Next Article in Journal
Adsorption Isotherm Analysis for Hybrid Molecularly Imprinted Polymeric Gold-Decorated Nanoparticles Suitable for Reliable Quantification of Gluconic Acid in Wine
Previous Article in Journal
Use of Antimicrobial Nanoparticles for the Management of Dental Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Performance of Nickel-Doped LaSrCoO3-SrCO3 Composite Materials for Alkaline Oxygen Evolution in Water Splitting

1
School of Mechanical and Electronic Engineering, Suzhou University, Suzhou 234000, China
2
Key Laboratory of Power Station Energy Transfer Conversion and System of MOE, School of Energy Power and Mechanical Engineering, North China Electric Power University, Beijing 102206, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(3), 210; https://doi.org/10.3390/nano15030210
Submission received: 28 December 2024 / Revised: 23 January 2025 / Accepted: 26 January 2025 / Published: 28 January 2025
(This article belongs to the Special Issue Development and Synthesis of New Nanostructured Catalysts)

Abstract

:
Perovskites exhibit catalytic properties on the oxygen evolution reaction (OER) in water electrolysis. Elemental doping by specific preparation methods is a good strategy to obtain highly catalytical active perovskite catalysts. In this work, La0.5Sr0.5Co1−xNixO3−δ perovskite materials doped with different ratios of nickel were successfully synthesized by the sol-gel method. The electrochemical measurement results show that for OER in 1 M KOH solution, La0.5Sr0.5Co0.8Ni0.2O3−δ prepared by the sol-gel method requires only a low overpotential of 213 mV to reach 10 mA cm−2, which is significantly lower than that of La0.5Sr0.5Co0.8Ni0.2O3−δ prepared by the hydrothermal method for the increasing about 45.24% (389 mV at 10 mA cm−2). In addition, La0.5Sr0.5Co0.8Ni0.2O3−δ by the sol-gel method can be kept stable in an alkaline medium tested for 30 h without degradation. This indicates that the prepared La0.5Sr0.5Co0.8Ni0.2O3−δ has better OER performance. The X-ray diffraction (XRD) results show that SrCO3 is the main phase formed, which is a disadvantage of this method. The performance improvement may be affected by the carbonate phase. The scanning electron microscopy (SEM) results show that layer structured La0.5Sr0.5Co0.8Ni0.2O3−δ by the sol-gel method has more surface pores with a pore diameter of about 0.362 μm than spherical granular structured La0.5Sr0.5Co0.8Ni0.2O3−δ by the hydrothermal method. X-ray photoelectronic spectroscopy (XPS) results reveal that the crystal lattice of La0.5Sr0.5Co0.8Ni0.2O3−δ by nickel doping is lengthened, and the electronic configuration of Co is also changed by the sol-gel preparation process. The improved electrocatalytic performance of La0.5Sr0.5Co0.8Ni0.2O3−δ may be attributed to the pore structure formed providing more active sites during the sol-gel process and the improved oxygen mobility with Ni doping by the sol-gel method. The doping strategy using the sol-gel method provides valuable insights for optimizing perovskite catalytic properties.

