Next Article in Journal
Green Goals, Financial Gains: SDG 7 “Affordable and Clean Energy” and Bank Profitability in Romania
Previous Article in Journal
Upgrading/Deacidification of Biofuels (Gasoline, Kerosene, and Diesel-like Hydrocarbons) by Adsorption Using Activated Red-Mud-Based Adsorbents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structure and Electrochemical Behavior of ZnLaFeO4 Alloy as a Negative Electrode in Ni-MH Batteries

1
Laboratoire de Mécanique, Matériaux et Procédés LR99ES05, Ecole Nationale Supérieure d’Ingénieurs de Tunis, Université de Tunis, Tunis 1008, Tunisia
2
ICB, UMR 6303 CNRS-Universite de Bourgogne, 9 Avenue Alain Savary, 47870 Dijon, Cedex, France
3
Laboratoire de Physico-Chimie de l’Atmosphère, Université du Littoral Côte d’Opale, 59140 Dunkerque, France
4
FEMTO-ST, MN2S, UMLP, UTBM, 90000 Belfort, Cedex, France
5
ICB-PMDM/FR FCLAB, UMLP, UBE, UTBM, 90000 Belfort, Cedex, France
*
Author to whom correspondence should be addressed.
Energies 2025, 18(13), 3251; https://doi.org/10.3390/en18133251 (registering DOI)
Submission received: 15 April 2025 / Revised: 6 June 2025 / Accepted: 11 June 2025 / Published: 21 June 2025

Abstract

:
This study focuses on the structural and electrochemical behavior of the compound ZnLaFeO4 as a negative electrode material for nickel–metal hydride (Ni-MH) batteries. The material was synthesized by a sol–gel hydrothermal method to assess the influence of lanthanum doping on the ZnFe2O4 spinel structure. X-ray diffraction revealed the formation of a dominant LaFeO3 perovskite phase, with ZnFe2O4 and La2O3 as secondary phases. SEM analysis showed agglomerated grains with an irregular morphology. Electrochemical characterization at room temperature and a discharge rate of C/10 (full charge in 10 h) revealed a maximum discharge capacity of 106 mAhg−1. Although La3+ doping modified the microstructure and slowed the activation process, the electrode exhibited stable cycling with moderate polarization behavior. The decrease in capacity during cycling is due mainly to higher internal resistance. These results highlight the potential and limitations of La-doped spinel ferrites as alternative negative electrodes for Ni-MH systems.

1. Introduction

Recently, mobile tools have become one of the most significant sources of pollution, consuming nearly half of the world’s oil production and accounting for a quarter of global CO2 emissions [1,2,3]. This dramatic increase in oil resource depletion, coupled with the growing environmental awareness, has created an urgent need to develop a new model of energy-efficient and eco-friendly vehicles, such as hybrid electric vehicles (HEVs) and electric vehicles (EVs) [4,5]. The key challenge in the development of EVs and HEVs lies in battery performance [6]. As a solution, lithium-ion (Li-ion) and nickel–metal hydride (Ni-MH) batteries have emerged as the most critical onboard energy storage systems for these vehicles [7]. While lithium batteries offer the highest energy density, their low thermal stability presents significant safety concerns, particularly in the transportation sector [8,9]. On the other hand, Ni-MH batteries are widely used to reduce CO2 emissions in transportation due to their high energy density, ease of application in series, series/parallel configurations, safe charge and discharge processes, tolerance to overcharge and over-discharge, excellent thermal properties, and the use of environmentally acceptable and recyclable materials [10,11,12,13,14,15,16].
However, the battery performance is influenced by the intrinsic properties of the electrode materials. In Ni-MH batteries, the negative electrode acts as a hydrogen-absorbing material, where hydrogen is reversibly inserted and released during the charging and discharging processes. During charging, hydrogen is stored in the form of metal hydrides (MHₓ), and during discharging, it is released in the form of hydrogen ions that participate in the redox reaction with the positive electrode. Various metal compounds with high discharge capacities, such as AB5 [17,18,19], AB2 [20], AB3 [21], AB [22], and A2B [23], have been studied since the invention of Ni-MH batteries. Despite their promising properties, these compounds are often expensive and difficult to produce, which limits their performance and viability in commercial applications. As a solution, oxide-based ABO3 compounds (e.g., perovskite alloys) have been investigated and developed as alternatives to the intermetallic alloys used in Ni-MH batteries [24,25,26]. However, one of the primary challenges in advancing this technology is the limited operating temperature range for effective hydrogen absorption; to overcome this challenge, ceramic oxide nanomaterials, including ferrites, have attracted the interest of the research community because of their distinctive crystal structures and magnetic characteristics [27,28,29]. These nanomaterials can be classified into three main types: hexaferrites, spinel ferrites, and garnets. In fact, nano-sized spinel ferrites having the chemical formula AB2O4, where A and B occupy the tetrahedral and octahedral cation sites, respectively, are the most widely utilized, due to their fascinating structural characteristics and outstanding electrical and magnetic properties [30,31,32]. This nanomaterial is extensively used in a wide range of applications, such as microwave and magnetic devices [33,34], wastewater treatment [35], sensors [36], targeted drug delivery [37], anticorrosion pigments [38,39], and optoelectronic systems [40]. It has also demonstrated a promising performance as an anode material in both lithium-ion [41,42] and sodium-ion batteries [43].
Given that the properties of nano-ferrites are highly dependent on the synthesis method, several techniques (e.g., co-precipitation [44], thermal decomposition [45], solvothermal methods [46], the sol–gel technique [47,48], etc.) were explored in previous studies, to prepare spinel ferrites. Among these, the sol–gel technique stands out for offering excellent textural characteristics, high chemical homogeneity, good crystallinity, and high purity of the final product [49].
These results have motivated research into the use of this type of nanomaterial as an electrode material in Ni-MH batteries. For instance, Wissem et al. [50] synthesized zinc spinel ferrites (ZnFe2O4) using the sol–gel method and investigated them as innovative anode materials in Ni-MH batteries. The performed electrochemical characterization revealed the excellent discharge capacity (up to 145 mAh/g) and high stability during cycling.
In complementary studies [51], substitution of the A site (Zn2+) with rare-earth cations such as Sm3⁺ in the ZnxSm1−xFe2O4 system also led to interesting results, highlighting the potential of cation substitution to enhance the electrochemical properties of spinel ferrites.
Based on these results, we propose a new approach based on B-site substitution in the spinel structure, where Fe3+ is partially replaced by La3+ to form the doped compound ZnLaFeO4. This substitution is expected to modify the structure and improve electrochemical behavior by enhancing hydrogen diffusion and electrode stability.
This work aims to investigate the structural, morphological, and electrochemical characteristics of ZnLaFeO4 as an innovative negative electrode material used in Ni-MH batteries.

