Next Article in Journal
Research on the Terahertz Modulation Performance of VO2 Thin Films with Surface Plasmon Polaritons Structure
Previous Article in Journal
A Review of Crack Sealing Technologies for Asphalt Pavement: Materials, Failure Mechanisms, and Detection Methods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of the Structure and Electrochemical Performance of Perovskite Oxide La1−xCaxCrO3 Utilized as Electrode Materials for Supercapacitors

School of Material Science and Engineering, Jiangsu University of Science and Technology, Zhenjiang 212003, China
*
Authors to whom correspondence should be addressed.
Coatings 2025, 15(7), 837; https://doi.org/10.3390/coatings15070837
Submission received: 4 June 2025 / Revised: 7 July 2025 / Accepted: 15 July 2025 / Published: 17 July 2025

Abstract

Lanthanide perovskite materials are promising candidates for supercapacitor applications. In this study, a series of La1−xCaxCrO3 (x = 0–0.2) materials were prepared by sol-gel method, incorporating bivalent ions calcium at A-site. La0.85Ca0.15CrO3 exhibited the lowest charge transfer resistance and highest specific surface area. At 1 A/g, La0.85Ca0.15CrO3 achieved a maximum specific capacitance of 306 F/g, about 2.3 times higher than that of the LaCrO3 (133 F/g). Based on the observed data, a mechanism involving oxygen anion charge storage during the charging-discharging process is proposed. After 5000 long cycle, the coulomb efficiency of the electrode remains above 94%. These results demonstrate that Ca-substituted compounds exhibit significant potential for A-site engineering in supercapacitor applications.

1. Introduction

In the contemporary era, the environmental energy crisis has emerged as a pivotal global concern [1,2,3]. Rapid population growth and industrialization have driven escalating global energy demands. This has led to a significant increase in the dependence on fossil fuels, triggering serious environmental degradation [4,5,6]. Nations pursue sustainable energy solutions like enhancing energy efficiency and developing renewable energy sources to balance economic growth with environmental protection [7,8]. Supercapacitors mitigate energy storage challenges with advantages like ultrafast charge and discharge, long cycle stability, and high-power density [9,10,11]. Unlike conventional batteries, supercapacitors are particularly well-suited for energy storage applications for intermittent energy sources, such as wind and solar energy [12]. Electrode material selection significantly impacts on supercapacitor performance. Consequently, a number of studies have investigated the potential of efficient electrode materials with excellent specific capacitance and long-term cycle stability. In recent years, lanthanide perovskite materials have attracted considerable attention in supercapacitors due to a number of advantageous properties, including high degrees of crystallinity, reversible redox capabilities, outstanding electrical characteristics, a profusion of oxygen vacancies, a highly stable architectural framework, straightforward synthesis procedures, and cost-efficiency [13,14].
ABO3-type perovskites have been studied as supercapacitors electrode materials, including LaCoO3 [15,16,17], LaMnO3 and LaCrO3 [18]. Among these materials, LaCrO3 exhibits the highest specific surface area and electrochemical properties, making it a prime research focus for supercapacitor research [19]. Nonetheless, LaCrO3’s ion diffusion and electron transfer speed are limited, restricting its practical use as an anode material [20]. LaCrO3 represents a prototypical ABO3 perovskite oxide structure. The basic building block of the structure is the CrO6 octahedron. The physical characteristics of the crystal structure depend on the deformation and configuration of these CrO6 octahedra [21]. Octahedron stability can be adjusted by doping the La site, thereby altering the performance of LaCrO3. LaCrO3 perovskite exhibits intrinsic ion vacancies that enhance their catalytic and electrical properties [22]. As a result, lanthanum chromite (LaCrO3) is widely utilized in fuel cell electrodes, catalytic substances, and electromagnetic apparatuses. In addition, chromium-containing materials hold considerable promise for energy storage uses due to their protons/electrons conductivity, multiple oxidation states (specifically Cr3+/Cr6+), high redox capacity, outstanding thermal and structural stability, and natural abundance [23]. Perovskite-type oxides can be modified to achieve desired performance. This tunability facilitates the development of specific supercapacitor applications. Oxygen vacancies play a crucial role in optimizing the specific capacitance. Effective strategies for generating oxygen vacancies include doping the ABO3 structure with A-site or B-site impurities to alter the oxidation state of the transition metal, or introducing oxygen sites for charge balance [24,25]. Transition metal dopants’ precise modulation of A/B-site cations significantly enhance the perovskite oxide electrodes’ capacitance, cycling stability, and the overall electrochemical characteristics. Doping with Ca, Sr, Co, and K at the A-site of LaMnO3 [26,27,28,29], LaAlO3 [30], LaFeO3 [31], LaNiO3 [32], and LaCoO3 [33], as well as Zn, Fe, and Ni at the B-site, has demonstrated significant performance enhancements. Partial A-site substitution with lower-cost elements results in B-site transition metal oxidation state instability and generates more oxygen vacancies. This modification enhances both electrical conductivity and electrochemical properties. The specific capacitances of LaMnO3, LaAlO3, LaFeO3 and LaNiO3 are 72, 122, 17.15 and 48.85 F/g, respectively. It is noteworthy that LaCrO3 exhibits a relatively high specific capacitance of 86.4 F/g [34]. To further enhance the specific capacitance of LaCrO3, researchers investigated the doping of Ca at the A-site as a potential avenue. For example, following the testing of the electrochemical properties of LaCrO3 and La0.7Ca0.3CrO3, Jiang et al. [35] demonstrated that the La0.7Ca0.3CrO3’s electrode polarization resistance decreased from 45.3 Ω cm2 to 6.3 Ω cm2 compared to undoped LaCrO3. La0.7Ca0.3CrO3 exhibits significantly lower electrode polarization resistance for oxygen reduction, enhancing the electrocatalytic activity of LaCrO3 for fuel cell reaction. Rashtchi et al. [36] prepared an AISI 430 interconnect metal coating by means of the sol-gel method and investigated the effect of doped calcium on the conductivity and oxidation properties of the coating. Their results show that the oxidation rate and antioxidant properties of the La0.8Ca0.2CrO3 coating are significantly reduced by about 2 times and 6 times than LaCrO3 metal coating. Their findings indicated that strontium doping improves the electrochemical properties of LaCrO3 materials. Consequently, it is crucial to examine the impact of Ca doping on the electrochemical characteristics of LaCrO3 materials.
This study introduces several innovations. Firstly, it systematically explores the effects of different calcium doping ratios (x = 0–0.2) on the microstructure, morphology and electrochemical performance of La1−xCaxCrO3. Previous studies have mostly focused on a single doping ratio. The present study combines XRD, SEM, TEM and XPS analysis, supplemented by electrochemical testing and EIS analysis, to provide new theoretical and experimental foundations for related applications. The results demonstrate that La0.85Ca0.15CrO3 exhibits the optimal electrochemical performance, characterized by a specific capacitance of 306 F/g. Despite the enhancement of the electrochemical performance of LaCrO3 material through the doping of Ca elements, the performance difference between Ca2+ and La3+ can readily result in uneven element distribution and composition segregation, thereby affecting the uniformity of performance. Future research endeavors could explore avenues for expanding the application prospects of the process by optimizing process parameters, developing new chelating agents to achieve uniform doping, and investigating green preparation routes to reduce costs.