1. Introduction

Hydrogen energy is a highly efficient and clean source of green energy, widely utilized for energy storage and conversion, which aids in reducing reliance on fossil fuels [1]. However, the slow kinetics of the alkaline oxygen evolution reaction (OER) remain a major challenge, limiting the efficiency of hydrogen production from water electrolysis and hindering the industrial scalability of related technologies [2,3]. It is known that transition metal oxides, hydroxides (e.g., NiFeOH), and nickel-iron alloys exhibit good OER electrocatalytic performance, but their limited stability still restricts their use [4]. Therefore, developing new catalysts that combine high performance with durability is essential. Among potential alternatives, cobalt-based perovskites, such as PrCoO3, BaCoO3, and SrCoO3, have emerged as highly effective electrocatalysts for OER due to their favorable physicochemical properties, tunable structures, and cost-effectiveness [5,6]. In particular, recent studies highlight the advantages of cobalt-based perovskites in achieving superior catalytic performance for OER. For example, Zhao et al. [7] demonstrated that the perovskite Pr0.7Sr0.3Co0.95Ru0.05O3 exhibited a lower OER overpotential than the commercial RuO2 in 1 M KOH. Similarly, Ba0.5Sr0.5Co0.8Fe0.2O3−δ was found to exhibit significantly better OER catalytic activity than iridium oxide catalysts [8]. These findings underscore the potential of cobalt-based perovskites as promising candidates for OER electrocatalysis in alkaline media.
Recent studies have demonstrated that doping and the incorporation of metal supports can significantly enhance the catalytic performance of perovskite materials [9]. For instance, combining Co3O4 with silver (Ag) or nickel (Ni) substrates has shown a synergistic effect, enhancing both oxygen evolution reaction (OER) and oxygen reduction reaction (ORR) activities [10]. Soltani et al. [11] demonstrated that metal-supported perovskite catalysts, such as Ag-supported Co3O4, exhibited significant improvements in electrocatalytic performance compared to bare metals or oxides, with the use of Ag as a substrate, reducing the overpotential of the OER by up to 40 mV. This enhanced performance can be attributed to several factors, including the higher electrical conductivity of the metal support, reduced contact resistance between the oxide and substrate, and electronic effects stemming from interactions between the metal and oxide components [12]. Additional studies have confirmed the potential of bimetallic systems. For instance, Hatem et al. [13] combined Ag with Co3O4 nanoparticles, demonstrating enhanced OER and ORR catalytic activity, with an ORR overpotential only 70 mV higher than that of commercial Pt catalysts. Moreover, metal-doped perovskites can significantly enhance OER activity by altering the electronic structure of the catalyst [14,15,16]. For example, Zhu et al. [17] synthesized a tetragonal structure of SrCo0.95P0.05O3−δ by doping phosphorus into SrCoO3−δ. The synthesized SrCo0.95P0.05O3−δ material achieved a current density of 10 mA cm−2 at a relatively small overpotential of 0.48 V. Luo et al. [18] synthesized Ba0.9Sr0.1Co0.8Fe0.1Ir0.1O3−δ, which demonstrated excellent OER activity under alkaline conditions. This enhancement is attributed to iridium substitution for iron, resulting in an overpotential of 300 mV at a current density of 10 mA cm−2, with stability maintained for 10 h in a 1.0 M potassium hydroxide electrolyte. It is noteworthy that among various dopants, nickel is the most effective candidate for enhancing the OER reaction [19]. Han et al. [20] synthesized nickel-doped perovskites (La5Ni3Co2) by optimizing the molar ratios of La, Ni, and Co. La5Ni3Co2 exhibited a current density of 10 mA cm−2 in 1.0 M potassium hydroxide at 0.360 V. These findings emphasize the importance of substrates and doping in optimizing the performance of perovskite-based catalysts. In contrast, there are relatively few studies on nickel-doped LaSrCoO3−δ for alkaline OER, and further research on the catalytic performance of nickel-doped LaSrCoO3−δ under various conditions is needed.
To date, various methods have been developed to synthesize these highly active perovskite materials, including electrostatic spinning, solution combustion, high-temperature methods, and hydrothermal techniques [21,22]. Niu et al. [23] prepared the La0.6Sr0.4Co0.8Ni0.2O3−δ (LSCN-0.8) electrocatalyst using the electrostatic spinning method. LSCN-0.8 exhibited an overpotential of 327 mV in 1 M KOH at a current density of 10 mA cm−2. Saraswati et al. [24] synthesized La0.6Sr0.4Co0.8Ni0.2O3 using the solution combustion method, achieving a low overpotential of 250 mV at a current density of 10 mA cm−2. However, the perovskite materials prepared by the methods described above often suffer from grain agglomerations and inhomogeneity during the sintered process. The hydrothermal method has been particularly effective in achieving uniform particle sizes, resulting in perovskite materials with enhanced catalytic properties [25]. For example, the LaNiO3 electrocatalysts synthesized using the hydrothermal method exhibited a porous, hollow structure and remarkable stability, outperforming Pt/C catalysts in long-term stability tests [26]. In addition, perovskite materials can also be prepared using the sol-gel method. For example, Omari et al. [27] prepared LaNi1−xCoxO3 nanoparticles with a porous structure using the sol-gel method. Their optical band gap exceeded 3 eV, indicating that these perovskite materials are good semiconductors and catalysts. Liang et al. [28] synthesized impurity-free La0.4Sr0.6Co0.8Ni0.2O3 using the sol-gel method, exhibiting fewer agglomerates and a higher density of small particles and interstitials on the La0.4Sr0.6Co0.8Ni0.2O3 surface. This result indicates that high-purity, less agglomerated perovskite materials with controllable particle sizes can be successfully prepared via the sol-gel method. However, there are limited studies regarding the preparation of the LaSrCoO3−δ system by the sol-gel method.
Herein, we synthesized nickel-doped perovskites La0.5Sr0.5Co1−xNixO3−δ (x = 0.2, 0.5, 0.8) as electrocatalysts for the alkaline oxygen evolution reaction (OER) using the sol-gel method. The sol-gel method is employed to controllably synthesize highly homogeneous materials, which is anticipated to enhance the electrocatalytic performance of perovskite structures. As a comparison, a series of La0.5Sr0.5Co1−xNixO3−δ (x = 0.2, 0.5, 0.8) was synthesized using the hydrothermal method. Electrochemical measurements were conducted to evaluate the OER properties of La0.5Sr0.5Co1−xNixO3−δ. Structural and morphological features were analyzed using various characterization techniques such as X-ray diffraction (XRD) and scanning electron microscopy (SEM). As the OER electrocatalyst, La0.5Sr0.5Co1−xNixO3−δ synthesized via the sol-gel method was expected to exhibit good performance at certain ratios of cobalt/nickel by doping Ni, demonstrating a high current density, low overpotential, and stability in 1.0 M KOH electrolytes. Additionally, X-ray photoelectron spectroscopy (XPS) was used to investigate the relationship between the electrocatalytic properties of La0.5Sr0.5Co1−xNixO3−δ by the sol-gel method with a specific ratio of cobalt/nickel and the change in the electronic configuration of Co. Furthermore, the reason for the good OER catalytic performance of the La0.5Sr0.5Co1−xNixO3−δ catalyst was explored.

2. Experimental Section

La0.5Sr0.5Co1−xNixO3−δ perovskite materials were prepared by the sol-gel method and the hydrothermal method, respectively. The raw materials, equipment, and calcination processes were the same except for the preparation methods. In addition, the ratios of nickel and cobalt were also consistent. The primary difference was the preparation process by the precursors.

2.1. Preparation of La0.5Sr0.5Co1−xNixO3−δ Using the Sol-Gel Method

The molar ratio of cobalt and nickel was 10:1. Lanthanum nitrate (5 mmol), strontium nitrate (5 mmol), nickel nitrate (8 mmol), and cobalt nitrate (2 mmol) were dissolved in 80 mL of deionized water. Citric acid (5 mmol) was added, and the solution was ultrasonically dispersed for 10 min. After complete dispersion, a magnetic stir bar was placed in the beaker, which was then positioned on a magnetic heating stirrer rotating at 400 rpm, maintaining a heating temperature of 95 °C. The solution evaporated until it became viscous. Heating and stirring were stopped, and the stir bars were removed once the beaker cooled to room temperature. Subsequently, the beaker underwent further drying in a blast drying oven at 180 °C for 6 h. The dried material was then ground in a mortar and transferred to an alumina crucible, followed by placement in a muffle furnace. The furnace temperature was ramped from room temperature to 750 °C at 3 °C/min, held at 750 °C for 4 h, and naturally cooled to room temperature. This synthesis process was repeated for perovskite materials with x values of 0.5 and 0.8, maintaining consistent conditions, named as La0.5Sr0.5Co0.5Ni0.5O3−δ-S, La0.5Sr0.5Co0.5Ni0.8O3−δ-S, La0.5Sr0.5CoO3−δ-S, and La0.5Sr0.5Co0.2Ni0.8O3−δ-S.