2. Materials and Methods

2.1. Method Applied to Synthesize the Alloys

The ZnLaFeO4 powders were synthesized using a sol–gel hydrothermal process, employing stoichiometric quantities of high-purity metal nitrates: zinc nitrate hexahydrate (Zn(NO3)2·6H2O, 99.9% purity, Sigma-Aldrich, St. Louis, MO, USA), iron nitrate nonahydrate (Fe(NO3)3·9H2O, ≥99.95% purity, Sigma-Aldrich, St. Louis, MO, USA), and lanthanum nitrate hexahydrate (La(NO3)3·6H2O, 99.999% purity, Sigma-Aldrich, St. Louis, MO, USA). These precursors were dissolved in demineralized water at room temperature, along with citric acid, to form a homogeneous solution. Ethylene glycol was then added to the solution under continuous stirring. After eight hours of agitation, the resulting gel was dried to remove water. Subsequently, it was ground into a fine powder. Finally, the powders were placed in the oven at 900 °C for 12 h to be calcined.

2.2. Characterization Methods

A Siemens X-ray diffractometer with Ni-filtered copper radiation (λ = 1.5404 Å) was used to obtain the X-ray diffraction data of the powders at room temperature. Then, the surface morphology of the alloy powder was analyzed using a scanning electron microscope (SEM).

2.3. Electrochemical Characterization

The negative electrodes were prepared applying latex technology [52,53], which consisted of mixing 90% target compound with 5% carbon black and 5% polytetrafluoroethylene (PTFE). The resulting latex was later dried under a vacuum at room temperature for 24 h. Two pieces of the dried latex were pressed onto each side of a nickel grid, which served as the current collector. Electrochemical measurements were conducted at room temperature in a three-electrode cell, where the coiled nickel wire, the Hg/HgO electrode and the prepared hydride negative electrode were used as auxiliary, reference and working electrodes, respectively. A concentrated potassium hydroxide (KOH) solution (6 M) was used as the electrolyte. The negative hydride electrode was characterized by various electrochemical methods utilizing an EC-Lab potentiostat–galvanostat. All measurements were performed at a C/10 rate. Chronopotentiometry was employed to evaluate the cycling properties, including the activation of the negative electrode, maximum electrochemical capacity, and polarization. This technique involves applying a constant current between the working and auxiliary electrodes while monitoring the potential change over time during a series of cycles [54] and potentiodynamic polarization, which involves applying a potential E on the LaZnFe2O4 negative electrode and varying it linearly with time at a slew rate of v = 1 mVs−1.

3. Results

3.1. SEM Observations

Figure 1 displays scanning electron microscopy (SEM) images showcasing the microstructures of ZnLaFeO4 spinel ferrites.
The micrographs of the ZnLaFeO4 alloys show substantial particle agglomeration, which hinders the accurate determination of grain sizes through SEM analysis.

3.2. Structural Properties

The X-ray diffractograms were refined by applying the Rietveld method, drawing from the Pearson crystal structure library and using the Maud program, as illustrated in Figure 2. This technique involves fitting the calculated diffraction pattern as closely as possible to the measured data, which allowed for identifying phases within the examined powder and determining their mass proportions and lattice parameters.
The analysis of the X-ray diffraction pattern of the ZnLaFeO4 sample reveals a complex composition comprising three main phases: LaFeO3 (54.88%), La2O3 (33.23%), and ZnFe2O4 (11.89%). The predominance of LaFeO3 suggests a strong reaction between lanthanum (La) and iron (Fe) during the preparation of the sample. However, the presence of La2O3 and ZnFe2O4 phases indicates an incomplete reaction, due to the uneven distribution of elements, which prevented the full formation of LaFeO3.
In their study, Chaudhary et al. [55] observed the formation of an ortho-ferrite phase. Similarly, Angari et al. [56] noticed that the characteristic peaks corresponding to the LaFeO3 phase became progressively more pronounced as the concentration of La ions increased in the NiFe2O4 spinel ferrite. This observation suggests that La cations did not easily form a solid solution with the spinel ferrite, likely due to the limited solubility of La ions or their incorporation into the spinel lattice. Indeed, at high dopant concentrations (La3+), only a small fraction of La3+ cations were substituted into the spinel lattice, while the excess cations accumulated at the grain boundaries, resulting in the formation of the orthoferrite LaFeO3 phase [57,58,59].
The average crystallite size D of the lanthanum-doped Zn-ferrite was determined using two techniques: the Debye–Scherrer method and the Williamson–Hall method were employed to calculate the crystallite size by applying the following formula:
D = k λ β c o s θ
where k is Scherrer’s constant, k = 0.9; θ (radian) denotes Bragg’s angle; λ represents the wavelength of the radiation CoKα; and β (radian) designates the half-height width of the peaks.
The Scherrer method assumes the isotropy of the crystallites, by attributing the broadening of diffraction peaks solely to the crystallite size, while neglecting significant microstructural characteristics such as crystallographic defects or stress-induced distortions. In contrast, the Williamson–Hall method offers a more comprehensive analysis by considering that the broadening of diffraction peaks results from a combination of factors, including both the crystallite size and the micro-strain (crystallographic defects). The Williamson–Hall equation is written as follows:
β cos θ = k λ D + 4 ε sin θ
The crystallite size can be determined using the intercept and slope of the linear fit from the plotted curve β   cos ( θ ) = f ( 4 sin ( θ ) ) .
Figure 3 represents the Williamson–Hall diagrams of the ZnLaFeO4 compound.
Table 1 presents the crystallographic details of the powders determined by Rietveld refinement analysis, including lattice parameters, phase composition, and crystallite size. The latter was measured using two different methods.

