrossRef]
  13. Pavlovska, O.B.; Vasylechko, L.O.; Lutsyuk, I.V.; Koval, N.M.; Zhydachevskii, Y.A.; Pieniazek, A. Structure Peculiarities of Micro- and Nanocrystalline Perovskite Ferrites La1−xSmxFeO3. Nanoscale Res. Lett. 2017, 12, 153. [Google Scholar] [CrossRef]
  14. Guo, B.-R.; Chen, M.-X.; Li, S.-W.; Gao, R.-H.; Sang, B.-H.; Ren, X.-Q.; Liu, Z.; Cao, X.; Liu, J.; Ding, Y.-N.; et al. Construction of iron oxyhydroxide/nickel sulfate hydroxide hybrid electrocatalyst for efficient oxygen evolution. Rare Met. 2024, 43, 6394–6404. [Google Scholar] [CrossRef]
  15. Qiao, Y.; Pan, Y.; Fan, W.; Long, G.; Zhang, F. Polyoxometalate-incorporated NiFe-based oxyhydroxides for enhanced oxygen evolution reaction in alkaline media. Chem. Commun. 2024, 60, 11287–11290. [Google Scholar] [CrossRef]
  16. Ahmad, M.H.; Islam, M.R.; Islam, M.R. Improved Electrochemical Performance of Co3O4 Incorporated MnO2 Nanowires for Energy Storage Applications. Arab. J. Sci. Eng. 2025, 50, 583–596. [Google Scholar] [CrossRef]
  17. Ahmed, S.; Lee, M.-C.; Rim, H.-R.; Lee, H.-K.; Shim, J.; Park, G. Highly porous Co3O4 and Ni-deposited Co3O4 nanoflowers as bifunctional catalysts for oxygen reduction and evolution reactions. Inorg. Chem. Commun. 2020, 113, 107799. [Google Scholar] [CrossRef]
  18. Quan, W.; Chen, T.; Sun, S.; Sun, C.; Wu, C. Facile hybrid strategy of SrCo0.5Fe0.3Mo0.2O3-dCo3O4 heterostructure for efficient oxygen evolution reaction. J. Alloys Compd. 2022, 917, 165508. [Google Scholar] [CrossRef]
  19. Zhang, C.; Antonietti, M.; Fellinger, T.-P. Blood Ties: Co3O4 Decorated Blood Derived Carbon as a Superior Bifunctional Electrocatalyst. Adv. Funct. Mater. 2014, 24, 7655–7665. [Google Scholar] [CrossRef]
  20. Lu, Z.; Zhou, H.; Qian, B.; Wang, S.; Zheng, Y.; Ge, L.; Chen, H. Y and Fe co-doped LaNiO3 perovskite as a novel bifunctional electrocatalyst for rechargeable zinc-air batteries. Int. J. Hydrogen Energy 2023, 48, 8082–8092. [Google Scholar] [CrossRef]
  21. Fan, Y.; Wang, W.; Chen, Y.; Xu, Z.; Cai, D.; Xu, M.; Tong, R. Cobalt-containing ZIF-derived catalysts for Zn-air batteries. Mater. Chem. Front. 2024, 8, 2394–2419. [Google Scholar] [CrossRef]
  22. Shang, Z.; Zhang, H.; Liang, J.; You, Z.; Wang, R.; Wan, L.; Lei, D.; Li, Z. Highly adhesive hydrogel electrolytes driven by adenosine monophosphate and Fe3+ for high-voltage asymmetric flexible zinc-air batteries. Energy Storage Mater. 2024, 71, 103599. [Google Scholar] [CrossRef]
  23. Chen, Z.; Chen, H.; Li, T.; Tian, X.; Zhang, K.; Miao, Y.; Xia, C.; Cai, L.; Hui, B.; Chen, C. Defective wood-based chainmail electrocatalysts boost performances of seawater-medium Zn-air batteries. J. Energy Chem. 2025, 102, 134–143. [Google Scholar] [CrossRef]
  24. Liu, T.; Dong, X.; Tang, B.; Zhao, R.; Xu, J.; Li, H.; Gao, S.; Fang, Y.; Chao, D.; Zhou, Z. Engineering electrolyte additives for stable zinc-based aqueous batteries: Insights and prospects. J. Energy Chem. 2024, 98, 311–326. [Google Scholar] [CrossRef]
  25. Harisha, B.S.; Akkinepally, B.; Shim, J.; Lim, J. NiO.Co3O4@TiO2: Multimetallic nanocomposite electrode for supercapacitors: Experimental evaluation and simulation insights. Chem. Eng. J. 2025, 508, 160207. [Google Scholar] [CrossRef]
  26. Yu, H.; Qin, S.; Qi, X.; Wang, K.; Li, Y.; Du, Y.; Feng, X.; Shan, W.; Xiong, Y. P123-assisted synthesis of Co3O4 catalyst for efficiently catalyzing N2O decomposition under simulated real tail-gases. Mol. Catal. 2024, 565, 114378. [Google Scholar] [CrossRef]
  27. Zheng, F.; Lu, J.; Liu, C.; Zheng, H.; Xu, Q. Synergistic engineering of oxygen vacancies, cation vacancies, and surface acidity in MOFs-derived Co-Mn spinel oxides via acid etching: A pathway to enhanced toluene oxidation performance. J. Hazard. Mater. 2025, 492, 138239. [Google Scholar] [CrossRef]
  28. Que, L.; Lu, L.; Xu, Y.; Xu, X.; Zhu, M.; Pan, J.; Cao, J.; Wang, J.; Zheng, Y.; Li, C. The Ni2+-LaNiO3/CdS hollow core-shell heterojunction towards enhanced visible light overall water splitting H2 evolution via HER/OER synergism of Ni2+/Ov. Chem. Eng. J. 2023, 469, 143902. [Google Scholar] [CrossRef]