2.2. Preparation of La0.5Sr0.5Co1−xNixO3−δ by the Hydrothermal Method

In a beaker containing 60 mL of deionized water, 5 mmol each of lanthanum nitrate and strontium nitrate, 8 mmol of nickel nitrate, and 2 mmol of cobalt nitrate were dissolved. Then, 5 mmol of citric acid was added, and the solution was stirred continuously at 400 rpm for 2 h at room temperature. The resulting solution was transferred to a 100 mL Teflon hydrothermal reactor vessel and heated in an oven at 180 °C for 24 h. After cooling naturally, the product in the reactor liner was removed and washed with deionized water and ethanol using centrifugation. Once washing was complete, the product was placed in a blast drying oven at 110 °C for 12 h. The dried powder was then transferred to an alumina crucible, which was placed in a muffle furnace and heated at a rate of 3 °C/min until 750 °C. The powder was held at 750 °C for 4 h before cooling naturally to room temperature. The resulting material was La0.5Sr0.5Co0.8Ni0.2O3−δ. In the same preparation process, the cobalt-nickel ratio was adjusted to produce La0.5Sr0.5CoO3−δ-H, La0.5Sr0.5Co0.5Ni0.5O3−δ-H, La0.5Sr0.5Co0.8Ni0.2O3−δ-H, and La0.5Sr0.5Co0.2Ni0.8O3−δ-H.

2.3. Physical Characterization

Thermo ScientificTM (Waltham, MA USA) Quattro S energy dispersive X-ray spectrum (EDS) and scanning electron microscopy (SEM) were used to determine the micromorphology and element content distribution of samples, respectively. A Panalytical X-ray diffractometer (Malvern, UK) equipped with an Al Kα radiation source (λ = 1.50) was used to get the X-ray diffraction (XRD) tests. A Thermo ScientificTM ESCALAB 250Xi (Waltham, MA USA) spectrometer with a monochromatic Al Kα X-ray source (1486.6 eV) running at 200 W was used to perform the X-ray photoelectron spectroscopy (XPS). Prior to fitting the experimental peak with the XPSPEAK software (Version 4.1), all XPS peaks were calibrated using C 1s peak binding energy for adventitious carbon, which is 284.8 eV. All characterization experiments were performed at room temperature and atmospheric pressure.

2.4. Electrochemical Measurements

A typical three-electrode system in 1.0 M KOH electrolyte solution was used for all electrochemical testing in the electrochemical workstation (Bio-Logic VMP3, Seyssinet-Pariset, France). The electrolytes were saturated with N2 bubbles for 10 min before the electrochemical experiments. At a sweep rate of 5 mV s−1, the linear sweep voltammetry (LSV) curves were obtained for OER with 90% iR correction. The Tafel slopes were determined by fitting the overpotential-current density curves using the formula (η = a + b log∣j∣), where j, η, and b stand for the current density, overpotential, and Tafel slope, respectively. Electrochemical impedance spectroscopy (EIS) was performed with an amplitude of 5 mV at overpotentials of −0.1 V and 0.3 V between 100 kHz and 100 MHz.

3. Results and Discussion

3.1. Materials Characterization

The crystal structures of La0.5Sr0.5Co1−xNixO3−δ-S prepared by the sol-gel method were analyzed using XRD. As depicted in Figure 1, the diffraction peaks observed at 23.2°, 33.1°, 40.7°, 47.4°, 58.9°, 69.3°, and 78.9° in the four samples correspond to crystallographic planes of (012), (104), (202), (024), (300), (208), and (128) of La0.5Sr0.5CoO2.91 (PDF#48-0122). The diffraction peaks in the nickel-doped samples exhibit a leftward shift compared to the standard card, indicating that nickel ion doping has increased the lattice constant and expanded the crystal lattice spacing. Additionally, peaks at 25.2°, 25.8°, 36.5°, 44.1°, and 49.9° in the four samples correspond to SrCO3 (PDF#05-0418), with crystallographic planes identified as (111), (021), (130), (221), and (113), respectively. These peaks are significantly stronger than those of La0.5Sr0.5CoO2.91, suggesting that SrCO3 is formed as a main phase. A leftward shift of peaks described above demonstrates the coexistence of SrCO3 alongside La0.5Sr0.5CoO2.91 during the preparation process. As a result, the formation of SrCO3 is a drawback of the sol-gel method, which could be improved to reduce carbonate formation. Moreover, nickel-doped samples also exhibit small diffraction peaks at 43.0° and 62.7°. However, the La0.5Sr0.5CoO2.91 crystalline structure remains predominant in all samples, indicating that the perovskite structure formed after nickel doping resembles that of the pre-doped materials. Notably, the diffraction peak of La0.5Sr0.5Co0.8Ni0.2O3−δ is more prominent in the prepared samples.
To obtain the morphologies of Sr0.5Co0.8Ni0.2O3−δ, Field emission scanning electron microscopy was used to analyze the morphology and surface element distribution of La0.5Sr0.5Co0.8Ni0.2O3−δ prepared by the sol-gel method and hydrothermal method. Figure 2a–d displays the morphology of La0.5Sr0.5Co1−xNixO3−δ-S series materials synthesized using the sol-gel method at low magnification, showing a lumpy and lamellar structure across all materials. During oven drying, citric acid in the gel structure swells and combusts, producing carbon dioxide that escapes. This results in the formation of porous voids in the material. The structure of La0.5Sr0.5Co1−xNixO3−δ consists of multiple laminated layers stacked on top of each other, with small debris on the surface without part of the main stack. Comparing materials with different nickel doping ratios, it is observed that La0.5Sr0.5Co0.8Ni0.2O3−δ-S and La0.5Sr0.5Co0.5Ni0.5O3−δ-S have similar surfaces at low magnification. However, most blocks have several irregularly distributed, more pronounced pores on their surfaces. The La0.5Sr0.5Co0.2Ni0.8O3−δ-S appears as broken blocks with few apparent holes, indicating gas escape during expansion and resulting in fragmentation.
Figure 3 illustrates the elemental distribution on the surface of La0.5Sr0.5Co0.8Ni0.2O3−δ-S. The elements La, Sr, Co, Ni, and O are evenly distributed on the surface. Some elements appear shaded or in darker colors, which mostly correspond to holes in the actual sample.
Figure 4 depicts the morphology of La0.5Sr0.5Co1−xNixO3−δ-H series materials synthesized using the hydrothermal method. These materials take on the shape of spherical particles. The overall volume is smaller than that of the blocks prepared by the sol-gel method; some particles are broken, and there is no apparent pore structure on their surfaces. Some particles are fragmented, and their surfaces lack apparent pore structures. A comparison of the materials with different cobalt-nickel doping ratios revealed that the particles of La0.5Sr0.5Co0.8Ni0.2O3−δ-H were significantly smaller than those with other ratios. The particle structures of all four samples are indistinguishable, with particles ranging in size from micrometers and covered with dendrites. La0.5Sr0.5Co0.8Ni0.2O3−δ-H is notable for its smaller particle size and dendrites measuring 50 to 60 nm upon closer examination.
The energy dispersive spectroscopy (EDS) of hydrothermally synthesized La0.5Sr0.5Co0.8Ni0.2O3−δ-H is shown in Figure 5. The sample surface exhibits a uniform distribution of five elements, with strontium being the least uniformly distributed among them. The prominent distribution of lanthanides suggests that a substantial fraction of these elements is concentrated on the surface of the sample.