3.3. Cycling Properties of the ZnLaFeO4 Electrode Using the Galvanostatic Method

3.3.1. Activation Process of the ZnLaFeO4 Electrode

The alloy activation during electrochemical cycling is mainly determined by its maximum discharge capacity, low electrode polarization, and stable half-discharge potential. It worth noting that some compounds activate after the first cycle, while others may require several cycles to achieve full activation.
Figure 4 shows the evolution of the discharge potential as a function of the electrochemical capacity variation in the ZnLaFeO4 electrode during the initial activation cycles in a 6 M KOH electrolyte at a C/10 rate.
The experimental results indicate that the substitution of lanthanum into the spinel structure significantly influenced the electrode’s activation process. The later was slowed by the electrode’s low cyclability and was only achieved after 46 cycles, with the maximum discharge capacity reaching approximately 106 mAh/g. Zayani et al. [51] systematically investigated the trivalent cation substitution in sol–gel-synthesized ZnFe2O4, revealing a strong correlation between Sm3⁺ substitution levels on the A-site of the spinel ZnFe2O4 and electrode activation kinetics. Their work demonstrated that higher Sm3+ doping concentrations progressively hindered the activation process. In case of high-rate substitution of La (another trivalent cation) for Fe in ZnFe2O4, excess La ions are partially incorporated into the spinel structure and interact with Fe, resulting in the formation of LaFeO3 at the grain boundaries. This phenomena created a thin insulating layer around the grains, which introduced additional resistance to the charge transfer process.
To assess the reversibility of the hydrogen absorption/desorption reactions, the polarization parameter, which is the difference between the half-charge and half-discharge potentials during electrochemical cycling, was introduced This parameter can be expressed using the following equation:
E = E 1 / 2   c h a r g e E 1 / 2   d i s c h a r g e
Figure 5 illustrates the evolution of the half-charge/half-discharge potentials as a function of number of cycles and the polarization curve for the ZnLaFeO4 electrode.
During the first cycle before activation, the half-discharge potential of the electrode shifted to less positive values, impeding the insertion of hydrogen into the interstitial sites of the electrode. As the activation cycle progressed, the half-discharge potential became more positive, facilitating hydrogen insertion into the electrode. Initially, the polarization values varied between approximately 663 and 817 mV. This value gradually decreased to 99 mV at the activation cycle before being stabilized during the subsequent cycles. These results are consistent with the electrochemical discharge capacity data, suggesting that the hydrogen absorption/desorption reactions at the sites became more reversible during the activation cycle.

3.3.2. Cycling Properties of the ZnLaFeO4 Electrode

Figure 6 shows the evolution of the electrochemical discharge capacity as a function of the number of cycles of the ZnLaFeO4 negative electrode.
The specific discharge capacity showed a progressive increase throughout the activation cycles, reaching a maximum value of 106 mAh·g−1, as illustrated in Figure 6. While this value remains lower than that of certain oxide-based materials such as ZnFe2O4 (145 mAh·g−1) [50] and intermetallic compounds like LaNi5, it still outperforms several conventional hydride materials, Ti2−xZrxNi (102 mAh·g−1) [60]. This demonstrates the promising balance between performance and stability offered by the La-substituted ZnLaFeO4.

3.3.3. Electrode Degradation Process

After several charge/discharge cycles, a decrease in capacity was commonly observed in most studied electrodes. This decrease can be explained by the first-order degradation kinetics, used to calculate the degradation rate ( K d e g ) . This parameter provides information about the stability of the electrochemical discharge capacity as a function of the number of cycles.
The discharge capacity ( Q d i c h , N ) and the concentration of the active material in the electrode are directly correlated with first-order degradation kinetics. This can be expressed as the following relationship:
Q d i c h , N = Q 0 exp ( K d e g   N )
l o g Q d i c h , N = l o g Q 0 0.434 exp K d e g   N
where Q0 represents the maximum capacity (mAhg−1), while K d e g denotes the degradation constant of the hydride material at a specific charge/discharge rate. This constant was derived from the slope of the linear part of the curve log(Q) = f(N), similarly to a first-order kinetic rate constant.
The following formula can be used to express the deterioration rate ( K d e g ), which is directly proportional to the corrosion rate ( r c o r r ):
r c o r r = K d e g C M
Figure 7 illustrates the dependence of l o g Q d i c h , N on the number of cycles, showing the capacity degradation curve of the negative electrode.
The maximum discharge capacity (Qdisch,max), the capacity at the 80th cycle (Q80), the retained discharge capacity (R80(%)), as well as the hydrogenation and corrosion characteristics obtained from the capacity fade curve are among the distinctive parameters summarized in Table 2.
It was noted that, after activation, the discharge capacity decreases with cycling, primarily due to the oxidation of the active surface. According to Pan et al. [61], the interactions of the working electrode with the electrolyte lead to the deterioration of the negative electrode material, which results in the decrease in the discharge capacity. The formation of corrosive byproducts, including iron (Fe) deposits that accumulate on the surface of the alloy, can be linked to this loss of capacity. It was also observed that the overall performance of the electrode was reduced because of the existence of these deposits, which drastically minimized the active surface area accessible to the hydrogen absorption.
Various approaches were examined in previous studies to minimize the degradation rates. Indeed, Dymek et al. [62] investigated the degradation mechanisms of a La1.5Mg0.5Ni6.5Co0.5 electrode coated with amorphous nickel. The obtained results showed that the nickel coating considerably reduced the deterioration of the electrode, most likely by acting as a barrier to prevent corrosion and mechanical wear. In a similar aim, Bala et al. [63] focused on the efficacy of encasing electrodes in Ni-P coatings to improve the durability and minimize the deterioration. Another study’s experimental results showed a significant decrease in the degradation with increasing Mn substitution levels, thereby enhancing structural stability and corrosion resistance, which consequently prolonged the electrode’s lifetime and improved its cycling performance. The above-stated research works demonstrated the important role of surface alterations and material substitution techniques in reducing the deterioration of the electrode used in electrochemical applications.
SEM micrographs obtained after cycling are shown in Figure 8. A comparative SEM analysis was carried out between ZnFe2O4 and La-doped ZnLaFe2O4 after electrochemical cycling. SEM micrographs highlight marked morphological differences between the ZnFe2O4 and ZnLaFe2O4 samples, resulting from lanthanum incorporation.
The ZnFe2O4 sample features a compact, dense microstructure composed of strongly agglomerated micrometric grains interspersed with cracks attributable to thermomechanical stresses. Its homogeneous, low-porosity surface is characteristic of a material that has undergone high-temperature sintering or calcination.
Conversely, lanthanum doping ZnLaFe2O4 leads to a nanostructured, porous morphology, where quasi-spherical particles are irregularly dispersed. This transition to a finer, less agglomerated structure suggests that the La3+ ion disrupts crystal growth, probably due to its higher ionic radius than that of Fe3+, inducing lattice distortions and limiting grain coalescence.