2. Experimental Section

The requisite pharmaceuticals for the preparation of La1−xCaxCrO3 (x = 0–0.2) series samples by sol-gel methodology are as follows: The requisite chemicals include citric acid monohydrate (≥99.5%), lanthanum nitrate hexahydrate (La(NO3)3·6H2O), chromium nitrate hydrate (≥99.9%) and Cr(NO3)2·9H2O (≥99%), as well as calcium nitrate Ca(NO3)2 (≥99.5%). The aforementioned reagents were all procured from Aladin Chemical Reagent Co., Ltd, Shanghai, China. These chemicals are of analytical grade and can be used directly without further purification.
Citric acid monohydrate, La(NO3)3·6H2O and Cr(NO3)2·9H2O were added in succession to a 30 ml beaker at a molar ratio of 1:1:1. A series of La1−xCaxCrO3 (x = 0–0.2) were prepared by altering the molar ratio of Ca(NO3)2. The solution should then be stirred for 24 h on a magnetic stirrer in order to form a transparent solution, resulting in the formation of a transparent solution. Following this, the solution underwent a drying procedure in a forced-air drying oven at a temperature of 80 °C for a duration of 72 h. Thereafter, the dried sample was removed and placed in a crucible and calcined in a muffle furnace. The Muffle furnace(KSL-1200, HF-Kejing, Hefei, China) was heated to 700 °C at a rate of 10 °C/min, maintained in an air atmosphere for a period of 2 h, and subsequently ground to produce a black powder sample. The synthesized black powder specimens underwent X-ray powder diffraction (XRD) examination. X-ray powder diffractometer (Bruker D8, Karlsruhe, Germany) was employed to conduct a phase analysis of these specimens. The X-ray powder diffractometer has a scanning range of 20–70° and a scanning speed of 6°/min. The microstructure of the samples was characterized by field emission scanning electron microscopy (Merlin Compact, Carl Zeiss, Munich, Germany). The elemental composition and chemical valence changes of the samples were analyzed by means of X-ray photoelectron spectroscopy (XPS, Thermofisher, Waltham, MA, USA). A specific surface area test was conducted using a specific surface area tester (BET, V-Sorb 2800TP, Hefei, China).
In order to conduct electrochemical testing of the prepared electrodes, a three-electrode system was employed, utilizing 3 M KOH as the electrolyte. The three electrodes are the working electrode, the counter electrode (platinum electrode), and the reference electrode (Hg/HgO electrode, Yueci, Shanghai, China). The electrode is prepared in the following manner: Firstly, the sample, comprising polyvinylidene fluoride (PVDF, Aladin Chemical Reagent Co., Ltd., Shanghai, China) and acetylene black, is mixed at a ratio of 8:1:1 in a specified quantity of N-methylpyrrolidone (NMP, Aladin Chemical Reagent Co., Ltd., Shanghai, China) and stirred for a designated period of time. Subsequently, the resulting paste is uniformly distributed on the 20 mm × 10 mm nickel foam sheet, with a coating area of 100 mm2. The mass of the active substance coated on the nickel foam is approximately 0.8 mg. After subjecting the electrode to a vacuum oven at 80 °C for 24 h, the nickel foam is ready for use. Prior to this, however, the foam nickel be deoiled with acetone and deoxidized with diluted hydrochloric acid. The electrochemical properties of the electrodes were investigated using an Electrochemistry Workstation (CHI660e, Chenhua, Shanghai, China), which allowed for the determination of cyclic voltammetry (CV), galvanostatic charge-discharge (GCD), and alternating-current impedance (EIS). The stability of the electrodes during cycling was assessed with a multi-channel battery testing apparatus (Land, CT2001A, Wuhan, China).