  29. Sun, Y.; Wu, C.-R.; Ding, T.-Y.; Gu, J.; Yan, J.-W.; Cheng, J.; Zhang, K.H.L. Direct observation of the dynamic reconstructed active phase of perovskite LaNiO3 for the oxygen-evolution reaction. Chem. Sci. 2023, 14, 5906–5911. [Google Scholar] [CrossRef]
  30. Zhu, M.; Sheng, Z.; Gao, J.; Li, Y.; Zhang, C.; Zhu, X. Synthesis of porous LaNiO3 thin films by chemical solution deposition for enhanced oxygen evolution reaction. Dalton Trans. 2023, 52, 9903–9907. [Google Scholar] [CrossRef]
  31. Kuchipudi, A.; Madhu, R.; Arunmuthukumar, P.; Sundarravalli, S.; Sreedhar, G.; Kundu, S. Decoration of Au Nanoparticles over LaFeO3: A High Performance Electrocatalyst for Total Water Splitting. Inorg. Chem. 2023, 62, 14448–14458. [Google Scholar] [CrossRef]
  32. Zhang, Y.; Xu, H.; Liu, M.; Qi, J.; Hu, L.; Feng, M.; Lu, W. The modulated oxygen evolution reaction performance of LaFeO3 with abundant electronic structures via a design of stoichiometry offset. Phys. Chem. Chem. Phys. 2023, 25, 11725–11731. [Google Scholar] [CrossRef]
  33. Yu, J.; Yang, G.; Li, Z.; Zhu, W.; Jiang, S.; Chen, D.; Shao, Z.; Ni, M. Boosting the lattice oxygen reactivity of perovskite electrocatalyst via less Ru substitution. Int. J. Hydrogen Energy 2024, 84, 650–657. [Google Scholar] [CrossRef]
  34. Inoue, Y.; Miyahara, Y.; Miyazaki, K.; Lee, C.; Sakamoto, R.; Abe, T. Synergistic Interplay between Fe-Based Perovskite Oxides and Co in Electrolyte for Efficient Oxygen Evolution Reaction. Chemsuschem 2025, 18, e202401982. [Google Scholar] [CrossRef]
  35. Chen, J.; Huang, J.; Wang, H.; Feng, W.; Luo, T.; Hu, Y.; Yuan, C.; Cao, L.; Jie, Y.; Kajiyoshi, K.; et al. Phase-mediated cobalt phosphide with unique core-shell architecture serving as efficient and bifunctional electrocatalyst for hydrogen evolution and oxygen reduction reaction. Chin. Chem. Lett. 2022, 33, 3752–3756. [Google Scholar] [CrossRef]
  36. Wang, C.; Zhang, K.; Yang, B. Architecting double-shelled hollow carbon nanocages embedded bimetallic sites as bifunctional oxygen electrocatalyst for zinc-air batteries. Chin. Chem. Lett. 2025, 36, 110538. [Google Scholar] [CrossRef]
  37. Huang, K.; Hu, J.; Cao, J.; Wei, X.; Liu, S.; Dai, Q.; Shen, F.; Zhang, X.; Zhao, X.; Peng, Y.; et al. Biomass-templated strategy to synthesize Fe2P/Co2P heterojunction bifunctional electrocatalyst for high performance flexible zinc-air batteries. Sci. China Chem. 2025, 68, 3056–3063. [Google Scholar] [CrossRef]
  38. Saha, P.; Shah, S.S.; Ali, M.; Shaikh, M.N.; Aziz, M.A.; Saleh Ahammad, A.J. Cobalt Oxide-Based Electrocatalysts with Bifunctionality for High-Performing Rechargeable Zinc-Air Batteries. Chem. Rec. 2024, 24, e202300216. [Google Scholar] [CrossRef]
  39. Liang, Q.; Brocks, G.; Bieberle-Hutter, A. Oxygen evolution reaction (OER) mechanism under alkaline and acidic conditions. J. Phys. Energy 2021, 3, 026001. [Google Scholar] [CrossRef]
  40. Zhang, R.; Liu, W.; Zhang, F.-M.; Yang, Z.-D.; Zhang, G.; Zeng, X.C. COF-C4N Nanosheets with uniformly anchored single metal sites for electrocatalytic OER: From theoretical screening to target synthesis. Appl. Catal. B Environ. Energy 2023, 325, 122366. [Google Scholar] [CrossRef]
  41. Jiang, T.; Jiang, H.; Wang, W.; Mu, H.; Zhang, Y.; Li, B. Atomically Dispersed High-Active Site Density Copper Electrocatalyst for the Reduction of Oxygen. Materials 2024, 17, 5030. [Google Scholar] [CrossRef]
  42. Yoo, D.; Park, S.; Oh, S.; Kim, M.P.; Park, K. In Situ Analysis of Binder Degradation during Catalyst-Accelerated Stress Test of Polymer Electrolyte Membrane Fuel Cells. Materials 2024, 17, 4425. [Google Scholar] [CrossRef]
  43. Han, Y.; Jin, K.; Zhao, T.; Yu, Y. Recent advances in solid-state zinc-air and zinc-ion batteries. Chin. Chem. Lett. 2025, 111387, in press. [Google Scholar] [CrossRef]
  44. Zhang, D.; Song, Y.; Du, Z.; Wang, L.; Li, Y.; Goodenough, J.B. Active LaNi1−xFexO3 bifunctional catalysts for air cathodes in alkaline media. J. Mater. Chem. A 2015, 3, 9421–9426. [Google Scholar] [CrossRef]