3.3.4. Kinetic Parameters of the ZnLaFeO4 Negative Electrode

Galvanostatic charge/discharge curves are commonly employed by researchers to determine the electrochemical discharge capacity based on the duration of each discharge cycle. However, Henryk Bala et al. [64] introduced innovative methods to determine some kinetic parameters, such as the equilibrium potential and the exchange current density. Their approach involves simplifying all interfacial processes between the negative electrode and electrolyte during charging and discharging to the H2O/H2 system.
The equilibrium potential (EeqH2O/H2) can be determined by applying the following formula [65,66].
E H 2 O / H 2 e q = E ( i + ) + E ( i ) b l o g ( i c i a ) 2
If |ic| = ia, Equation (2) was simplified to represent the midpoint of the potential jump (ΔE), resulting in the following equation:
E ( i + ) + E ( i ) 2
The potentials E(i+) and E(i−) correspond to the start of the charge and the end of the discharge, respectively. The Tafel slope of the straight lines on both sides (cathodic and anodic) of the hydrogen electrode is represented by b, while icharge and idischarge are the charge and discharge currents applied between the working electrode and the auxiliary electrode, respectively.
Figure 9 illustrates the variation in the equilibrium potential for the H2O/H2 system as a function of the number of cycles for the ZnLaFeO4 negative electrode.
As shown in Figure 9, the equilibrium potential of the ZnLaFeO4 electrode declines gradually with the increase in the number charge/discharge cycles. During the initial cycles, the potential shifts towards more negative values. As the number of cycles increases, the equilibrium potential gradually turns towards more positive values, resulting in a value of −0.827 V at the end of the cycling process.
The obtained findings reveal that the ZnLaFeO4 electrode underwent an initial activation or modification, followed by a stabilization phase. Clearly, the performance and potential of the electrode were influenced by changes in its electrochemical properties or structure during cycling.

3.4. Redox Properties of the Electrode ZnLaFeO4 Studied by the Potentiodynamic Method

During electrochemical cycling, the ZnLaFeO4 negative electrode cycling was studied by determining voltammograms, I = f(E), using the linear voltammetry technique. Using the first Stern method and origin software, the Tafel polarization curves were drawn using the obtained voltammograms.
Figure 10 shows the progression of both experimental and theoretical Tafel curves for ZnLaFeO4 negative electrodes during cycling.
As illustrated in Figure 10, the Tafel polarization curves of the ZnLaFeO4 electrode exhibit minimal variation during the initial cycles, indicating stable electrochemical behavior. However, a distinct shift appears around the 46th activation cycle, with the curves moving toward more positive potential values. This change suggests that the activation process improves the electrode’s ability to facilitate hydrogen insertion, likely due to increased surface activity and a greater number of active sites. After this activation process, the Tafel curves gradually shift back toward less positive potential values, indicating a reduction in hydrogen insertion efficiency. This behavior may be attributed to a decline in surface reactivity or a partial saturation of active sites, which makes hydrogen incorporation more difficult.
Adjusting the theoretical Butler–Volmer equation to the Tafel polarization curves allowed for determining the redox variables, such as the current density and the Nernst potential [67].
Figure 11 depicts the progress of the Nernst potential (E0) as the ZnLaFeO4 electrode underwent multiple charge/discharge cycles.
It was observed that the Nernst potential of the electrode increased as the activation cycles of the electrode increased during the activation cycle and shifted toward less negative values. It reached its maximum value on the 40th cycle, which is attributed to the fast insertion of hydrogen into the electrodes, especially during the activation process. After this activation phase, the Nernst potential shifted to more negative values, indicating that hydrogen insertion became more difficult due to the deterioration of the active material through oxidation [51].
Figure 12 demonstrates the variation in the exchange current density I0 as a function of the number of cycles of the ZnLaFeO4 electrode.
As shown in Figure 12, during cycling, the exchange current density of the ZnLaFeO4 electrode shows the same evolution as the electrochemical discharge capacity. As the activation cycles proceeded, the exchange current density of the ZnLaFeO4 electrode increased and reached its highest values. This rise can be explained by the improved hydrogen adsorption occurring at the electrode–electrolyte interface during cycling. However, achieving the maximum value required more cycles because the LaFeO3 phase, which appeared at the grain boundaries, had poor intrinsic conductivity.
It was also noted that the current density of the ZnLaFeO4 electrode decreased due to corrosion after activation, which resulted in fewer interstitial sites available for hydrogen insertion.

4. Conclusions

Due to its simplicity and cost-effectiveness, the sol–gel method was used to synthesize the ZnLaFeO4 compound.
Structural analysis revealed a multiphase composition in the synthesized ZnLaFeO4 material, consisting of a spinel structure accompanied by LaFeO3 and La2O3 secondary phases. The electrochemical behavior of the electrode was investigated using techniques such as chronoamperometry and cyclic voltammetry. The activation process was relatively slow, with the maximum discharge capacity of 106 mAh/g achieved only after 46 cycles. This sluggish activation is likely due to structural rearrangements.
The gradual decrease in discharge capacity after the activation process was attributed to mechanical degradation of the electrode, particularly cracking and fracturing caused by volumetric expansion and contraction during charge/discharge cycles. These defects likely contributed to the loss or detachment of active material, thereby reducing the number of electrochemically accessible sites.
A strong correlation was observed between discharge capacity, polarization, and current density throughout cycling. The high La3+ concentration was found to destabilize the spinel structure, facilitating the emergence of the LaFeO3 perovskite phase. This structural transformation negatively affected both the integrity and the electrochemical performance of the electrode.
Compared to analogous compounds, the ZnLaFeO4 alloy developed in this study demonstrates a significantly enhanced electrochemical performance. Undoped ZnFe2O4, for instance, commonly suffers from a low discharge capacity and poor cycling stability due to its limited electrical conductivity and inherent structural limitations. Similarly, while LaFeO3 is known for its chemical stability, its perovskite structure is not well-suited to hydrogen storage, leading to a suboptimal performance in Ni-MH batteries. In contrast, the ZnLaFeO4 compound exhibits a maximum discharge capacity of 106 mAh/g along with stable cycling behavior. These improvements underscore the beneficial role of La doping, which appears to enhance charge transport and corrosion resistance within the spinel structure. Overall, ZnLaFeO4 emerges as a promising candidate for use as an anode material in Ni-MH battery systems.
Future work will focus on a more controlled and gradual substitution of La3+ into the spinel lattice, to investigate its effects on the structural stability and overall electrochemical performance of the material.