3. Results and Discussion

Figure 1a depicts the XRD patterns of the La1−xCaxCrO3 (x = 0–0.2), which are compared to the standard card PDF#75–0441. It reveals the presence of peaks corresponding to pure LaCrO3. No impurity peaks were detected, indicating that Ca2+ has been completely dissolved into the LaCrO3 lattice. These results demonstrated the successful synthesis of the La1−xCaxCrO3 sample. Diffraction peaks observed at 2θ = 22.90°, 32.61°, 40.22°, 46.78°, 52.70°, 58.19°, and 68.32°. The diffraction peaks point to the (100), (110), (111), (200), (210), (211) and (220) crystal faces, respectively. As illustrated in Figure 1b, the diffraction peak of the locally enlarged (110) crystal plane demonstrates a shift towards higher angles with Ca2+ doping. The angular shift is more pronounced than that observed in pure LaCrO3. This shift originates from doped Ca2+ occupying the A-site to minimize lattice stress. As Ca doping level increases, smaller Ca ion (0.99 A) replaces La ion (1.06 Å), inducing some Cr3+ to Cr6+ to maintain electrical neutrality. The ionic radius of Cr6+ (0.52 Å) is smaller than that of Cr3+ (0.69 Å). According to Bragg’s formula, increasing Ca doping shifts the diffraction peak to a larger angle. Calcium introduction affects ion transport pathways primarily through the lattice parameter alterations, defect formation, and microstructure change. Lattice expansion alters the particle size and pore structure, which in turn affects the ion transport efficiency. Ca doping in LaCrO3 enhances redox activity and reaction rate by changing electronic structure, increasing active sites, and reducing charge transfer resistance [37]. As illustrated in Figure 1c, the XPS full spectrum of La1−xCaxCrO3 (x = 0–0.2) samples demonstrates the presence of La, Cr, O and Ca. The outcomes of XRD and XPS validate that Ca has been successfully integrated into the samples.
As illustrated in the SEM diagram of Figure 2, La1−xCaxCrO3 (x = 0–0.2) sample exhibits a microstructure comprising the aggregation of nanoparticles. This structure possesses a high specific surface area and small grain size, which facilitate enhanced surface activity and reaction rate [38]. Homogeneous particle distribution ensures consistent material properties and structural stability. Grain size decreases with elevated calcium doping levels correlates with lattice distortion induced by Ca2⁺ incorporation, which inhibits crystal growth [39]. The increased lattice defects are the primary factor reducing grain size. These defects result in the incompleteness and distorted crystal structure, thereby influencing the crystal growth process. During crystal growth, lattice defects may impede grain growth, ultimately reducing the overall grain size [40,41]. Grain size reduction increases grain boundary density, hindering ion diffusion and stabilizing the electrolyte interface. The increased surface-area-to-volume ratio in smaller grains mitigates material degradation. Calcium doping relieves internal stresses and enhances mechanical stability. Reduced grains shorten ion diffusion paths, while Ca doping improves both ionic and electronic conductivity. With more active sites, rate capability was enhanced at high current densities [42].
To investigate the impact of Ca doping on the specific surface area of La1−xCaxCrO3 (x = 0–0.2) series samples, the N2 adsorption-desorption method was employed. The particular surface area and the distribution of pore sizes of LaCrO3 and La0.85Ca0.15CrO3 nanoparticles were determined. Figure 3a,c illustrate the N2 adsorption-desorption isotherm curves of LaCrO3 and La0.85Ca0.15CrO3 samples. The results demonstrate that with an increase in relative pressure, the adsorption capacity continues to rise. Above a relative pressure of 0.8, adsorption capacity increases sharply, exhibiting an evident hysteresis ring [43]. The observed hysteresis confirms mesoporous architectures in both samples [44]. Figure 3b,d show both samples have pore sizes predominantly in the range of 1–10 nm, confirming their mesoporous structure. Mesoporous structures enhance ion diffusion kinetics and surface redox reactions in electrodes. Interconnected pores serve as rapid ion pathways, reducing ion transport distances. Additionally, the high specific surface area and porosity of mesoporous structures improve electrolyte infiltration and ion transmission efficiency, while exposing more active sites for redox reactions [45]. Enhanced accessibility and chemical activity of the mesoporous surfaces promote redox reactions [46]. Specific surface area measurements demonstrate that the specific surface area of the LaCrO3 and La0.85Ca0.15CrO3 samples are 38.01 and 48.21 m2g−1, respectively. These results suggest that the Ca doping enhances the specific surface area of the LaCrO3 system.
Figure 4a–d depict La 3d, Cr 2p, O 1s, and Ca 3d high-resolution XPS spectra of La1−xCaxCrO3 (x = 0–0.2). The high-resolution La 3d XPS spectrum is presented in Figure 4a. Peaks at 856.0 and 851.0 eV attribute to the orbital peak of La 3d3/2, while peaks at 839.0 and 834.0 eV correspond to the orbital peak of La 3d5/2 [47]. The difference between the two La 3d peaks is approximately 17 eV, indicating that the La element exists in a +3 valence state in the La1−xCaxCrO3 (x = 0–0.2) nanoparticles [48]. Figure 4b depicts the Cr 2p high-resolution XPS diagram. The spectrum exhibits two distinct peaks. The peak with the lower binding energy is identified as Cr 2p1/2, while the other is Cr 2p3/2. The Cr 2p1/2 and Cr 2p3/2 peaks are composed of two smaller peaks [49]. The Cr 2p1/2 peak occurs at 586.6 eV, while the Cr 2p3/2 peak occurs at 576.7 eV. Similarly, the Cr 2p1/2 peak occurs at 589.8 eV, while the Cr 2p3/2 peak occurs at 579.9 eV. In the Cr 2p1/2 and Cr 2p3/2 spectra, the peaks with lower binding energies are attributed to Cr3+, while those with higher binding energies are ascribed to Cr6+. This suggests that chromium ions in LaCrO3 exhibit two distinct valence states [50]. XPS analysis shows that changes in oxygen vacancy concentration and Cr oxidation state in La1−xCaxCrO3 are directly linked to its charge storage mechanism in supercapacitors [51]. Oxygen vacancies can alter the material’s electronic structure, enhancing conductivity and ion diffusion pathways, thereby improving charge storage efficiency. Meanwhile, changes in Cr oxidation state directly impact electrochemical activity by providing additional redox pairs, increasing specific capacitance and pseudocapacitive properties [52]. Figure 4c illustrates the sub-peak fitting outcomes for the O 1s characteristic peaks of La1−xCaxCrO3 (x = 0–0.2) nanoparticles. Three distinctive O 1s peaks are observed at 532.9, 531.5 and 529.4 eV, respectively [53]. The peak observed at 532.9 eV is attributed to the presence of an oxygen-containing functional group, specifically a hydroxyl group (OH). The characteristic peaks observed at 531.5 eV are attributed to the presence of adsorbed oxygen, which may be in the forms of O, O2−, or O22−. The peak at 529.4 eV is attributed to lattice oxygen (O2−) [53]. Table 1 presents the relative concentrations of the three types of oxygen present in the La1−xCaxCrO3 (x = 0–0.2) sample. The oxygen vacancy of the electrode material is intimately associated with its electrical conductivity. High adsorbed oxygen (O, O2−, O22−) concentration enhances OH adsorption, surface redox reactions, and electrochemical performance [54]. The concentration of oxygen vacancies is determined by calculating the ratio of the amount of adsorbed oxygen to the amount of lattice oxygen present. Table 1 illustrates that the introduction of Ca into the La1−xCaxCrO3 (x = 0–0.2) series increases oxygen vacancy concentration. Among these samples, La0.85Ca0.15CrO3 exhibits the highest oxygen vacancy concentration. Compared with LaCrO3, Ca doping reduce the oxygen vacancy formation energy in La0.75Ca0.25CrO3, as the charge of Ca replacing La is opposite to that of oxygen vacancies. Oxygen sites are inequivalent, making O sites easier to form oxygen vacancies. Under tensile strain, oxygen vacancy formation energy decreases without obvious minimum value; Under compressive strain, the O 1/O 2 formation energy decreases and the O 3 formation energy increases [55]. As illustrated in Figure 4d, the high-resolution XPS diagram of the La1−xCaxCrO3 (x = 0–0.2) nanoparticle sample Ca 2p demonstrates a characteristic peak at 350.5eV, corresponding to the orbital peak of Ca 2p1/2. The presence of a characteristic peak at 347.1 eV is indicative of the Ca 2p3/2 orbital peak. The results obtained thus far demonstrate that the Ca elements present in the La1−xCaxCrO3 (x = 0–0.2) nanoparticle samples exist at + 2 valence [56].
Figure 5a–d illustrates the cyclic voltammetry curves of the La1−xCaxCrO3 (x = 0–0.2) electrodes at a scanning rate of 10–100 mV/s. The electrodes exhibit excellent reversibility, which is evidenced by the unaltered shape of the cyclic voltammetry curve despite increasing current density. A distinct redox peak suggests that redox reactions contribute to the charge storage mechanism [57]. Figure 5e illustrates the CV curve of nickel foam. The active substance is carried on the nickel foam, which may affect electrochemical tests. As demonstrated in Figure 5e, the nickel foam CV curve area is negligible at 10–100 mV/s. This observation indicates that nickel foam exhibits no inherent capacity to facilitate charge or material transfer. Consequently, the earlier electrochemical performance enhancement is exclusively attributable to La1−xCaxCrO3. As the Ca doping concentration increased, the enclosed area of the cyclic voltammetry curves for the La1−xCaxCrO3 series electrodes (x = 0–0.2) exhibited a trend of initial increase followed by subsequent decrease. Ca doping has an impact on the electrochemical characteristics of electrode materials. La0.85Ca0.15CrO3 electrode displays the most favorable electrochemical performance, as evidenced by the largest enclosed cyclic voltammetry curve area. La0.85Ca0.15CrO3 electrode exhibits enhanced capacitance and superior cycle stability during capacitor discharge and charging [58]. This phenomenon can be accounted for by referring to the oxygen insertion process of perovskite oxide substances. The equations for the reaction are as follows [59]:
L a C r 2 δ 3 + ; C r 1 2 δ 6 + O 3 δ + 2 δ O H L a C r 3 + O 3 + δ H 2 O + 2 δ e
L a C r 3 + O 3 + 2 δ O H L a C r 2 δ 6 + ; C r 1 2 δ 3 + O 3 + δ + δ H 2 O + 2 δ e
L a 0.85 C a 0.15 C r 2 δ 3 + ; C r 1 2 δ 3 + O 2.925 δ + 2 δ O H L a 0.85 C a 0.15 C r 3 + O 2.925 + δ H 2 O + 2 δ e
L a 0.85 C a 0.15 C r 3 + O 2.925 + 2 δ O H L a 0.85 C a 0.15 C r 2 δ 6 + ; C r 1 2 δ 3 + O 2.925 + δ + δ H 2 O + 2 δ e
E R H E = E H g / H g O + 0.0592 p H + 0.098
According to Equation (5), the voltage window of the working electrode during measurement is determined based on the electrode of the reference electrode. The CV and XPS results show that the electrochemical storage mechanism of LaCrO3 perovskite in alkaline electrolytes involves oxygen ion intercalation mediated by Cr oxidation. This can be expressed by Equations (1) and (2). Oxygen ions adsorbed on the surface of perovskite release a proton (H+) to insert into the oxygen vacancies. Subsequently, the proton (H+) combines with another oxygen ion (OH) to form H2O. The oxygen ion diffusion along the octahedron edge is responsible for the filling of oxygen vacancies. This process leads to the intercalation of each oxygen ion, thereby oxidizing Cr3+ to Cr6+. As illustrated by Equations (3) and (4), the excess oxygen ions diffuse to the surface via the cobalt centers, concurrently oxidizing the Cr3+ atoms to the Cr6+ state. These results confirm that the LaCrO3 electrodes contains a greater number of oxygen vacancies, aligning with prior XPS findings. With increasing in Ca doping content, as illustrated in Table 1, oxygen vacancy concentration within the crystal structure rises correspondingly. Enhanced oxygen ion conductivity of the electrode material improves its electrochemical performance. Nevertheless, excessive Ca doping alters lattice parameter, leading to crystal structure distortion and defect site formation [60]. Such alterations degrade electrochemical performance, including a reduction in cycle stability and energy storage efficiency in capacitors [61,62,63].
Figure 6a–d illustrate the galvanostatic charge-discharge curve of the La1−xCaxCrO3 (x = 0–0.2) electrodes sample at current densities of 0.5–5 A/g. As the current density rises, the duration of discharge diminishes. This trend is due to the increased internal resistance at higher current densities, leading to a decrease in the specific capacitance [64]. At a constant current density, the discharge duration the La1−xCaxCrO3 (x = 0–0.2) electrodes first increases then decreases with increasing Ca doping concentration. The specific capacitance of the La1−xCaxCrO3 (x = 0–0.2) electrodes sample at a current density of 0.5 A/g is 133, 211, 306 and 211 F/g, respectively. The specific capacitance shows a volcano-shaped dependence on Ca content, peaking at x = 0.15. The electrode composed of La0.85Ca0.15CrO3 demonstrates the greatest specific capacitance, and this result is consistent with the outcomes obtained from the cyclic voltammetry curve test [65].
The electrode performance contingents upon the redox reactive sites, as evidenced by the charge-discharge curve. Table 2 summarizes the coulomb efficiency and the number of available redox reactive sites of the La1−xCaxCrO3 (x = 0–0.2) electrode at a current density of 0.5 A/g. La1−xCaxCrO3 (x = 0–0.2) electrodes participated in the redox reaction to a limited extent, with participation rates of 7.2%, 8.3%, 11.8% and 9.4%, respectively. La0.85Ca0.15CrO3 electrode achieves optimal active site density (11.8%), correlating with its highest specific capacitance. Current density changes significantly affect the coulomb efficiency and active site utilization of La1−xCaxCrO3 electrodes. At low current density, the electrochemical reaction is stable with high coulomb efficiency but lower active site utilization. Conversely, high current density increases reaction irreversibility and reduces coulomb efficiency, while significantly improving active site utilization. Optimizing the electrode structure, material modification, electrolyte, and battery design can reduce internal resistance at high current density and enhance electrochemical performance [66].
The AC impedance spectra of La1−xCaxCrO3 (x = 0–0.2) series, as illustrated in Figure 7a, reveals two distinct regions: a semi-circular section at high frequencies and a linear region at low frequencies [67,68,69]. The semi-circular portion of the high-frequency spectrum represents the redox reaction process occurring at the electrode surface, while the linear portion represents the capacitive effect generated at the electrode surface [70,71]. The charge transfer resistance is determined by fitting a semicircular curve, with the magnitude of the resistance directly reflected in the diameter of the semicircle [72]. Figure 7b,c show the bode plots for La1−xCaxCrO3. bode plots show that the low-frequency impedance mode value increases and the phase angle changes, providing evidence of the adsorption phenomenon on the electrode surface. Calculation concluded that LaCrO3 exhibits the largest transfer resistance. This finding is consistent with the results of the previous GCD curve. Electrochemical impedance spectroscopy reveals impedance values of 0.66, 0.56, 0.52, and 0.58 Ω for x = 0, 0.1, 0.15, and 0.2, respectively. La0.85Ca0.15CrO3 electrode exhibited the lowest equivalent resistance, confirming superior electrochemical performance. The stability of the La0.85Ca0.15CrO3 electrode was evaluated over 5000 continuous charge and discharge cycles at a constant current density of 1 A·g−1 in 0.1–0.55 V. Figure 7d confirms the exceptional cyclability of La0.85Ca0.15CrO3 electrode. The capacitance remains at 94% demonstrates even after 5000 cycles. To verify the cycling performance of La1−xCaxCrO3 in practical applications, a symmetrical supercapacitor was assembled using two La0.85Ca0.15CrO3 electrodes, PET films and 1 M KOH solution. Long-cycle test was conducted at a current density of 1 A·g−1, as shown in Figure 7e. After 5000 long-cycle tests, the electrode’s specific capacity gradually increased. This increase is due to initial electrolyte penetration and ion intercalation/deintercalation reopening pores. This impeded the effective entry of ions, consequently leading to an incomplete release of the designated capacity. As the cycle progresses, electrolyte penetration and ion intercalation/deintercalation progressively open closed pores, increasing specific surface area and active sites, thereby enhancing capacity. The coulombic efficiency remained above 94%, and the partial charge and discharge curves of the electrode also proved La1−xCaxCrO3’s excellent stability. These findings suggest that La0.85Ca0.15CrO3 nanoparticles are a promising negative electrode material for supercapacitors.