3.2. XPS Analysis of La0.5Sr0.5Co0.8Ni0.2O3−δ by the Sol-Gel Method

X-ray photoelectron spectroscopy (XPS) analysis of La0.5Sr0.5Co0.8Ni0.2O3−δ-S, prepared via the sol-gel method, shows significant shifts in the binding energies of La, Sr, Co, Ni, and O, indicating changes in the material’s electronic structure due to nickel doping. The binding energy peaks for La 3d, Sr 3d, Co 2p, Ni 2p, and O 1s, shown in Figure 6, were observed at characteristic values corresponding to their respective oxidation states, confirming the presence of La3+, Sr2+, and Co3+ in the samples. The La 3d peaks at 833.5 eV and 837.1 eV for the 3d5/2 orbitals and at 849.7 eV and 854.2 eV for the 3d3/2 orbitals, along with the Sr 3d peaks at 133.5 eV and 135.1 eV, matched the expected oxidation states [29,30]. Co 2p peaks at 779.6 eV and 794.9 eV indicate Co3+, with a shift in the Co 2p binding energy, suggesting an enhancement in Co oxidation state linked to improved electronic conductivity and catalytic activity [31]. Although the overlapping La peaks obscure the Ni 2p features, the positions for Ni2+ can be approximately identified [32]. The O 1s spectrum showed three distinct components: lattice oxygen (O2−) at 528.4 eV, surface-adsorbed oxygen at 531.6 eV, and CO32− at 532 eV, with the oxygen p-band center near the Fermi level [33]. Nickel substitution is thought to increase the B–O bond distance, which enhances both ionic and electronic conductivity by modifying Co electronic structure and shifting the diffraction peaks left [34,35]. These structural changes, including lattice extension, oxygen mobility enhancement, and electronic configuration modification, are consistent with previous reports [36,37] on doped perovskite-type oxides, where similar doping strategies improved catalytic performance. In conclusion, the synergistic effect of lattice extension and electronic modification improves the performance of La0.5Sr0.5Co0.8Ni0.2O3−δ-S.

3.3. Electrochemical Performance

The OER catalytic performance of La0.5Sr0.5Co1−xNixO3−δ-S was evaluated, as shown in Figure 7. Figure 7a shows that the LSV curves indicate that La0.5Sr0.5Co0.8Ni0.2O3−δ-S requires only 213 mV and 320 mV overpotentials to achieve current densities of 10 mA cm−2 and 50 mA cm−2, respectively. La0.5Sr0.5Co0.8Ni0.2O3−δ-S demonstrated excellent OER performance. Subsequently, La0.5Sr0.5CoO3−δ-S shows the next best performance, with current densities of 10 mA cm−2 at 397 mV overpotential and 50 mA cm−2 at 492 mV overpotential. La0.5Sr0.5Co0.5Ni0.5O3−δ-S and La0.5Sr0.5Co0.2Ni0.8O3−δ-S exhibited comparatively poor performance. The former exhibited an initial overpotential of 390 mV at 10 mA cm−2, but the current density did not increase proportionately thereafter. The latter displayed weak trends in both initial overpotential and the subsequent increase in current density, suggesting a potential deficiency in OER catalytic performance. Tafel plots were also used to analyze reaction kinetics. La0.5Sr0.5Co0.8Ni0.2O3−δ-S showed a Tafel slope of 180 mV dec−1, while La0.5Sr0.5CoO3−δ-S exhibited a lower slope of 150 mV dec−1. La0.5Sr0.5Co0.5Ni0.5O3−δ-S had a Tafel slope of 538 mV dec−1, indicating severely restricted reaction kinetics. In the impedance test, La0.5Sr0.5CoO3−δ-S, La0.5Sr0.5Co0.8Ni0.2O3−δ-S, and La0.5Sr0.5Co0.5Ni0.5O3−δ-S displayed similar electrical impedance values of 23.58 Ω,17.56 Ω, and 36.07 Ω, respectively. The fitting graph and relevant parameters of EIS are given in Figure S2 and Table S1 (see Supplementary Materials). In contrast, La0.5Sr0.5Co0.2Ni0.8O3−δ-S had a much higher impedance. This result mirrors the pattern observed in samples prepared via the hydrothermal method, suggesting that difficult electron transfer during catalysis significantly impacts the performance of the material. Additionally, as shown in Figure 7d, the overpotential of this type of material at a current density of 10 mA cm−2 is summarized by other researchers. This work has the lowest overpotential, proving that La0.5Sr0.5Co0.8Ni0.2O3−δ-S has excellent catalytic performance.
The OER catalytic performance of the La0.5Sr0.5Co1−xNixO3−δ-H was tested. The results are shown in Figure 8. The LSV performance curves reveal that La0.5Sr0.5Co0.8Ni0.2O3−δ-H exhibits good performance, with current densities of 10 mA cm−2 and 50 mA cm−2 and overpotentials of 389 mV and 490 mV, respectively. The Tafel value of La0.5Sr0.5Co0.8Ni0.2O3−δ-H is 179 mV dec−1. La0.5Sr0.5Co0.2Ni0.8O3−δ-H has the worst negligible OER performance. Overall, the catalytic performance through La0.5Sr0.5Co1−xNixO3−δ-H is much worse than that of La0.5Sr0.5Co1−xNixO3−δ-S, with a lower current density. The EIS results show that the catalytic performances of La0.5Sr0.5Co0.5O3−δ-H, La0.5Sr0.5Co0.8Ni0.2O3−δ-H, and La0.5Sr0.5Co0.5Ni0.5O3−δ-H have electrical impedance values of 32.22 Ω, 33.31 Ω, and 31.29 Ω, respectively, and the differences between them are minor. The fitting graph and relevant parameters of EIS are given in Figure S3 and Table S1. However, La0.5sr0.5Co0.2Ni0.8O3−δ-H also has a higher resistance, and the poor performance may be due to its extremely high electrical impedance, which hinders charge transfer during electrolysis.
La0.5Sr0.5Co1−xNixO3−δ-S was scanned at various speeds to generate cyclic voltammetry curves for each series and to plot the double-layer capacitance slope. According to Figure 9, La0.5Sr0.5Co0.8Ni0.2O3−δ-S exhibits the highest capacitance among the catalysts in this series, measuring 13.7 mF cm−2. La0.5Sr0.5Co0.2Ni0.8O3−δ-S follows closely with 8.2 mF cm−2. La0.5Sr0.5Co0.5Ni0.5O3−δ-S and La0.5Sr0.5CoO3−δ-S exhibit lower capacitances of 4.0 mF cm−2 and 1.7 mF cm−2, respectively. Combined SEM and double-layer capacitance analyses revealed that the morphology of the catalyst significantly affects the electrochemically active surface area, thereby influencing the OER activity.
During the cyclic voltammetry (CV) test, the cyclic voltammetry curves of the La0.5Sr0.5Co1−xNixO3−δ-H show remarkably low current densities. Despite its superior oxygen evolution performance, La0.5Sr0.5Co0.8Ni0.2O3−δ-H exhibits current densities less than one-tenth of those La0.5Sr0.5Co0.8Ni0.2O3−δ-S. The double-layer capacitance for La0.5Sr0.5Co0.8Ni0.2O3−δ-H is negligible according to CV calculations.
The comparison of SEM results from the two methods, using the double-layer capacitance test, revealed that the La0.5Sr0.5Co0.8Ni0.2O3−δ-S forms a porous structure during preparation. This improvement enhances the catalyst’s specific surface area, thereby increasing the number of active sites. In contrast, La0.5Sr0.5Co0.8Ni0.2O3−δ-H prepared by the hydrothermal method shows either complete or broken particle morphologies, with larger particle sizes and no change in specific surface area. Comprehensive analyses of bilayer capacitance tests indicate that the controllable preparation of the material morphology can have an impact on the catalytic activity of the material.
Stability testing of La0.5Sr0.5Co0.8Ni0.2O3−δ-S demonstrated the best OER catalytic performance. Nickel foam was selected as the substrate to prevent catalyst detachment and oxidation of the glassy carbon electrode surface during stability testing for oxygen evolution reactions. The catalyst was drop-coated onto the nickel foam, ensuring a consistent loading per unit area equivalent to that on the glassy carbon electrode during testing. Figure 10 shows that La0.5Sr0.5Co0.8Ni0.2O3−δ-S exhibits excellent catalytic stability even without pre-activation. Its potential remained stable, varying by approximately 10 mV over 30 h, demonstrating its robustness in alkaline conditions for OER.