Author Contributions

Validation, N.F.; Formal analysis, Y.D.; Resources: J.L., Supervision J.L.; Writing—original draft, H.G.; Writing—review & editing, W.Z.; Supervision, Validation, review C.K., O.E. and N.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Dataset available on request from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gielen, D.; Boshell, F.; Saygin, D.; Bazilian, M.D.; Wagner, N.; Gorini, R. The role of renewable energy in the global energy transformation. Energy Rev. 2019, 24, 38–50. [Google Scholar] [CrossRef]
  2. Campanari, S.; Manzolini, G.; De la Iglesia, F.G. Energy analysis of electric vehicles using batteries or fuel cells through well-to-wheel driving cycle simulations. J. Power Sources 2009, 186, 464–477. [Google Scholar] [CrossRef]
  3. Smith, W.J. Can EV (electric vehicles) address Ireland’s CO2 emissions from transport? Energy 2010, 35, 4514–4521. [Google Scholar] [CrossRef]
  4. Sun, F.; Hu, X.; Zou, Y.; Li, S. Adaptive unscented Kalman filtering for state of charge estimation of a lithium-ion battery for electric vehicles. Energy 2011, 36, 3531–3540. [Google Scholar] [CrossRef]
  5. Johansson, B.; Mårtensson, A. Energy and environmental costs for electric vehicles using CO2-neutral electricity in Sweden. Energy 2000, 25, 777–792. [Google Scholar] [CrossRef]
  6. Chau, K.T.; Wu, K.C.; Chan, C.C. A new battery capacity indicator for lithium-ion battery powered electric vehicles using adaptive neuro-fuzzy inference system. Energy. Conver. Manag. 2004, 45, 1681–1692. [Google Scholar] [CrossRef]
  7. Rao, Z.; Wang, S. A review of power battery thermal energy management. Renew. Sustain. Energy Rev. 2011, 15, 4554–4571. [Google Scholar] [CrossRef]
  8. Kang, J.; Yan, F.; Zhang, P.; Du, C. Comparison of comprehensive properties of Ni-MH (nickel-metal hydride) and Li-ion (lithium-ion) batteries in terms of energy efficiency. Energy 2014, 70, 618–625. [Google Scholar] [CrossRef]
  9. Niu, H.; Zhang, N.; Lu, Y.; Zhang, Z.; Li, M.; Liu, J.; Zhang, N.; Song, W.; Zhao, Y.; Miao, Z. Strategies toward the development of high-energy-density lithium batteries. J. Energy Storage 2024, 88, 111666. [Google Scholar] [CrossRef]
  10. Fetcenko, M.A.; Ovshinsky, S.R.; Reichman, B.; Young, K.; Fierro, C.; Koch, J.; Zallen, A.; Mays, W.; Ouchi, T. Recent advances in NiMH battery technology. J. Power Sources 2007, 165, 544–551. [Google Scholar] [CrossRef]
  11. Guo, Y.; Wang, W.; Su, H.; Lu, H.; Li, Y.; Peng, Q.; Shumin, H.; Zhang, L. Insights into the effect of Y substitution on superlattice structure and electrochemical performance of A5B19-type La-Mg-Ni-based hydrogen storage alloy for nickel metal hydride battery. J. Mater. Sci. Technol. 2025, 207, 60–69. [Google Scholar] [CrossRef]
  12. Salighe, Z.; Arabi, H.; Ghorbani, S.; Komeili, M. Microstructural, hydrogenation and electrochemical properties of La2Mg1−xYxNi10Mn0.5 (x = 0.1, 0.38) alloys for Ni-MH battery anode. J. Solid. State Chem. 2025, 348, 125340. [Google Scholar] [CrossRef]
  13. Zhang, Q.; Bian, Z.; Liu, X.; Lan, X.; Liu, J.; Ma, Z.; Zhang, H.; Luo, Y. Effects of thickness and gas hydrogenation on the electrochemical performances of a-Si thin film as anode for Ni-MH battery. Int. J. Hydrogen Energy 2024, 82, 959–967. [Google Scholar] [CrossRef]
  14. Sun, W.; Sheng, P.; Zhang, X.; Sun, H.; Li, J.; Cao, Z.; Qi, Y.; Zhang, Y. Metal Hydride Electrodes Applied to Ni-MH Battery Using Mg-Y-Ni-Cu-Based Alloys. Energy Technol. 2025, 24, 02252. [Google Scholar] [CrossRef]
  15. Marins, A.A.; Boasquevisque, L.M.; Muri, E.J.; Freitas, M.B. Enviromentally friendly recycling of spent Ni–MH battery anodes and electrochemical characterization of nickel and rare earth oxides obtained by sol–gel synthesis. Mater. Chem. Phys. 2022, 280, 125821. [Google Scholar] [CrossRef]
  16. Zhang, Y.H.; Wei, X.; Gao, J.L.; Hu, F.; Qi, Y.; Zhao, D.L. Electrochemical hydrogen storage behaviors of as-milled Mg–Ti–Ni–Co–Al-based alloys applied to Ni-MH battery. Electrochim. Acta 2020, 342, 136123. [Google Scholar] [CrossRef]
  17. Kalita, G.; Otsuka, R.; Endo, T.; Furukawa, S. Recent advances in substitutional doping of AB5 and AB2 type hydrogen storage metal alloys for Ni-MH battery applications. J. Alloys Compd. 2025, 1020, 179352. [Google Scholar] [CrossRef]
  18. Ouyang, L.; Huang, J.; Wang, H.; Liu, J.; Zhu, M. Progress of hydrogen storage alloys for Ni-MH rechargeable power batteries in electric vehicles: A review. Mater. Chem. Phys. 2017, 200, 164–178. [Google Scholar] [CrossRef]
  19. Joubert, J.M.; Paul-Boncour, V.; Cuevas, F.; Zhang, J.; Latroche, M. LaNi5 related AB5 compounds: Structure, properties and applications. J. Alloys. Comp. 2021, 862, 158163. [Google Scholar] [CrossRef]
  20. Gamo, T.; Moriwaki, Y.; Yanagihara Yamashita, T.; Iwaki, T. Formation and properties of titanium-manganese alloy hydrides. Int. J. Hydrogen Energy 1985, 10, 39–47. [Google Scholar] [CrossRef]
  21. Chen, J.; Takeshita, H.T.; Tanaka, H.; Kuriyama, N.; Sakai, T.; Uehara, I.; Haruta, M. Hydriding properties of LaNi3 and CaNi3 and their substitutes with PuNi3-type structure. J. Alloy. Comp. 2000, 302, 304–313. [Google Scholar] [CrossRef]
  22. Cuevas, F.; Joubert, J.M.; Latroche, M.; Percheron-Guegan, A. Intermetallic compounds as negative electrodes of Ni/MH batteries. App. Phys. A 2001, 72, 225–238. [Google Scholar] [CrossRef]
  23. Arya, S.; Verma, S. Nickel-metal hydride (Ni-MH) batteries. Rechar. Batter. Histo. Prog. Appl. 2020, 131–175. [Google Scholar] [CrossRef]
  24. Jin, S.; Ren, K.; Liang, J.; Kong, J. A novel sheet perovskite type oxides LaFeO3 anode for nickel-metal hydride batteries. J. Mater. Sci. Technol. 2024, 195, 218–226. [Google Scholar] [CrossRef]
  25. Esaka, T.; Sakaguchi, H.; Kobayashi, S. Hydrogen storage in proton-conductive perovskite-type oxides and their application to nickel–hydrogen batteries. Solid State Ioni. 2004, 166, 351–357. [Google Scholar] [CrossRef]
  26. Stoumpos, C.C.; Kanatzidis, M.G. Halide Perovskites: Poor Man’s high-performance semiconductors. Adv. Mater. 2016, 28, 5778–5793. [Google Scholar] [CrossRef]
  27. Bini, M.; Ambrosetti, M.; Spada, D. ZnFe2O4, a green and high-capacity anode material for lithium-ion batteries: A review. Appl. Sci. 2021, 11, 11713. [Google Scholar] [CrossRef]
  28. Brabers, V.A.M. Progress in spinel ferrite research. Hand. Mater. 1995, 8, 189–324. [Google Scholar]
  29. Narang, S.B.; Pubby, K. Nickel spinel ferrites: A review. J. Magn. Magn. Mater. 2021, 519, 167163. [Google Scholar] [CrossRef]
  30. Thakur, D.; Latwal, M.; Singh, J.P.; KGupta, L.; Srivastava, R.C. Zinc ferrite nanoparticles and their biomedical applications. In Oxides for Medical Applications; Woodhead Publishing: Sawston, UK, 2023; pp. 233–255. [Google Scholar]