4. Conclusions

In this study, a series of La1−xCaxCrO3 (x = 0–0.2) samples were successfully synthesized by the sol-gel method, and their microstructure, morphology and electrochemical properties were systematically studied. Results indicate that the incorporation of calcium has a notable impact on both the crystal structure and the electrochemical properties of LaCrO3. Calcium doping leads to changes in the lattice parameters of LaCrO3 and simultaneously increases the concentration of oxygen vacancies. These oxygen vacancies act as active sites, promoting the adsorption of OH and the diffusion of O2 ions in the electrolyte, thereby enhancing the specific capacitance of the material. In addition, calcium doping significantly increased the specific surface area of LaCrO3 from 38.01 m2/g to 48.21 m2/g, providing more active sites for electrochemical reactions and further enhancing the electrochemical performance. In electrochemical tests, La0.85Ca0.15CrO3 sample exhibited the highest specific capacitance of 306 F/g at 1 A/g, which was approximately 2.3 times higher than the undoped sample (133 F/g). Meanwhile, its charge transfer resistance is the lowest, only 0.52 Ω, indicating the fastest electrode reaction kinetics. After 5000 charging-discharge cycles, the specific capacitance retention rate of La0.85Ca0.15CrO3 electrode was 94%, demonstrating excellent cycling stability. In conclusion, calcium doping significantly enhances the electrochemical performance of LaCrO3 by altering the crystal structure, increasing oxygen vacancies, and improving the specific surface area, making it a promising candidate for supercapacitor electrode materials. Future research could explore other dopants or composite structures to further enhance lanthanide perovskite performance.

Author Contributions

Conceptualization: X.G. and X.S.; methodology: X.G.; software: X.G.; validation: X.G.; formal analysis: X.G.; investigation: Y.Q.; resources: S.D.; data curation: L.W.; writing—original draft preparation: X.G.; writing—review and editing: L.W.; visualization: S.D.; supervision: S.D.; project administration: Y.Q.; funding acquisition: Y.Q. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Jiangsu Practice Innovation Program (SJCX24_2508), the National Natural Science Foundation of China with the number of 52171071 and 51501074.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

No potential conflict of interest was reported by the authors.

