7.9
CiteScore
 
3.6
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
ABUNDANCE ESTIMATION IN AN ARID ENVIRONMENT
Case Study
Correspondence
Corrigendum
Editorial
Full Length Article
Invited review
Letter to the Editor
Original Article
Research Article
Retraction notice
REVIEW
Review Article
SHORT COMMUNICATION
Short review
7.2
CiteScore
3.7
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
ABUNDANCE ESTIMATION IN AN ARID ENVIRONMENT
Case Study
Correspondence
Corrigendum
Editorial
Full Length Article
Invited review
Letter to the Editor
Original Article
Research Article
Retraction notice
REVIEW
Review Article
SHORT COMMUNICATION
Short review
View/Download PDF

Translate this page into:

Research Article
2025
:37;
1232024
doi:
10.25259/JKSUS_123_2024

Synthesis and characterization of Sr-doped Ce/Mn nanocomposites for fuel cell applications

Department of Physics, Government college University, Kotwali Rd, Gurunanakpura, Faisalabasd, 38000, Pakistan
Department of Physics, King Saud University, PO Box 2455, 11451, Riyadh, Saudi Arabia
Department of Physics, Florida Agriculture and Mechanical University, Talllahassee, 32309, United States

* Corresponding authors E-mail addresses: tanveer.bukhari@yahoo.com (TH Bokhari), muhammad1.aslam@famu.edu (MJ Aslam)

Licence
This is an open-access article distributed under the terms of the Creative Commons Attribution-Non Commercial-Share Alike 4.0 License, which allows others to remix, transform, and build upon the work non-commercially, as long as the author is credited and the new creations are licensed under the identical terms.

Abstract

In this study, the feasibility of strontium (Sr)-doped cerium/manganese (Ce/Mn) nanocomposites as potential electrolyte materials for solid oxide fuel cells (SOFCs) has been checked. This study aims to improve ionic conductivity and catalytic activity, essential for the overall performance of fuel cells. The Sr-doped Ce/Mn nanocomposites were synthesized using a sol-gel method followed by thermal treatment. Structural investigation was done by X-ray diffraction (XRD) analysis, crystalline size synthesized samples was recorded as approximately 54.08 nm. The grown sample size was compatible and less than 100 nm. Evidence shows particle size has an important influence on the various high-tech uses. Morphological analysis shows the distinguishable grain boundaries on the samples and a similar size across the surface. These grain boundaries impede electrical conduction, and the energy band gap of the prepared material, which was investigated by using UV-Visible spectroscopy. The synthesized material’s band gap decreased steadily as its composition grew, which is characteristic of semiconducting materials. An electrochemical workstation and current-voltage (I-V) measurements show that the conductivity of the synthesized samples enhances with the steady increase in concentration of Ce and Mn dopants in the sample. The performance of the electrolyte material’s specific capacitance of all synthetic samples was determined to be (382.39, 436.97, and 500.312) F/g. The Csp value was greatest for the sample having the highest concentration of dopant.

GRAPHICAL ABSTRACT

Highlights of research work

  • Synthesis of SrxCe1-xMny (x= 0.03, 0.05 and 0.07, y=1) nanocomposites were synthesized by using sol-gel method.

  • Metal-doped Ce/Mn nanocomposites enhanced the storage capacity and electrical conductivity.

Keywords

(Ce/Mn) nanocomposites
Fuel cells
ionic conductivity and catalytic activity
UV-Visible spectroscopy
XRD analysis

1. Introduction

Worldwide research and development efforts have focused on solid oxide fuel cells (SOFCs) because of rising demand for energy and concern for the environment (Dogdibegovic et al., 2019; Hao et al., 2017; Gan et al., 2019; He et al., 2017). Researchers have experienced economic challenges in lowering standard SOFC operating temperature, especially below 600°C, due to the device’s poor ionic conductivity in recent years (Li et al., 2019; Gu et al., 2018). However, the economic viability of SOFCs was delayed due to problems with high temperatures that are linked to high costs (Preethi et al. 2019). As a result, global research and development projects focus on producing outstanding SOFCs that operate at temperatures below 600°C (Hou et al., 2019). More significant than their role in the ionic conduction mechanism is the oxide phase materials’ catalytic activity (Kim et al. 2019; Rocha et al. 2019).

A fuel cell is an electrochemical energy conversion technology that has great potential for producing clean, efficient electricity while also having a significant positive impact on the environment (Lohmann et al., 2017). Despite traditional heat engines that recognize problems with lubrication, leakage, and heat loss, SOFC technology is unrestricted. The advantages include reduced greenhouse gas emissions, water management, fuel variety, high energy conversion efficiency, and constant quality (Ali et al., 2016; Fergus et al., 2016). The ceria-based rare earth doped materials have been studied as a potential electrolyte for IT-SOFCs due to their capacity to provide high ionic conductivity at lower temperatures (Ahmad et al. 2017). The ceria doped with Gd, Sm, and Y, which had a high dopant concentration, exhibited improved ionic conductivity at higher temperatures (Momin et al. 2022).

The alkaline earth-doped ceria materials are supposed to be much more efficient in giving extremely high electrical conductivity at lower temperatures. The performance of nano-sized La0.4Sr0.6Co0.8Fe0.2O3 cathodes was examined utilizing the assisted ammonium combustion process, and it was found that this material is an inexpensive cathode for IT-SOFCs, with the maximum peak power density of 1.23 W cm-2 at 650°C (Mogni et al., 2015). Moreover, SmBaCo2-xNixO5 (x=0 to 0.5), the electrical conductivities of SBCNx decrease as the Ni content increases at a given temperature, while SBCO has a maximum conductivity of 1091 Scm-1 at 250°C (Xia et al., 2016). The Ce is gaining a lot of attention due to its increased ionic conductivity and morphological changes, which could be useful in solid oxide fuel cells (Coles-Aldridge et al., 2018).