4. Conclusions

In summary, nickel-doped perovskite oxides La0.5Sr0.5Co1−xNixO3−δ-S (x = 0.2, 0.5, 0.8) were synthesized using the sol-gel method. SEM and EDS results indicate that the La0.5Sr0.5Co1−xNixO3−δ-S presents numerous surface pores with a multilayered laminar structure and a homogeneous distribution of surface elements. La0.5Sr0.5Co1−xNixO3−δ-H prepared by the hydrothermal method shows spherical granular. La0.5Sr0.5Co0.8Ni0.2O3−δ-S requires only 213 mV and 320 mV overpotentials to achieve current densities of 10 mA cm−2 and 50 mA cm−2, respectively, for the oxygen evolution reaction (OER). La0.5Sr0.5Co0.8Ni0.2O3−δ-S shows excellent stability in alkaline solutions, and the potential change remains within about 10 mV after 30 h of stabilization tests. Comprehensive XRD and XPS analyses show that nickel doping extended the lattice and increased the B-O spacing, which changed the electronic configuration of cobalt and increased the oxygen mobility in the sol-gel process. As a result, the catalytic performance of the La0.5Sr0.5Co0.8Ni0.2O3−δ-S material for alkaline OER is improved. La0.5Sr0.5Co0.8Ni0.2O3−δ-S is a highly efficient, durable, and cost-effective electrocatalyst for water splitting in alkaline solutions. It offers valuable insights for improving the alkaline OER catalytic performance of non-precious metal perovskite materials in the future.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano15030210/s1, Figure S1. The oxidation peak images of La0.5Sr0.5Co0.5Ni0.5O3−δ-S and La0.5Sr0.5Co0.5Ni0.8O3−δ-S; Figure S2. Electrochemical impedance spectroscopy (EIS) plots of La0.5Sr0.5CoO3−δ-S, La0.5Sr0.5Co0.8Ni0.2O3−δ-S, La0.5Sr0.5Co0.5Ni0.5O3−δ-S and La0.5Sr0.5Co0.8Ni0.2O3−δ-S; Figure S3. Electrochemical impedance spectroscopy (EIS) plots of La0.5Sr0.5CoO3−δ-H, La0.5Sr0.5Co0.8Ni0.2O3−δ-H, La0.5Sr0.5Co0.5Ni0.5O3−δ-H and La0.5Sr0.5Co0.8Ni0.2O3−δ-H; Table S1. Relevant parameters in electrochemical impedance spectroscopy (EIS) fitting.