References

  1. Anil, N.; Rao, P.K.; Sarkar, A.; Kubavat, J.; Vadivel, S.; Manwar, N.R.; Paul, B. Advancements in sustainable biodiesel production: A comprehensive review of bio-waste derived catalysts. Energy Convers. Manag. 2024, 318, 1188874. [Google Scholar] [CrossRef]
  2. Farghali, M.; Osman, A.I.; Mohamed, I.M.A.; Chen, Z.; Chen, L.; Ihara, I.; Yap, P.-S.; Rooney, D.W. Strategies to save energy in the context of the energy crisis: A review. Environ. Chem. Lett. 2023, 21, 2003–2039. [Google Scholar] [CrossRef] [PubMed]
  3. Tan, H.; Li, J.; He, M.; Li, J.; Zhi, D.; Qin, F.; Zhang, C. Global evolution of research on green energy and environmental technologies: A bibliometric study. J. Environ. Manag. 2021, 297, 113382. [Google Scholar] [CrossRef] [PubMed]
  4. Soeder, D.J. Fossil fuels and climate change. In Fracking and the Environment; Springer International Publishing: Cham, Switzerland, 2021; pp. 155–185. [Google Scholar]
  5. Borowski, P.F. Mitigating Climate Change and the Development of Green Energy versus a Return to Fossil Fuels Due to the Energy Crisis in 2022. Energies 2022, 15, 9289. [Google Scholar] [CrossRef]
  6. Akalin, G.; Erdogan, S.; Sarkodie, S.A. Do dependence on fossil fuels and corruption spur ecological footprint. Environ. Impact Assess. Rev. 2021, 90, 106641. [Google Scholar] [CrossRef]
  7. Lu, Y.; Khan, Z.A.; Alvarez-Alvarado, M.S.; Zhang, Y.; Huang, Z.; Imran, M. A Critical Review of Sustainable Energy Policies for the Promotion of Renewable Energy Sources. Sustainability 2020, 12, 5078. [Google Scholar] [CrossRef]
  8. Yadav, G.; Ahmaruzzaman, M. Multi anion-based materials: Synthesis and catalytic applications. Mater. Res. Bull. 2022, 152, 111836. [Google Scholar] [CrossRef]
  9. Tan, K.M.; Babu, T.S.; Ramachandaramurthy, V.K.; Kasinathan, P.; Solanki, S.G.; Raveendran, S.K. Empowering smart grid: A comprehensive review of energy storage technology and application with renewable energy integration. J. Energy Storage 2021, 39, 102591. [Google Scholar] [CrossRef]
  10. Sharma, M.; Chauhan, A.; Vaish, R. Energy harvesting using piezoelectric cementitious composites for water cleaning applications. Mater. Res. Bull. 2021, 137, 111205. [Google Scholar] [CrossRef]
  11. Karthikeyan, S.; Narenthiran, B.; Sivanantham, A.; Bhatlu, L.D.; Maridurai, T. Supercapacitor: Evolution and review. Mater. Today Proc. 2021, 46, 3984–3988. [Google Scholar] [CrossRef]
  12. Karabacak, K. Towards sustainable solar energy solutions: Harnessing supercapacitors in PV systems. Energy Storage Convers. 2024, 2, 436. [Google Scholar] [CrossRef]
  13. Karunakaran, S.K.; Arumugam, G.M.; Yang, W.; Ge, S.; Khan, S.N.; Lin, X.; Yang, G. Research Progress on the Application of Lanthanide-Ion-Doped Phosphor Materials in Perovskite Solar Cells. ACS Sustain. Chem. Eng. 2021, 9, 1035–1060. [Google Scholar] [CrossRef]
  14. Arumugam, G.M.; Karunakaran, S.K.; Galian, R.E.; Pérez-Prieto, J. Recent Progress in Lanthanide-Doped Inorganic Perovskite Nanocrystals and Nanoheterostructures: A Future Vision of Bioimaging. Nanomaterials 2022, 12, 2130. [Google Scholar] [CrossRef] [PubMed]
  15. Qu, J.; Guo, X.; Tian, Z.; Wang, H.; Ye, X.; Wang, L.; Dong, S. Effect of Sr Doping at the A Site on the Electrochemical Properties of LaCoO3 Used for Supercapacitor Applications. J. Electron. Mater. 2025, 54, 3851–3862. [Google Scholar] [CrossRef]
  16. Sun, X.; Pei, Z.; Guo, X.; Ye, X.; Wang, L.; Zhang, Y.; Dong, S. Impact of Ca ions substitution at A-site on LaCoO3 perovskite energy applications. Mater. Sci. Eng. 2025, 319, 118343. [Google Scholar] [CrossRef]
  17. Guo, X.; Tian, Z.; Qu, J.; Ye, X.; Wang, L.; Zhang, Y.; Dong, S.; Ju, H. Enhanced the electrochemical performance in Sr-doped LaCoO3 nanofibers. J. Alloys Compd. 2025, 1031, 181003. [Google Scholar] [CrossRef]
  18. Ashok C, S.; Vazhayil, A.; Vinaykumar, R.; Thomas, J.; Jeffery, A.A.; Hasan, I.; Thomas, N.; Yadav, A.K.; Ahn, Y.-H. The effect of B–site cation on the supercapacitive properties of LaBO3 (B = Cr, Mn, Fe and Co) porous perovskites. J. Energy Storage 2024, 86, 111145. [Google Scholar] [CrossRef]
  19. Dutta, S. Investigation of Photocatalytic Properties of Molybdenum Disulfide Incorporated Lanthanum Ferrite Nanocomposites. Master’s Thesis, Bangladesh University of Engineering and Technology, Dhaka, Bangladesh, February 2020. [Google Scholar]
  20. Sun, H.; Song, J.; Shi, P.; Cheng, Z.; Tian, L.; Zhou, M.; Qi, T. Selective detection of NO using the perovskite-type oxide LaMO3 (M = Cr, Mn, Fe, Co, and Ni) as the electrode material for yttrium-stabilized zirconia-based electrochemical sensors. J. Mater. Chem. A 2024, 12, 18234–18247. [Google Scholar] [CrossRef]
  21. Yadagiri, K.; Anvay, Y.; Dinakar, D.; Narasaiah, N.; Sathe, V.G.; Kumar, K.; Haranath, D. Structural transformation and magnetic properties of Fe-substituted nano CuCr2O4 spinel structure. Ceram. Int. 2023, 50, 4987–4993. [Google Scholar] [CrossRef]
  22. Qahtan, A.A.; Zarrin, N.; Fatema, M.; Al-Shamiri, H.A.; Khan, W.; Husain, S. A brief review on rare-earth chromites: Synthesis, properties and applications. Phys. Scr. 2024, 99, 072001. [Google Scholar] [CrossRef]
  23. Al-Awadi, A.S.; Al-Zahrani, S.M.; El-Toni, A.M.; Abasaeed, A.E. Dehydrogenation of Ethane to Ethylene by CO2 over Highly Dispersed Cr on Large-Pore Mesoporous Silica Catalysts. Catalysts 2020, 10, 97. [Google Scholar] [CrossRef]
  24. Bachankar, S.; Malavekar, D.; Lokhande, V.; Ji, T. Empowering supercapacitors: Strategic B-site modulations of transition metal atoms in perovskite oxide based electrodes. J. Alloys Compd. 2024, 1003, 175708. [Google Scholar] [CrossRef]
  25. Li, S.; Irvine, J.T. Non-stoichiometry, structure and properties of proton-conducting perovskite oxides. Solid State Ion. 2021, 361, 115571. [Google Scholar] [CrossRef]
  26. Sun, X.; Meng, Z.; Hao, Z.; Du, Z.; Xu, J.; Nan, H.; Shi, W.; Zeng, F.; Hu, X.; Tian, H. Efficient fabrication of flower-like core–shell nanochip arrays of lanthanum manganate and nickel cobaltate for high-performance supercapacitors. J. Colloid Interface Sci. 2022, 630, 618–628. [Google Scholar] [CrossRef] [PubMed]
  27. Mo, H.; Nan, H.; Lang, X.; Liu, S.; Qiao, L.; Hu, X.; Tian, H. Influence of calcium doping on performance of LaMnO3 supercapacitors. Ceram. Int. 2018, 44, 9733–9741. [Google Scholar] [CrossRef]
  28. Shafi, P.M.; Mohapatra, D.; Reddy, V.P.; Dhakal, G.; Kumar, D.R.; Tuma, D.; Brousse, T.; Shim, J.-J. Sr- and Fe-substituted LaMnO3 Perovskite: Fundamental insight and possible use in asymmetric hybrid supercapacitor. Energy Storage Mater. 2022, 45, 119–129. [Google Scholar] [CrossRef]
  29. Dong, S.; Jin, X.; Ye, X.; Wei, J.; Wang, L.; Zhang, Y. Effects of Calcium Substitution for La on the Electrochemical Performance of LaMnO3 Nanoparticles. ChemistrySelect 2023, 8, e202203977. [Google Scholar] [CrossRef]
  30. Dong, S.-T.; Ye, X.; Fu, Z.; Jin, X.; Wei, J.; Wang, L.; Zhang, Y.-M. Effects of strontium substitution for La on the electrochemical performance of LaAlO3 perovskite nanotubes. J. Mater. Res. Technol. 2022, 19, 91–100. [Google Scholar] [CrossRef]