  45. Gong, C.; Zhao, L.; Li, S.; Wang, H.; Gong, Y.; Wang, R.; He, B. Atomic layered deposition iron oxide on perovskite LaNiO3 as an efficient and robust bi-functional catalyst for lithium oxygen batteries. Electrochim. Acta 2018, 281, 338–347. [Google Scholar] [CrossRef]
  46. Chandrappa, S.G.; Moni, P.; Chen, D.; Karkera, G.; Prakasha, K.R.; Caruso, R.A.; Prakash, A.S. The influence of ruthenium substitution in LaCoO3 towards bi-functional electrocatalytic activity for rechargeable Zn-air batteries. J. Mater. Chem. A 2020, 8, 20612–20620. [Google Scholar] [CrossRef]
  47. Zheng, Q.L.; Zhang, Y.D.; Wang, C.C.; Zhang, C.H.; Guo, Y.M. CoO Enhanced Oxygen Evolution Kinetics of LaMnO3 Perovskite As a Potential Cathode for Rechargeable Zn-Air Batteries. Energy Fuels 2022, 36, 1091–1099. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of LYNF, Co3O4, and LYNF-xCo3O4 (x = 0.25,0.5,0.75).
Figure 1. XRD patterns of LYNF, Co3O4, and LYNF-xCo3O4 (x = 0.25,0.5,0.75).
Catalysts 15 00785 g001
Figure 2. (a) TEM images of LYNF-0.5Co3O4; (b) HRTEM images of LYNF-0.5Co3O4.
Figure 2. (a) TEM images of LYNF-0.5Co3O4; (b) HRTEM images of LYNF-0.5Co3O4.
Catalysts 15 00785 g002
Figure 3. (a) SEM images of LYNF-0.5Co3O4; (b–h) EDS images of LYNF-0.5Co3O4.
Figure 3. (a) SEM images of LYNF-0.5Co3O4; (b–h) EDS images of LYNF-0.5Co3O4.
Catalysts 15 00785 g003
Figure 4. XPS spectra of (a) Ni 2p, (b) Fe 2p, and (c) O 1s of LYNF and LYNF-0.5Co3O4.
Figure 4. XPS spectra of (a) Ni 2p, (b) Fe 2p, and (c) O 1s of LYNF and LYNF-0.5Co3O4.
Catalysts 15 00785 g004
Figure 5. (a) ORR LSV polarization curves were recorded in O2-saturated 0.1 M KOH at 1600 rpm and 5 mV s−1. (b) Tafel plots of prepared electrocatalysts for ORR. (c) OER LSV polarization curves were recorded in 0.1 M KOH at a scan rate of 5 mV s−1. (d) Tafel plots of prepared electrocatalysts for OER. (e) Nyquist plots of LYNF, LYNF-0.25Co3O4, LYNF-0.5Co3O4, and LYNF-0.75Co3O4 obtained from EIS measurements.
Figure 5. (a) ORR LSV polarization curves were recorded in O2-saturated 0.1 M KOH at 1600 rpm and 5 mV s−1. (b) Tafel plots of prepared electrocatalysts for ORR. (c) OER LSV polarization curves were recorded in 0.1 M KOH at a scan rate of 5 mV s−1. (d) Tafel plots of prepared electrocatalysts for OER. (e) Nyquist plots of LYNF, LYNF-0.25Co3O4, LYNF-0.5Co3O4, and LYNF-0.75Co3O4 obtained from EIS measurements.
Catalysts 15 00785 g005
Figure 6. CV curves for (a) LYNF, (b) LYNF-0.25Co3O4, (c) LYNF-0.5Co3O4, and (d) LYNF-0.75Co3O4 in 0.1 M KOH with different scan rates; (e) capacitive current densities of the prepared electrocatalysts at different scan rates. Scan rates are 20, 40, 60, 80, 100, and 120 mV s−1.
Figure 6. CV curves for (a) LYNF, (b) LYNF-0.25Co3O4, (c) LYNF-0.5Co3O4, and (d) LYNF-0.75Co3O4 in 0.1 M KOH with different scan rates; (e) capacitive current densities of the prepared electrocatalysts at different scan rates. Scan rates are 20, 40, 60, 80, 100, and 120 mV s−1.
Catalysts 15 00785 g006
Figure 7. (a,b) ORR LSV curves of LYNF-0.5Co3O4 and LYNF, respectively, before and after CV cycles; (c,d) corresponding OER LSV curves. All measurements are performed in O2-saturated 0.1 M KOH at a rotating speed of 1600 rpm and a scan rate of 5 mV s−1.
Figure 7. (a,b) ORR LSV curves of LYNF-0.5Co3O4 and LYNF, respectively, before and after CV cycles; (c,d) corresponding OER LSV curves. All measurements are performed in O2-saturated 0.1 M KOH at a rotating speed of 1600 rpm and a scan rate of 5 mV s−1.
Catalysts 15 00785 g007
Figure 8. (a) The schematic diagram of the zinc–air battery; (b) discharge polarization curves and corresponding power density curve for LYNF and LYNF-0.5Co3O4; (c) constant current cycle curve at 5 mA cm−2 for LYNF and LYNF-0.5Co3O4.
Figure 8. (a) The schematic diagram of the zinc–air battery; (b) discharge polarization curves and corresponding power density curve for LYNF and LYNF-0.5Co3O4; (c) constant current cycle curve at 5 mA cm−2 for LYNF and LYNF-0.5Co3O4.
Catalysts 15 00785 g008
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