Author Contributions

Conceptualization, B.Z. and X.F.; methodology, X.P. and L.Z.; validation, X.P., L.Z. and D.C.; investigation, B.Z., L.Z. and X.F.; resources, F.Y.; data curation, B.W., X.F. and M.G.; writing—original draft, X.P., B.W. and D.C.; writing—reviewing and editing, F.Y.; visualization, M.G.; supervision, B.Z.; project administration, F.Y.; funding acquisition, F.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 22278125), the Quality Engineering Project of Suzhou University (No: szxy2023xxhz02), and the key scientific research project of Suzhou University (No: 2022xhx251, 2024yzd18).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chen, X.; Zhang, X.; Zhao, Y. Metal-organic framework-based hybrids with photon upconversion. Chem. Soc. Rev. 2024, 54, 152–177. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, H.; Wang, H.; Hu, Q.; Wang, B.; Lei, X.; You, J.; Guo, R. The latest advances in the deep reconstruction of pre-catalysts for the oxygen evolution reaction. J. Alloys Compd. 2025, 1010, 177225. [Google Scholar] [CrossRef]
  3. Zhu, S.; Song, L.; Xu, Z.; Chen, F.; Tao, H.; Tang, X.; Wang, Y. Theoretical and experimental aspects of electrocatalysts for oxygen evolution reaction. Chemistry 2024, 30, e202303672. [Google Scholar] [CrossRef]
  4. Xie, X.; Du, L.; Yan, L.; Park, S.; Qiu, Y.; Sokolowski, J.; Wang, W.; Shao, Y. Oxygen evolution reaction in alkaline environment: Material challenges and solutions. Adv. Funct. Mater. 2022, 32, 2110036. [Google Scholar] [CrossRef]
  5. Liang, X.; Zhang, K.-X.; Shen, Y.-C.; Sun, K.; Shi, L.; Chen, H.; Zheng, K.-Y.; Zou, X.-X. Perovskite-type water oxidation electrocatalysts. J. Electrochem. 2022, 28, 2214004. [Google Scholar]
  6. Hwang, J.; Rao, R.R.; Giordano, L.; Katayama, Y.; Yu, Y.; Shao-Horn, Y. Perovskites in catalysis and electrocatalysis. Science 2017, 358, 751–756. [Google Scholar]
  7. Zhao, Y.-N.; Liu, C.; Xu, S.; Min, S.; Wang, W.; Mitsuzaki, N.; Chen, Z. A/B-site management strategy to boost electrocatalytic overall water splitting on perovskite oxides in an alkaline medium. Inorg. Chem. 2023, 62, 12590–12599. [Google Scholar] [CrossRef]
  8. Suntivich, J.; May, K.J.; Gasteiger, H.A.; Goodenough, J.B.; Shao-Horn, Y. A perovskite oxide optimized for oxygen evolution catalysis from molecular orbital principles. Science 2011, 334, 1383–1385. [Google Scholar] [CrossRef]
  9. Flores-Lasluisa, J.X.; Huerta, F.; Cazorla-Amorós, D.; Morallón, E. LaNi1-xCoxO3 perovskites for application in electrochemical reactions involving molecular oxygen. Energy 2023, 273, 127256. [Google Scholar] [CrossRef]
  10. Lee, D.U.; Li, J.; Park, M.G.; Seo, M.H.; Ahn, W.; Stadelmann, I.; Ricardez-Sandoval, L.; Chen, Z. Self-assembly of spinel nanocrystals into mesoporous spheres as bifunctionally active oxygen reduction and evolution electrocatalysts. ChemSusChem 2017, 10, 2258–2266. [Google Scholar] [CrossRef]
  11. Soltani, M.; Amin, H.M.A.; Cebe, A.; Ayata, S.; Baltruschat, H. Metal-supported perovskite as an efficient bifunctional electrocatalyst for oxygen reduction and evolution: Substrate effect. J. Electrochem. Soc. 2021, 168, 034504. [Google Scholar] [CrossRef]
  12. Amin, H.M.A.; Zan, L.; Baltruschat, H. Boosting the bifunctional catalytic activity of Co3O4 on silver and nickel substrates for the alkaline oxygen evolution and reduction reactions. Surf. Interfaces 2024, 54, 105218. [Google Scholar] [CrossRef]
  13. Amin, H.M.A.; Baltruschat, H.; Wittmaier, D.; Friedrich, K.A. A highly efficient bifunctional catalyst for alkaline air-electrodes based on a Ag and Co3O4 hybrid: RRDE and online DEMS insights. Electrochim. Acta 2015, 151, 332–339. [Google Scholar] [CrossRef]
  14. Wei, Y.; Hu, Y.; Da, P.; Weng, Z.; Xi, P.; Yan, C.-H. Triggered lattice-oxygen oxidation with active-site generation and self-termination of surface reconstruction during water oxidation. Proc. Natl. Acad. Sci. USA 2023, 120, e2312224120. [Google Scholar] [CrossRef] [PubMed]
  15. Hua, B.; Li, M.; Pang, W.; Tang, W.; Zhao, S.; Jin, Z.; Zeng, Y.; Amirkhiz, B.S.; Luo, J.-L. Activating p-blocking centers in perovskite for efficient water splitting. Chem 2018, 4, 2902–2916. [Google Scholar] [CrossRef]
  16. Mefford, J.T.; Rong, X.; Abakumov, A.M.; Hardin, W.G.; Dai, S.; Kolpak, A.M.; Johnston, K.P.; Stevenson, K.J. Water electrolysis on La1-xSrxCoO3−δ perovskite electrocatalysts. Nat. Commun. 2016, 7, 11053. [Google Scholar] [CrossRef]
  17. Zhu, Y.; Zhou, W.; Sunarso, J.; Zhong, Y.; Shao, Z. Phosphorus-doped perovskite oxide as highly efficient water oxidation electrocatalyst in alkaline solution. Adv. Funct. Mater. 2016, 26, 5862–5872. [Google Scholar] [CrossRef]
  18. Luo, Q.; Lin, D.; Zhan, W.; Zhang, W.; Tang, L.; Luo, J.; Gao, Z.; Jiang, P.; Wang, M.; Hao, L.; et al. Hexagonal perovskite Ba0.9Sr0.1Co0.8Fe0.1Ir0.1O3−δ as an efficient electrocatalyst towards the oxygen evolution reaction. ACS Appl. Energy Mater. 2020, 3, 7149–7158. [Google Scholar] [CrossRef]
  19. Cai, W.; Zhou, M.; Cao, D.; Yan, X.; Li, Q.; Lü, S.; Mao, C.; Li, Y.; Xie, Y.; Zhao, C.; et al. Ni-doped A-site-deficient La0.7Sr0.3Cr0.5Mn0.5O3−δ perovskite as anode of direct carbon solid oxide fuel cells. Int. J. Hydrogen Energy 2020, 45, 21873–21880. [Google Scholar] [CrossRef]
  20. Han, Y.; Zhu, Z.; Huang, L.; Guo, Y.; Zhai, Y.; Dong, S. Hydrothermal synthesis of polydopamine-functionalized cobalt-doped lanthanum nickelate perovskite nanorods for efficient water oxidation in alkaline solution. Nanoscale 2019, 11, 19579–19585. [Google Scholar] [CrossRef]
  21. Sung, M.-C.; Kim, C.H.; Hwang, B.; Kim, D.-W. Rationalizing the catalytic surface area of oxygen vacancy-enriched layered perovskite LaSrCrO4 nanowires on oxygen electrocatalyst for enhanced performance of LiO2 batteries. Carbon Energy 2024, 6, e550. [Google Scholar] [CrossRef]
  22. Cao, C.; Shang, C.; Li, X.; Wang, Y.; Liu, C.; Wang, X.; Zhou, S.; Zeng, J. Dimensionality control of electrocatalytic activity in perovskite nickelates. Nano Lett. 2020, 20, 2837–2842. [Google Scholar] [CrossRef] [PubMed]
  23. Niu, Y.; Chang, X.; Zhang, M.; Mu, J. Surface reconstruction of La0.6Sr0.4Co0.8Ni0.2O3−δ perovskite nanofibers for oxygen evolution reaction. Ceram. Int. 2024, 50, 13014–13021. [Google Scholar] [CrossRef]
  24. Roy, S.; Yoshida, T.; Kumar, A.; Yusuf, S.M.; Chakraborty, C.; Roy, S. Tailoring Co site reactivity via Sr and Ni doping in LaCoO3 for enhanced water splitting performance. Catal. Today 2024, 441, 114885. [Google Scholar] [CrossRef]