  31. Alexander, C.T.; Mefford, J.T.; Saunders, J.; Forslund, R.P.; Johnston, K.P.; Stevenson, K.J. Anion-Based Pseudocapacitance of the Perovskite Library La1−xSrxBO3−δ (B = Fe, Mn, Co). ACS Appl. Mater. Interfaces 2019, 11, 5084–5094. [Google Scholar] [CrossRef] [PubMed]
  32. Qayyum, A.; Rehman, M.O.U.; Ahmad, F.; Khan, M.A.; Ramay, S.M.; Atiq, S. Performance optimization of Nd-doped LaNiO3 as an electrode material in supercapacitors. Solid State Ion. 2023, 395, 116227. [Google Scholar] [CrossRef]
  33. Liu, P.; Weng, X.; Liu, Z.; Zhang, Y.; Qiu, Q.; Wang, W.; Zhou, M.; Cai, W.; Ni, M.; Liu, M.; et al. High-Performance Quasi-Solid-State Supercapacitor Based on CuO Nanoparticles with Commercial-Level Mass Loading on Ceramic Material La1−xSrxCoO3−δ as Cathode. ACS Appl. Energy Mater. 2019, 2, 1480–1488. [Google Scholar] [CrossRef]
  34. Zhao, H.; Zhu, Q.; Ye, X.; Wang, L.; Dong, S.-T. Electrochemical Properties of LaMO3 (M= Cr, Mn, and Co) Perovskite Materials. Coatings 2024, 14, 147. [Google Scholar] [CrossRef]
  35. Jiang, S.P.; Liu, L.; Ong, K.P.; Wu, P.; Li, J.; Pu, J. Electrical conductivity and performance of doped LaCrO3 perovskite oxides for solid oxide fuel cells. J. Power Sources 2008, 176, 82–89. [Google Scholar] [CrossRef]
  36. Rashtchi, H.; Sani, M.A.F.; Dayaghi, A.M. Effect of Sr and Ca dopants on oxidation and electrical properties of lanthanum chromite-coated AISI 430 stainless steel for solid oxide fuel cell interconnect application. Ceram. Int. 2013, 39, 8123–8131. [Google Scholar] [CrossRef]
  37. Deshmukh, V.; Tejashwini, D.; Nagaswarupa, H.; Naik, R.; Al-Kahtani, A.A.; Kumar, Y.A. Sr and Fe substituted LaCoO3 nano perovskites: Electrochemical energy storage and sensing applications. J. Energy Storage 2024, 89, 111724. [Google Scholar] [CrossRef]
  38. Leng, L.; Xiong, Q.; Yang, L.; Li, H.; Zhou, Y.; Zhang, W.; Jiang, S.; Li, H.; Huang, H. An overview on engineering the surface area and porosity of biochar. Sci. Total. Environ. 2021, 763, 144204. [Google Scholar] [CrossRef] [PubMed]
  39. Tang, Z.; Feng, Z.; Deng, D.; Sun, D.; Chen, M.; Han, K.; Ye, H.; Wang, H.; Tang, Y. Optimization of Sr-doping boosting the structural stability for single crystalline LiNi0.8Co0.1Mn0.1O2 cathode to enhance its electrochemical performance at elevated voltage and temperature. J. Power Sources 2022, 545, 231915. [Google Scholar] [CrossRef]
  40. Wang, K.; Xu, Z.; Guo, Z.; Wang, H.; Qaid, S.M.H.; Yang, K.; Zang, Z. Phosphonate Diacid Molecule Induced Crystallization Manipulation and Defect Passivation for High-Performance Inverted MA-Free Perovskite Solar Cells. Adv. Energy Mater. 2024, 14, 2402249. [Google Scholar] [CrossRef]
  41. Wang, H.; Yang, Y.; Yang, L. Optimization of dielectric loss in calcium copper titanate based on different doping modification strategies. Ceram. Int. 2023, 49, 38399–38419. [Google Scholar] [CrossRef]
  42. Ashfaq, A.; Sabugaa, M.M.; Ben Moussa, M.; Almousa, N.; Shokralla, E.A.; Capangpangan, R.Y.; Alguno, A.C.; Hossain, A.; Alanazi, A.M.; Abboud, M. High thermoelectric power factor of Sr doped Bi2Te3 thin film through energy filtering effect. Int. Commun. Heat Mass Transf. 2023, 143, 106719. [Google Scholar] [CrossRef]
  43. Zhang, W.; Li, F.; Zhang, W.; Wang, Y.; Liu, S.; Mo, X.; Li, K. B-site cation synergy in Mn doped LaCoO3 for enhanced electrocatalysts for electrochemical reduction of nitrate. Sep. Purif. Technol. 2023, 317, 123775. [Google Scholar] [CrossRef]
  44. Xu, C.; Lei, C.; Wang, Y.; Yu, C. Dendritic Mesoporous Nanoparticles: Structure, Synthesis and Properties. Angew. Chem. Int. Ed. Engl. 2021, 61, e202112752. [Google Scholar] [CrossRef] [PubMed]
  45. Ye, X.; Dong, S.; Jin, X.; Wei, J.; Wang, L.; Zhang, Y. Enhancement in the Electrochemical Performance of Strontium (Sr)-Doped LaMnO3 as Supercapacitor Materials. Coatings 2022, 12, 1739. [Google Scholar] [CrossRef]
  46. Zhang, D.; Ma, W.; Wu, S.; Lu, G.; Zhang, Y.; Zheng, K.; Wang, Y.-J.; Li, K. Rational engineering of meso-macroporous structured carbon materials for revealing capacitive mechanism. J. Energy Storage 2024, 104, 114478. [Google Scholar] [CrossRef]
  47. Takegami, D.; Fujinuma, K.; Nakamura, R.; Yoshimura, M.; Tsuei, K.-D.; Wang, G.; Wang, N.N.; Cheng, J.-G.; Uwatoko, Y.; Mizokawa, T. Absence of Ni2+/Ni3+ charge disproportionation and possible roles of O 2p holes in La3Ni2O7−δ revealed by hard x-ray photoemission spectroscopy. Phys. Rev. B Condens. Matter. 2024, 109, 125119. [Google Scholar] [CrossRef]
  48. Mena, J.A.; Mendoza-Estrada, V.; Sabolsky, E.M.; Sierros, K.A.; Sabolsky, K.; González-Hernández, R.; Idhaiam, K.S.V. Exploring the impact of multivalent substitution on high-temperature electrical conductivity in doped lanthanum chromite perovskites: An experimental and ab-initio study. J. Eur. Ceram. Soc. 2024, 44, 7040–7058. [Google Scholar] [CrossRef]
  49. Silva, R.S., Jr.; Ferreira, N.; Fontes, J.F.D.; da Costa, M.E.H.M.; Barrozo, P. Toward the stabilization of perovskite phase at low temperature and decrease of the magnetic ordering by Sr2+-doping in LaCrO3. Chem. Phys. Lett. 2022, 787, 139278. [Google Scholar] [CrossRef]
  50. Mazza, A.R.; Skoropata, E.; Lapano, J.; Zhang, J.; Sharma, Y.; Musico, B.L.; Keppens, V.; Gai, Z.; Brahlek, M.J.; Moreo, A. Charge doping effects on magnetic properties of single-crystal La1−xSrx (Cr0.2Mn0.2Fe0.2Co0.2Ni0.2)O3 (0≤ x ≤ 0.5) high-entropy perovskite oxides. Phys. Rev. B Condens. Matter. 2021, 104, 094204. [Google Scholar] [CrossRef]
  51. Hu, M.; Jing, Y.; Zhang, X. Low thermal conductivity of graphyne nanotubes from molecular dynamics study. Phys. Rev. 2015, 91, 155408. [Google Scholar] [CrossRef]
  52. Dhandapani, P.; Nayak, P.K.; Maruthapillai, A. Improved electrochemical performance and charge storage mechanism of NiMnCoO4 by XPS study. Mater. Chem. Phys. 2023, 297, 127287. [Google Scholar] [CrossRef]
  53. Tapia-P, J.; Gallego, J.; Gamba, O.; Espinal, J. Insight into the Interaction of Perovskite-Like Surfaces (LaMnO3 and LaCoO3) with Ar, H2, CO, and O2 through NAP-XPS Analysis. Catal. Lett. 2024, 154, 6133–6142. [Google Scholar] [CrossRef]
  54. Wang, Z.; Hu, N.; Wang, L.; Zhao, H.; Zhao, G. In Situ Production of Hydroxyl Radicals via Three-Electron Oxygen Reduction: Opportunities for Water Treatment. Angew. Chem. 2024, 63, e202407628. [Google Scholar] [CrossRef] [PubMed]
  55. Gan, L.-Y.; Akande, S.O.; Schwingenschlögl, U. Anisotropic O vacancy formation and diffusion in LaMnO3. J. Mater. Chem. 2014, 2, 19733–19737. [Google Scholar] [CrossRef]
  56. Xia, W.; Leng, K.; Tang, Q.; Yang, L.; Xie, Y.; Wu, Z.; Yi, K.; Zhu, X. Structural characterization, magnetic and optical properties of perovskite (La1−xLnx)0.67Ca0.33MnO3 (Ln= Nd and Sm; x= 0.0–0.5) nanoparticles synthesized via the sol-gel process. J. Alloys Compd. 2021, 867, 158808. [Google Scholar] [CrossRef]
  57. Tamirat, A.G.; Guan, X.; Liu, J.; Luo, J.; Xia, Y. Redox mediators as charge agents for changing electrochemical reactions. Chem. Soc. Rev. 2020, 49, 7454–7478. [Google Scholar] [CrossRef] [PubMed]
  58. Li, C.; Liu, J.; Lin, L.; Bai, W.; Wu, S.; Zheng, P.; Zhang, J.; Zhai, J. Superior Energy Storage Capability and Stability in Lead-Free Relaxors for Dielectric Capacitors Utilizing Nanoscale Polarization Heterogeneous Regions. Small 2023, 19, e2206662. [Google Scholar] [CrossRef] [PubMed]