  31. Valenzuela, R. Novel applications of ferrites. Phys. Res. Int. 2012, 2012, 591839. [Google Scholar] [CrossRef]
  32. Yadav, R.S.; Kuřitka, I.; Vilcakova, J.; Urbánek, P.; Machovsky, M.; Masař, M.; Holek, M. Structural, magnetic, optical, dielectric, electrical and modulus spectroscopic characteristics of ZnFe2O4 spinel ferrite nanoparticles synthesized via honey-mediated sol-gel combustion method. J. Phys. Chem. Solids 2017, 110, 87–99. [Google Scholar] [CrossRef]
  33. Baykal, A.; Kasapoğlu, N.; Durmuş, Z.; Kavas, H.; Toprak, M.S.; Köseoğlu, Y. CTAB-Assisted Hydrothermal Synthesis and Magnetic Characterization of NixCo1−xFe2O4 Nanoparticles (x = 0.0, 0.6, 1.0). Turkish. J. Chem. 2009, 33, 33–45. [Google Scholar] [CrossRef]
  34. Selvan, R.K.; Krishnan, V.; Augustin, C.O.; Bertagnolli, H.; Kim, C.S.; Gedanken, A. Investigations on the structural, morphological, electrical, and magnetic properties of CuFe2O4−NiO nanocomposites. Chem. Mater. 2008, 20, 429–439. [Google Scholar] [CrossRef]
  35. Kefeni, K.K.; Mamba, B.B.; Msagati, T.A. Application of spinel ferrite nanoparticles in water and wastewater treatment: A review. Sep. Purif. Technol. 2017, 188, 399–422. [Google Scholar] [CrossRef]
  36. Šutka, A.; Gross, K.A. Spinel ferrite oxide semiconductor gas sensors. Sens. Actuators B Chem. 2016, 222, 95–105. [Google Scholar] [CrossRef]
  37. Lima-Tenório, M.K.; Tenório-Neto, E.T.; Hechenleitner, A.A.W.; Fessi, H.; Pineda, E.A.G. CoFe2O4 and ZnFe2O4 nanoparticles: An overview about structure, properties, synthesis and biomedical applications. J. Coll. Sci. 2016, 5, 45–54. [Google Scholar] [CrossRef]
  38. Amini, M.M.; Yadavi, M. Influence of metal core of mixed-metal carboxylates in preparation of spinel: ZnFe2O(O2CCF3)6 as a single-source precursor for preparation of ZnFe2O4. Appl. Organo. Chem. 2005, 19, 1164–1167. [Google Scholar] [CrossRef]
  39. Kalendova, A.; Veselý, D.; Brodinova, J. Anticorrosive spinel-type pigments of the mixed metal oxides compared to metal polyphosphates. Anti-Corr. Meth. Mater. 2004, 51, 6–17. [Google Scholar] [CrossRef]
  40. Dhineshbabu, N.R.; Vettumperumal, R.; Narendrakumar, A.; Manimala, M.; Kanna, R.R. Optical properties of lanthanum-doped copper spinel ferrites nanoparticles for optoelectronic applications. Adv. Sci. Eng. Med. 2017, 9, 377–383. [Google Scholar] [CrossRef]
  41. Mujahid, M.; Khan, R.U.; Mumtaz, M.; Soomro, S.A.; Ullah, S. NiFe2O4 nanoparticles/MWCNTs nanohybrid as anode material for lithium-ion battery. Ceram. Int. 2019, 45, 8486–8493. [Google Scholar] [CrossRef]
  42. Lavela, P.; Tirado, J.L. CoFe2O4 and NiFe2O4 synthesized by sol–gel procedures for their use as anode materials for Li ion batteries. J. Pow. Sou 2007, 172, 379–387. [Google Scholar] [CrossRef]
  43. Kumar, P.R.; Jung, Y.H.; Bharathi, K.K.; Lim, C.H.; Kim, D.K. High capacity and low cost spinel Fe3O4 for the Na-ion battery negative electrode materials. Electrochim. Acta 2014, 146, 503–510. [Google Scholar] [CrossRef]
  44. Shenoy, S.D.; Joy, P.A.; Anantharaman, M.R. Effect of mechanical milling on the structural, magnetic and dielectric properties of coprecipitated ultrafine zinc ferrite. J. Mag. Mag. Mater. 2004, 269, 217–226. [Google Scholar] [CrossRef]
  45. Yao, C.; Zeng, Q.; Goya, G.F.; Torres, T.; Liu, J.; Wu, H.; Ge, M.; Zeng, Y.; Wang, Y.; Jiang, J.Z. ZnFe2O4 nanocrystals: Synthesis and magnetic properties. J. Phys. Chem. 2007, 111, 12274–12278. [Google Scholar] [CrossRef]
  46. Rameshbabu, R.; Ramesh, R.; Kanagesan, S.; Karthigeyan, A.; Ponnusamy, S. Synthesis and study of structural, morphological and magnetic properties of ZnFe2O4 nanoparticles. J. Super. Nov. Mag. 2014, 27, 1499–1502. [Google Scholar] [CrossRef]
  47. Haija, M.A.; Abu-Hani, A.F.; Hamdan, N.; Stephen, S.; Ayesh, A.I. Characterization of H2S gas sensor based on CuFe2O4 nanoparticles. J. Alloy. Comp. 2017, 690, 461–468. [Google Scholar] [CrossRef]
  48. Soussi, A.; Haounati, R.; Taoufiq, M.; Baoubih, S.; Jellil, Z.; Elfanaoui, A.; Ihlal, A. Investigating structural, morphological, electronic, and optical properties of SnO2 and Al-doped SnO2: A combined DFT calculation and experimental study. Phys. B Condens. Matter 2024, 690, 416242. [Google Scholar] [CrossRef]
  49. Renuka, L.; Anantharaju, K.S.; Sharma, S.C.; Vidya, Y.S.; Nagaswarupa, H.P.; Prashantha, S.C.; Nagabhushana, H. Synthesis of ZnFe2O4 nanoparticle by combustion and sol gel methods and their structural, photoluminescence and photocatalytic performance. Mater. Today Proc. 2018, 5, 20819–20826. [Google Scholar] [CrossRef]
  50. Zayani, W.; Azizi, S.; El-Nasser, K.S.; Othman Ali, I.; Molière, M.; Fenineche, N.; Mathlouthi, H.; Lamloumi, J. Electrochemical behavior of a spinel zinc ferrite alloy obtained by a simple sol-gel route for Ni-MH battery applications. Int. J. Ener. Res. 2021, 45, 5235–5247. [Google Scholar] [CrossRef]
  51. Zayani, W.; Azizi, S.; El-Nasser, K.S.; Belgacem, Y.B.; Ali, I.O.; Fenineche, N.; Mathlouthi, H. New nanoparticles of (Sm, Zn)-codoped spinel ferrite as negative electrode in Ni/MH batteries with long-term and enhanced electrochemical performance. Int. J. Hydrogen Energy 2019, 44, 11303–11310. [Google Scholar] [CrossRef]
  52. Epp, J. X-ray diffraction (XRD) techniques for materials characterization. In Materials Characterization Using Nondestructive Evaluation (NDE) Methods; Woodhead Publishing: Sawston, UK, 2016; pp. 81–124. [Google Scholar] [CrossRef]
  53. Huang, L.W.; Elkedim, O.; Moutarlier, V. Synthesis and characterization of nanocrystalline Mg2Ni prepared by mechanical alloying: Effects of substitution of Mn for Ni. J. Alloy. Comp. 2010, 504, S311–S314. [Google Scholar] [CrossRef]
  54. Karaoud, I.; Dabaki, Y.; Khaldi, C.; ElKedim, O.; Fenineche, N.; Lamloumi, J. Electrochemical properties of the CaNi4.8M0.2 (M = Mg, Zn, and Mn) mechanical milling alloys used as anode materials in nickel-metal hydride batteries. Environ. Prog. Sustain. Energy 2023, 42, e14118. [Google Scholar] [CrossRef]
  55. Chaudhari, V.; Shirsath, S.E.; Mane, M.L.; Kadam, R.H.; Shelke, S.B.; Mane, D.R. Crystallographic, magnetic and electrical properties of Ni0.5Cu0.25Zn0.25LaxFe2−xO4 nanoparticles fabricated by sol–gel method. J. Alloy. Comp. 2013, 549, 213–220. [Google Scholar] [CrossRef]
  56. Al Angari, Y.M. Magnetic properties of La-substituted NiFe2O4 via egg-white precursor route. J. Mag. Mag. Mater. 2011, 323, 1835–1839. [Google Scholar] [CrossRef]
  57. Aslam, A.; Rehman, A.U.; Amin, N.; un Nabi, M.A.; ul ain Abdullah, Q.; Morley, N.A.; Mehmood, K. Lanthanum doped Zn0.5Co0.5LaxFe2−xO4 spinel ferrites synthesized via co-precipitation route to evaluate structural, vibrational, electrical, optical, dielectric, and thermoelectric properties. J. Phys. Chem. Solids 2021, 154, 110080. [Google Scholar] [CrossRef]