  25. Rendón-Angeles, J.C.; Yanagisawa, K.; Matamoros-Veloza, Z.; Pech-Canul, M.I.; Mendez-Nonell, J.; la Torre, S.D. Hydrothermal synthesis of perovskite strontium doped lanthanum chromite fine powders and its sintering. J. Alloys Compd. 2010, 504, 251–256. [Google Scholar] [CrossRef]
  26. Thanh, T.D.; Chuong, N.D.; Balamurugan, J.; Van Hien, H.; Kim, N.H.; Lee, J.H. Porous hollow-structured LaNiO3 stabilized N, S-Codoped graphene as an active electrocatalyst for oxygen reduction reaction. Small 2017, 13, 1701884. [Google Scholar] [CrossRef]
  27. Omari, E.; Makhloufi, S.; Omari, M. Preparation by sol-gel method and characterization of Co-doped LaNiO3 perovskite. J. Inorg. Organomet. Polym. 2017, 27, 1466–1472. [Google Scholar]
  28. Liang, D.; Huang, H.; Liu, J.; Wang, H. Preparation and modification of La0.4Sr0.6Co1-xNixO3 (x = 0-0.8) perovskite oxide for bi-functional catalysis in alkaline medium. Inorg. Chem. Commun. 2021, 127, 108533. [Google Scholar] [CrossRef]
  29. Ramana, C.V.; Vemuri, R.S.; Kaichev, V.V.; Kochubey, V.A.; Saraev, A.A.; Atuchin, V.V. X-ray photoelectron spectroscopy depth profiling of La2O3/Si thin films deposited by reactive magnetron sputtering. ACS Appl. Mater. Interfaces 2011, 3, 4370–4373. [Google Scholar] [CrossRef]
  30. Komai, S.; Hirano, M.; Ohtsu, N. Spectral analysis of Sr 3d XPS spectrum in Sr-containing hydroxyapatite. Surf. Interface Anal. 2020, 52, 823–828. [Google Scholar] [CrossRef]
  31. Yang, J.; Liu, H.; Martens, W.N.; Frost, R.L. Synthesis and characterization of cobalt hydroxide, cobalt oxyhydroxide, and cobalt oxide nanodiscs. J. Phys. Chem. C 2010, 114, 111–119. [Google Scholar] [CrossRef]
  32. Zhao, Q.; Fang, C.; Tie, F.; Luo, W.; Peng, Y.; Huang, F.; Ku, Z.; Cheng, Y.-B. Regulating the Ni3+/Ni2+ ratio of NiOx by plasma treatment for fully vacuum-deposited perovskite solar cells. Mater. Sci. Semicond. Process. 2022, 148, 106839. [Google Scholar] [CrossRef]
  33. Wang, J.; Mueller, D.N.; Crumlin, E.J. Recommended strategies for quantifying oxygen vacancies with X-ray photoelectron spectroscopy. J. Eur. Ceram. Soc. 2024, 44, 116709. [Google Scholar] [CrossRef]
  34. Amin, H.M.A.; Bondue, C.J.; Eswara, S.; Kaiser, U.; Baltruschat, H. A carbon-free Ag-Co3O4 composite as a bifunctional catalyst for oxygen reduction and evolution: Spectroscopic, microscopic and electrochemical characterization. Electrocatalysis 2017, 8, 540–553. [Google Scholar] [CrossRef]
  35. Hu, Z.; Yan, Q.; Wang, Y. Dynamic surface reconstruction of perovskite oxides in oxygen evolution reaction and its impacts on catalysis: A critical review. Mater. Today Chem. 2023, 34, 101800. [Google Scholar] [CrossRef]
  36. Dong, F.; Li, L.; Kong, Z.; Xu, X.; Zhang, Y.; Gao, Z.; Dongyang, B.; Ni, M.; Liu, Q.; Lin, Z. Materials engineering in perovskite for optimized oxygen evolution electrocatalysis in alkaline condition. Small 2020, 17, 2006638. [Google Scholar] [CrossRef]
  37. Chen, L.; Yan, C.; Zhang, F.; Chi, J.; Xiong, W.; Liu, P.; Hao, F. Modulated oxygen mobility of perovskite oxides for efficient redox oxidative cracking of cycloalkane to light olefins. Chem. Eng. J. 2023, 477, 146894. [Google Scholar] [CrossRef]
  38. Wang, Y.; Wu, J.; Lu, X.; Guo, Y.; Zhao, H.; Tang, X. A-site doped ruthenium perovskite bifunctional electrocatalysts with high OER and ORR activity. J. Alloys Compd. 2022, 920, 165770. [Google Scholar] [CrossRef]
  39. 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]
  40. Cao, J.; Riaz, S.; Qi, Z.; Zhao, K.; Qi, Y.; Wei, P.; Xie, Y. Superhydrophilic self-supported perovskite oxides for oxygen evolution reactions in oilfield wastewater. Catal. Lett. 2024, 154, 5350–5358. [Google Scholar] [CrossRef]
Figure 1. XRD images of La0.5Sr0.5Co1−xNixO3−δ-S by the sol-gel method.
Figure 1. XRD images of La0.5Sr0.5Co1−xNixO3−δ-S by the sol-gel method.
Nanomaterials 15 00210 g001
Figure 2. SEM images of (a) La0.5Sr0.5Co0.8Ni0.2O3−δ-S, (b,c) La0.5Sr0.5Co0.2Ni0.8O3−δ-S, and (d) La0.5Sr0.5CoO3−δ-S by the sol-gel method.
Figure 2. SEM images of (a) La0.5Sr0.5Co0.8Ni0.2O3−δ-S, (b,c) La0.5Sr0.5Co0.2Ni0.8O3−δ-S, and (d) La0.5Sr0.5CoO3−δ-S by the sol-gel method.
Nanomaterials 15 00210 g002
Figure 3. EDS mapping of La0.5Sr0.5Co0.8Ni0.2O3−δ-S by the sol-gel method at a 10 μm scale by SEM (a) SEM image of La0.5Sr0.5Co0.8Ni0.2O3−δ-S, (b) La, (c) Sr, (d) O, (e) Co, (f) Ni.
Figure 3. EDS mapping of La0.5Sr0.5Co0.8Ni0.2O3−δ-S by the sol-gel method at a 10 μm scale by SEM (a) SEM image of La0.5Sr0.5Co0.8Ni0.2O3−δ-S, (b) La, (c) Sr, (d) O, (e) Co, (f) Ni.
Nanomaterials 15 00210 g003
Figure 4. SEM images of (a) La0.5Sr0.5CoO3−δ-H, (b) La0.5Sr0.5Co0.8Ni0.2O3−δ-H, (c) La0.5Sr0.5Co0.5Ni0.5O3−δ-H, and (d) La0.5Sr0.5Co0.2Ni0.8O3−δ-H by the hydrothermal method.
Figure 4. SEM images of (a) La0.5Sr0.5CoO3−δ-H, (b) La0.5Sr0.5Co0.8Ni0.2O3−δ-H, (c) La0.5Sr0.5Co0.5Ni0.5O3−δ-H, and (d) La0.5Sr0.5Co0.2Ni0.8O3−δ-H by the hydrothermal method.
Nanomaterials 15 00210 g004
Figure 5. EDS mapping of La0.5Sr0.5Co0.8Ni0.2O3−δ-H by the hydrothermal method at a 2 μm scale by SEM (a) SEM image of La0.5Sr0.5Co0.8Ni0.2O3−δ-H, (b) La, (c) Sr, (d) O, (e) Co, (f) Ni.
Figure 5. EDS mapping of La0.5Sr0.5Co0.8Ni0.2O3−δ-H by the hydrothermal method at a 2 μm scale by SEM (a) SEM image of La0.5Sr0.5Co0.8Ni0.2O3−δ-H, (b) La, (c) Sr, (d) O, (e) Co, (f) Ni.
Nanomaterials 15 00210 g005
Figure 6. XPS spectra of La0.5Sr0.5Co0.8Ni0.2O3−δ-S by the sol-gel method. (a) La 3d, (b) Sr 3d, (c) Co 2p, (d) Ni 2p, and (e) O 1s.
Figure 6. XPS spectra of La0.5Sr0.5Co0.8Ni0.2O3−δ-S by the sol-gel method. (a) La 3d, (b) Sr 3d, (c) Co 2p, (d) Ni 2p, and (e) O 1s.
Nanomaterials 15 00210 g006
Figure 7. OER tests of La0.5Sr0.5Co1−xNixO3−δ-S (a) LSV curves, (b) Tafel curves, (c) EIS, (d) comparison of OER overpotentials of the catalysts in 1.0 M KOH solution at 10 mA cm−2 [7,8,17,20,23,24,38,39,40].
Figure 7. OER tests of La0.5Sr0.5Co1−xNixO3−δ-S (a) LSV curves, (b) Tafel curves, (c) EIS, (d) comparison of OER overpotentials of the catalysts in 1.0 M KOH solution at 10 mA cm−2 [7,8,17,20,23,24,38,39,40].
Nanomaterials 15 00210 g007
Figure 8. OER tests of La0.5Sr0.5Co1−xNixO3−δ-H. (a) LSV curves, (b) EIS.
Figure 8. OER tests of La0.5Sr0.5Co1−xNixO3−δ-H. (a) LSV curves, (b) EIS.
Nanomaterials 15 00210 g008
Figure 9. (a) CV curves of La0.5Sr0.5Co0.8Ni0.2O3−δ-S, (b) capacitive currents.
Figure 9. (a) CV curves of La0.5Sr0.5Co0.8Ni0.2O3−δ-S, (b) capacitive currents.
Nanomaterials 15 00210 g009
Figure 10. The stability test of La0.5Sr0.5Co0.8Ni0.2O3−δ by the sol-gel method.
Figure 10. The stability test of La0.5Sr0.5Co0.8Ni0.2O3−δ by the sol-gel method.
Nanomaterials 15 00210 g010
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