Yi, X.; Zhuang, G.; Bai, J.; Yan, J.; Zheng, Y. Boosting the Bifunctional Catalytic Activity of La0.85Y0.15Ni0.7Fe0.3O3 Perovskite Air Electrode with Facile Hybrid Strategy of Metallic Oxide for Rechargeable Zn–Air Batteries. Catalysts 2025, 15, 785. https://doi.org/10.3390/catal15080785

AMA Style

Yi X, Zhuang G, Bai J, Yan J, Zheng Y. Boosting the Bifunctional Catalytic Activity of La0.85Y0.15Ni0.7Fe0.3O3 Perovskite Air Electrode with Facile Hybrid Strategy of Metallic Oxide for Rechargeable Zn–Air Batteries. Catalysts. 2025; 15(8):785. https://doi.org/10.3390/catal15080785

Chicago/Turabian Style

Yi, Xiankai, Guangwei Zhuang, Junhua Bai, Jiaxing Yan, and Yifeng Zheng. 2025. "Boosting the Bifunctional Catalytic Activity of La0.85Y0.15Ni0.7Fe0.3O3 Perovskite Air Electrode with Facile Hybrid Strategy of Metallic Oxide for Rechargeable Zn–Air Batteries" Catalysts 15, no. 8: 785. https://doi.org/10.3390/catal15080785

APA Style

Yi, X., Zhuang, G., Bai, J., Yan, J., & Zheng, Y. (2025). Boosting the Bifunctional Catalytic Activity of La0.85Y0.15Ni0.7Fe0.3O3 Perovskite Air Electrode with Facile Hybrid Strategy of Metallic Oxide for Rechargeable Zn–Air Batteries. Catalysts, 15(8), 785. https://doi.org/10.3390/catal15080785

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

Article Metrics

Back to TopTop