  58. Wang, Y.; Xu, F.; Li, L.; Liu, H.; Qiu, H.; Jiang, J. Magnetic properties of La-substituted Ni–Zn–Cr ferrites via rheological phase synthesis. Mater. Chem. Phys. 2008, 112, 769–773. [Google Scholar] [CrossRef]
  59. Ganure, K.A.; Dhale, L.A.; Katkar, V.T.; Lohar, K.S. Synthesis and characterization of lanthanum-doped Ni-Co-Zn spinel ferrites nanoparticles via normal micro-emulsion method. Int. J. Nanotechnol. Appl. 2017, 11, 189–195. [Google Scholar]
  60. Li, X.D.; Elkedim, O.; Nowak, M.; Jurczyk, M.; Chassagnon, R. Structural characterization and electrochemical hydrogen storage properties of Ti2−xZrxNi (x = 0, 0.1, 0.2) alloys prepared by mechanical alloying. Int. J. Hydrogen Energy 2013, 38, 12126–12132. [Google Scholar] [CrossRef]
  61. Pan, Y.H.; Srinivasan, V.; Wang, C.Y. An experimental and modeling study of isothermal charge/discharge behavior of commercial Ni–MH cells. J. Power. Sources 2002, 112, 298–306. [Google Scholar] [CrossRef]
  62. Dymek, M.; Nowak, M.; Jurczyk, M.; Bala, H. Encapsulation of La1.5Mg0.5Ni7 nanocrystalline hydrogen storage alloy with Ni coatings and its electrochemical characterization. J. Alloys. Comp. 2018, 749, 534–542. [Google Scholar] [CrossRef]
  63. Bala, H.; Dymek, M. Corrosion degradation of powder composite hydride electrodes in conditions of long-lasting cycling. Mater. Chem. Phys. 2015, 167, 265–270. [Google Scholar] [CrossRef]
  64. Bala, H.; Giza, K.; Kukula, I. Determination of hydrogenation ability and exchange current of H2O/H2 system on hydrogen-absorbing metal alloys. J. Appl. Electrochem. 2010, 40, 791–797. [Google Scholar] [CrossRef]
  65. Bala, H.; Dymek, M.; Adamczyk, L.; Giza, K.; Drulis, H. Hydrogen diffusivity, kinetics of H2O/H2 charge transfer and corrosion properties of LaNi5-powder, composite electrodes in 6 M KOH solution. J. Solid. State Electrochem. 2014, 18, 3039–3048. [Google Scholar] [CrossRef]
  66. Bala, H.; Kukuła, I.; Giza, K.; Marciniak, B.; Różycka-Sokołowska, E.; Drulis, H. Evaluation of electrochemical hydrogenation and corrosion behavior of LaNi5-based materials using galvanostatic charge/discharge measurements. Int. J. Hydrogen Energy 2012, 37, 16817–16822. [Google Scholar] [CrossRef]
  67. Ren, K.; Miao, J.; Shen, W.; Su, H.; Pan, Y.; Zhao, J.; Han, S. High temperature electrochemical discharge performance of LaFeO3 coated with C/Ni as anode material for NiMH batteries. Prog. Nat. Sci. Mat. Int. 2022, 32, 684–692. [Google Scholar] [CrossRef]
Figure 1. SEM micrograph of ZnLaFeO4.
Figure 1. SEM micrograph of ZnLaFeO4.
Energies 18 03251 g001
Figure 2. X-ray diffractograms refined by the Rietveld method of ZnLaFeO4 powder synthesized by the sol–gel method calcined at 900 °C. The black circles represent the experimental data, the red line is the fit by Rietveld refinement, and the lower curve shows the difference between the observed and calculated patterns. Vertical markers correspond to the Bragg positions of LaFeO3 (red), La2O3 (blue) and ZnFe2O4 (green).
Figure 2. X-ray diffractograms refined by the Rietveld method of ZnLaFeO4 powder synthesized by the sol–gel method calcined at 900 °C. The black circles represent the experimental data, the red line is the fit by Rietveld refinement, and the lower curve shows the difference between the observed and calculated patterns. Vertical markers correspond to the Bragg positions of LaFeO3 (red), La2O3 (blue) and ZnFe2O4 (green).
Energies 18 03251 g002
Figure 3. The Williamson–Hall diagrams of the ZnLaFeO4 compound.
Figure 3. The Williamson–Hall diagrams of the ZnLaFeO4 compound.
Energies 18 03251 g003
Figure 4. Evolution of the discharge potential curves as a function of the variation in the discharge capacity of the ZnLaFeO4 electrode during the first activation cycles, and at a discharge current of 0.8 mA.
Figure 4. Evolution of the discharge potential curves as a function of the variation in the discharge capacity of the ZnLaFeO4 electrode during the first activation cycles, and at a discharge current of 0.8 mA.
Energies 18 03251 g004
Figure 5. (A) Half-charge and half-discharge potentials vs. the number of cycles, and (B) polarization curve for the ZnLaFeO4 electrode determined under a discharge current of 0.8 mA.
Figure 5. (A) Half-charge and half-discharge potentials vs. the number of cycles, and (B) polarization curve for the ZnLaFeO4 electrode determined under a discharge current of 0.8 mA.
Energies 18 03251 g005
Figure 6. Evolution of the discharge capacity as a function of the variation in the number of cycles of the ZnLaFeO4 electrode.
Figure 6. Evolution of the discharge capacity as a function of the variation in the number of cycles of the ZnLaFeO4 electrode.
Energies 18 03251 g006
Figure 7. Capacity fade of the ZnLaFeO4 electrode.
Figure 7. Capacity fade of the ZnLaFeO4 electrode.
Energies 18 03251 g007
Figure 8. SEM images of the ZnLaFeO4 (a) and the undoped ZnFe2O4 (b) [50] electrode surface after cycling.
Figure 8. SEM images of the ZnLaFeO4 (a) and the undoped ZnFe2O4 (b) [50] electrode surface after cycling.
Energies 18 03251 g008
Figure 9. Evolution of the equilibrium potential of the H2O/H2 system versus cycling, associated with the ZnLaFeO4 negative electrode.
Figure 9. Evolution of the equilibrium potential of the H2O/H2 system versus cycling, associated with the ZnLaFeO4 negative electrode.
Energies 18 03251 g009
Figure 10. Theoretical and experimental Tafel curves of the ZnLaFeO4 electrodes during cycling.
Figure 10. Theoretical and experimental Tafel curves of the ZnLaFeO4 electrodes during cycling.
Energies 18 03251 g010
Figure 11. Evolution of the E0 of the ZnLaFeO4 electrode during cycling.
Figure 11. Evolution of the E0 of the ZnLaFeO4 electrode during cycling.
Energies 18 03251 g011
Figure 12. Change in the exchange current density of the ZnLaFeO4 electrode during cycling.
Figure 12. Change in the exchange current density of the ZnLaFeO4 electrode during cycling.
Energies 18 03251 g012
Table 1. Crystallographic parameters obtained by Rietveld refinement of the powders and the average crystallite sizes obtained by the two methods.
Table 1. Crystallographic parameters obtained by Rietveld refinement of the powders and the average crystallite sizes obtained by the two methods.
SamplePhases FoundLattice Parameters (Å)Abundance of Phase (wt%)Crystallite Size
Ds (nm)
Crystallite Size
DW-H (nm)
SigFit Parameters
ZnLaFeO4LaFeO3a = 5.562
b = 7.841
c = 5.549
54.8836.4554.321.8Rwp = 6.11;
Rb = 5.77
Rexp = 5.36
La2O3a = 3.956
c = 6.137
33.23
ZnFe2O4a = 8.43911.89
Table 2. Degradation parameters derived from the capacity fade curve.
Table 2. Degradation parameters derived from the capacity fade curve.
MaterialNmaxQdisch,max
(mAh/g)
Qdich80
(mAh/g)
R80
(%)
l o g Q d i s c h N K
(cycle−1)
CM
(g·cm−3)
rcorr
(g·cm−3·cycle−1)
ZnLaFeO4461067469.81−0.004950.011402.0280.02311
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