Zong, B.; Pan, X.; Zhang, L.; Wei, B.; Feng, X.; Guo, M.; Cao, D.; Ye, F. Preparation and Performance of Nickel-Doped LaSrCoO3-SrCO3 Composite Materials for Alkaline Oxygen Evolution in Water Splitting. Nanomaterials 2025, 15, 210. https://doi.org/10.3390/nano15030210

AMA Style

Zong B, Pan X, Zhang L, Wei B, Feng X, Guo M, Cao D, Ye F. Preparation and Performance of Nickel-Doped LaSrCoO3-SrCO3 Composite Materials for Alkaline Oxygen Evolution in Water Splitting. Nanomaterials. 2025; 15(3):210. https://doi.org/10.3390/nano15030210

Chicago/Turabian Style

Zong, Bangfeng, Xiaojun Pan, Lifang Zhang, Bo Wei, Xiangxiong Feng, Miao Guo, Duanhao Cao, and Feng Ye. 2025. "Preparation and Performance of Nickel-Doped LaSrCoO3-SrCO3 Composite Materials for Alkaline Oxygen Evolution in Water Splitting" Nanomaterials 15, no. 3: 210. https://doi.org/10.3390/nano15030210

APA Style

Zong, B., Pan, X., Zhang, L., Wei, B., Feng, X., Guo, M., Cao, D., & Ye, F. (2025). Preparation and Performance of Nickel-Doped LaSrCoO3-SrCO3 Composite Materials for Alkaline Oxygen Evolution in Water Splitting. Nanomaterials, 15(3), 210. https://doi.org/10.3390/nano15030210

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