  59. Coslop, G. Experimental Investigation of La(0.6)Sr(0.4)Mn(0.6)Cr(0.4)O3 Perovskite Oxidation Kinetics for Thermochemical Carbon Dioxide Splitting Redox Cycles. Master’s Thesis, The Politecnico di Torino, Torino, Italy, March 2023. [Google Scholar]
  60. Huang, L.; Cheng, L.; Pan, S.; He, Y.; Tian, C.; Yu, J.; Zhou, H. Effects of Sr doping on the structure, magnetic properties and microwave absorption properties of LaFeO3 nanoparticles. Ceram. Int. 2020, 46, 27352–27361. [Google Scholar] [CrossRef]
  61. Yadlapalli, R.T.; Alla, R.R.; Kandipati, R.; Kotapati, A. Super capacitors for energy storage: Progress, applications and challenges. J. Energy Storage 2022, 49, 104194. [Google Scholar] [CrossRef]
  62. Zhu, Q.; Li, J.; Simon, P.; Xu, B. Two-dimensional MXenes for electrochemical capacitor applications: Progress, challenges and perspectives. Energy Storage Mater. 2021, 35, 630–660. [Google Scholar] [CrossRef]
  63. Guo, X.; Sun, X.; Ju, H.; Jin, L.; Wang, L.; Dong, S. Facile synthesis of perovskite La2CoMnO6 nanoparticles for high-performance supercapacitor electrodes. J. Energy Storage 2025, 130, 117444. [Google Scholar] [CrossRef]
  64. Pu, L.; Zhang, J.; Jiresse, N.K.L.; Gao, Y.; Zhou, H.; Naik, N.; Gao, P.; Guo, Z. N-doped MXene derived from chitosan for the highly effective electrochemical properties as supercapacitor. Adv. Compos. Hybrid Mater. 2021, 5, 356–369. [Google Scholar] [CrossRef]
  65. Singh, N.; Tanwar, S.; Kumar, P.; Sharma, A.; Yadav, B. Advanced sustainable solid state energy storage devices based on FeOOH nanorod loaded carbon@PANI electrode: GCD cycling and TEM correlation. J. Alloys Compd. 2023, 947, 169580. [Google Scholar] [CrossRef]
  66. An, S.; Gao, X.; Zheng, M.; Zhang, M.; Liu, P.; Zhu, M.; Hou, Y. Boosting power output performance of Ba0.85Ca0.15SnZr0.1-Ti0.9O3 piezoelectric nanogenerators via internal resistance regulation strategy. Nano Energy 2024, 123, 109420. [Google Scholar] [CrossRef]
  67. Chen, W.; Li, Y. Energy harvesting performance of an elastically mounted semi-circular cylinder. Renew. Energy 2024, 229, 120692. [Google Scholar] [CrossRef]
  68. Fujiwara, K.; Sriram, R.; Kontis, K. Experimental investigations on the sharp leading-edge separation over a flat plate at zero incidence using particle image velocimetry. Exp. Fluids 2020, 61, 205. [Google Scholar] [CrossRef]
  69. Sreematha, B.; Arundhathi, N.; Ravinder, D. Influence of Cerium substitution on structural, optical, and electrical transport properties of Ni-Nano ferrites prepared by Citrate gel auto combustion method. Results Chem. 2023, 5, 100929. [Google Scholar] [CrossRef]
  70. de Sá, M.; Pereira, C.M. The relevance of the initial conditions in glassy carbon electrode sensing applications: The ferri/ferrocyanide redox reaction model system in aqueous solution. Electrochim. Acta 2024, 489, 141158. [Google Scholar] [CrossRef]
  71. Mallikarjuna, K.; Reddy, G.R.; Ghanem, M.A.; Dillip, G.; Joo, S. Insights of oxygen vacancies engineered structural, morphological, and electrochemical attributes of Cr-doped Co3O4 nanoparticles as redox active battery-type electrodes for hybrid supercapacitors. J. Energy Storage 2024, 100, 113751. [Google Scholar] [CrossRef]
  72. Pfeiffer, N.; Wachter, T.; Frickel, J.; Halima, H.; Hofmann, C.; Errachid, A.; Heuberger, A. Determination of Charge Transfer Resistance from Randles Circuit Spectra Using Elliptical Fitting. In Biomedical Engineering Systems and Technologies, Proceedings of the 14th International Joint Conference on Biomedical Engineering Systems and Technologies (BIOSTEC 2021), Vienna, Austria, 11–13 February 2021; Springer: Berlin, Germany, 2023; pp. 61–79. [Google Scholar]
Figure 1. (a) XRD patterns of La1−xCaxCrO3 (x = 0–0.2) series samples; (b) A partial enlarged view of the (110) diffraction peak of La1−xCaxCrO3 (x = 0–0.2) series samples; (c) XPS full spectrum of La1−xCaxCrO3 (x = 0–0.2) nanoparticles.
Figure 1. (a) XRD patterns of La1−xCaxCrO3 (x = 0–0.2) series samples; (b) A partial enlarged view of the (110) diffraction peak of La1−xCaxCrO3 (x = 0–0.2) series samples; (c) XPS full spectrum of La1−xCaxCrO3 (x = 0–0.2) nanoparticles.
Coatings 15 00837 g001
Figure 2. SEM images of (a) LaCrO3; (b) La0.9Ca0.1CrO3; (c) La0.85Ca0.15CrO3; (d) La0.8Ca0.2CrO3.
Figure 2. SEM images of (a) LaCrO3; (b) La0.9Ca0.1CrO3; (c) La0.85Ca0.15CrO3; (d) La0.8Ca0.2CrO3.
Coatings 15 00837 g002
Figure 3. (a) N2 adsorption-desorption isotherms of LaCrO3; (b) pore size distributions of the LaCrO3; (c) N2 adsorption-desorption isotherms of La0.85Ca0.15CrO3; (d) pore size distributions of the La0.85Ca0.15CrO3.
Figure 3. (a) N2 adsorption-desorption isotherms of LaCrO3; (b) pore size distributions of the LaCrO3; (c) N2 adsorption-desorption isotherms of La0.85Ca0.15CrO3; (d) pore size distributions of the La0.85Ca0.15CrO3.
Coatings 15 00837 g003
Figure 4. (a) high-resolution XPS map of La 3d; (b) high-resolution XPS map of Cr 2p; (c) high-resolution XPS map of O 1s; (d) high-resolution XPS map of Ca 2p.
Figure 4. (a) high-resolution XPS map of La 3d; (b) high-resolution XPS map of Cr 2p; (c) high-resolution XPS map of O 1s; (d) high-resolution XPS map of Ca 2p.
Coatings 15 00837 g004
Figure 5. CV curves of (a) LaCrO3; (b) La0.9Ca0.1CrO3; (c) La0.85Ca0.15CrO3; (d) La0.8Ca0.2CrO3; (e) bare Ni-form.
Figure 5. CV curves of (a) LaCrO3; (b) La0.9Ca0.1CrO3; (c) La0.85Ca0.15CrO3; (d) La0.8Ca0.2CrO3; (e) bare Ni-form.
Coatings 15 00837 g005
Figure 6. GCD curves of (a) LaCrO3; (b) La0.9Ca0.1CrO3; (c) La0.85Ca0.15CrO3; (d) La0.8Ca0.2CrO3 samples.
Figure 6. GCD curves of (a) LaCrO3; (b) La0.9Ca0.1CrO3; (c) La0.85Ca0.15CrO3; (d) La0.8Ca0.2CrO3 samples.
Coatings 15 00837 g006
Figure 7. (a) Impedance comparison diagram of La1−xCaxCrO3 (x = 0–0.2) electrodes; Bode plots of (b,c); (d) Cycle stability diagram of electrodes prepared by La0.85Ca0.15CrO3 using foamed nickel as current; (e) Long-cycle tests and partial charge-discharge curves during cycling of symmetrical supercapacitors assembled with La0.85Ca0.15CrO3 at a current density of 1 A·g−1.
Figure 7. (a) Impedance comparison diagram of La1−xCaxCrO3 (x = 0–0.2) electrodes; Bode plots of (b,c); (d) Cycle stability diagram of electrodes prepared by La0.85Ca0.15CrO3 using foamed nickel as current; (e) Long-cycle tests and partial charge-discharge curves during cycling of symmetrical supercapacitors assembled with La0.85Ca0.15CrO3 at a current density of 1 A·g−1.
Coatings 15 00837 g007
Table 1. Relative concentration of three oxygen in La1−xCaxCrO3 (x = 0–0.2).
Table 1. Relative concentration of three oxygen in La1−xCaxCrO3 (x = 0–0.2).
LaCrO3La0.9Ca0.1CrO3La0.85Ca0.15CrO3La0.8Ca0.2CrO3
O 1(%)45.844.742.143.9
O 2(%)47.649.655.649.2
O 3(%)6.65.72.36.9
O 2/O 11.041.111.321.12
Table 2. The columbic efficiency and available redox reaction active sites of La1−xCaxCrO3 (x = 0–0.2) at a current density of 0.5 A/g.
Table 2. The columbic efficiency and available redox reaction active sites of La1−xCaxCrO3 (x = 0–0.2) at a current density of 0.5 A/g.
Coulombic Efficiency (%)Active Site
LaCrO388.40.072
La0.9Ca0.1CrO387.70.083
La0.85Ca0.15CrO391.40.118
La0.8Ca0.2CrO391.10.094
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