Gharbi, H.; Zayani, W.; Dabaki, Y.; Khaldi, C.; ElKedim, O.; Fenineche, N.; Lamloumi, J. Structure and Electrochemical Behavior of ZnLaFeO4 Alloy as a Negative Electrode in Ni-MH Batteries. Energies 2025, 18, 3251. https://doi.org/10.3390/en18133251

AMA Style

Gharbi H, Zayani W, Dabaki Y, Khaldi C, ElKedim O, Fenineche N, Lamloumi J. Structure and Electrochemical Behavior of ZnLaFeO4 Alloy as a Negative Electrode in Ni-MH Batteries. Energies. 2025; 18(13):3251. https://doi.org/10.3390/en18133251

Chicago/Turabian Style

Gharbi, Houyem, Wissem Zayani, Youssef Dabaki, Chokri Khaldi, Omar ElKedim, Nouredine Fenineche, and Jilani Lamloumi. 2025. "Structure and Electrochemical Behavior of ZnLaFeO4 Alloy as a Negative Electrode in Ni-MH Batteries" Energies 18, no. 13: 3251. https://doi.org/10.3390/en18133251

APA Style

Gharbi, H., Zayani, W., Dabaki, Y., Khaldi, C., ElKedim, O., Fenineche, N., & Lamloumi, J. (2025). Structure and Electrochemical Behavior of ZnLaFeO4 Alloy as a Negative Electrode in Ni-MH Batteries. Energies, 18(13), 3251. https://doi.org/10.3390/en18133251

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

Article Metrics

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