Guo, X.; Sun, X.; Wang, L.; Qiao, Y.; Dong, S. Investigation of the Structure and Electrochemical Performance of Perovskite Oxide La1−xCaxCrO3 Utilized as Electrode Materials for Supercapacitors. Coatings 2025, 15, 837. https://doi.org/10.3390/coatings15070837

AMA Style

Guo X, Sun X, Wang L, Qiao Y, Dong S. Investigation of the Structure and Electrochemical Performance of Perovskite Oxide La1−xCaxCrO3 Utilized as Electrode Materials for Supercapacitors. Coatings. 2025; 15(7):837. https://doi.org/10.3390/coatings15070837

Chicago/Turabian Style

Guo, Xu, Xin Sun, Lei Wang, Yanxin Qiao, and Songtao Dong. 2025. "Investigation of the Structure and Electrochemical Performance of Perovskite Oxide La1−xCaxCrO3 Utilized as Electrode Materials for Supercapacitors" Coatings 15, no. 7: 837. https://doi.org/10.3390/coatings15070837

APA Style

Guo, X., Sun, X., Wang, L., Qiao, Y., & Dong, S. (2025). Investigation of the Structure and Electrochemical Performance of Perovskite Oxide La1−xCaxCrO3 Utilized as Electrode Materials for Supercapacitors. Coatings, 15(7), 837. https://doi.org/10.3390/coatings15070837

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