The remarkable oxide ion conductivity found in doped ceria makes it possible to lower the working temperature of SOFCs, which eliminates various technological challenges (Babu et al., 2016). These days, the composite effect is also used to newly developed, cutting-edge Low Temperature Solid Oxide Fuel Cell (LTSOFC) electrolyte materials. Multi-doped or composite electrolytes have many contact areas between the two constituent phases as opposed to single-doped electrolytes (SDC, YSZ) (Artini et al., 2018). In comparison to the bulk, the interface offers an excellent conductivity channel for ionic conduction, which can disperse mobile ion concentration (Coduri et al., 2018).

The proton-conducting perovskite electrolytes unquestionably have the benefit of allowing for low operating temperatures. Unfortunately, a significant obstacle they must overcome is high grain boundary resistance that results in limited ionic conductivity (Souza et al., 2010; Iguchi et al., 2010). The doped barium cerate is an example of a proton conductor with limited proton conductivity (Barison et al., 2008). Moreover, perovskites with semiconducting properties have received interest as semiconductor materials that are employed as the electrolyte membrane in LT-SOFCs (Zhou et al., 2016; Xing et al., 2019; Zhu et al., 2018). Designing semiconductor-based high ion-conducting electrolytes, particularly perovskite oxides, with the aforementioned advantages in mind may have great practical promise for LT-SOFC applications.

The nanosized SrxCe1-xMny nanocomposite was synthesized via a sol-gel process, with x = 0.03, 0.05, 0.07, and y = 1. The electrochemical properties of the produced material were investigated using conductivity measurements and fuel cell performance tests, while its structure was examined using X-ray diffraction (XRD) and SEM.

2. Materials and Methods

2.1 Chemicals and reagents

Strontium nitrate Sr(NO3)2 (99.9% pure), cerium nitrate hexahydrate Ce(NO3)3.6H2O (99.9% pure), manganese nitrate hexahydrate Mn(NO3)2.4H2O (99.9% pure), citric acid, ethanol, ammonia solution, and deionized water of analytical grade were purchased from Sigma Aldrich and used for synthesis process without any further reaction/purification.

2.2 Synthesis of Sr doped Ce-Mn nanocomposites

The SrxCe1-xMny (x= 0.03, 0.05, and 0.07, y=1) nanocomposites were synthesized using the sol-gel method. The prepared samples were named SCM1, SCM2, and SCM3 for x = 0.03, 0.05, and 0.07, respectively. The strontium nitrate [Sr(NO3)2], cerium nitrate hexahydrate [Ce(NO3)3.6H2O], manganese nitrate hexahydrate [Mn(NO3)3.4H2O], and citric acid were used as precursors. The stoichiometric amount of the material was dissolved in deionized water. This reaction mixture was subjected to ultrasonication to homogenize the solution. The mixed solution was continuously stirred and heated at 80°C until the formation of a gel. The pH value was maintained around 7 by the addition of NH4OH solution. As the final step, the whole mixture was constantly stirred at 120°C to allow for the sol-gel auto-combustion process to obtain the final product. The powder product was collected, ground, and calcined at 800°C for 5 h to release the remaining nitrates and organics gel (Fig. 1). For further investigation, scanning electron microscopy (SEM) and XRD were employed for the desired crystalline structure and surface morphological analysis (Ahmad et al. 2017).

Schematic diagram for the synthesis of Sr-doped Ce/Mn Nanocomposites.
Fig. 1.
Schematic diagram for the synthesis of Sr-doped Ce/Mn Nanocomposites.

2.3 Fuel cell construction and performance

The cathode, anode, and solid electrolyte materials are the basic parts of a fuel cell. The electrolyte, which is the central part of a fuel cell, plays a crucial function in a solid oxide fuel cell as an electrode, allowing oxide ions to migrate via oxygen vacancies. The materials must have high conductivity of oxide ions and high density as well to avoid gas mixing and low electronic conductivity (Ferkhi et al., 2016). For understanding of single-cell performance, the power density of the membrane plays an important role in cell performance. The power density must be calculated via a particular voltage, RH, and temperature. However, in a reported fuel cell study, the overall calculated performance density at high temperature was recorded comparatively high, which is costly and hard to achieve the required temperature in portable devices (Wang et al., 2024). Further details have been shown in the given schematic 1.

Working diagram of solid oxide fuel cell
Schematic 1.
Working diagram of solid oxide fuel cell

The basic principle of a fuel cell depends upon the galvanostatic mechanism, as shown in the schematic diagram. Two types of conductivities are involved in a solid oxide fuel cell. Electronic conductivity and ionic conductivity are involved in a fuel cell. Self-explained fuel cell schematic depicts the open merit process via electrolyte and electronic conduction (Hao et al., 2017). Many young researchers focusing on the efficacy of fuel cells are facing challenges as well. One basic/crucial challenge in a solid oxide fuel cell is the achievement of a high desired temperature of 700-750°C. The second crucial challenge of the fuel cell is the efficacy of the fuel cell. Till now, 68.6% of fuel cell efficacy has been reported (Ferkhi et al., 2016).

3. Results and Discussion

3.1 Structural analysis

The phase and crystallinity of the deposited form of Sr-doped Ce/Mn Nano-composites determine the dependency of the working efficiency of the fuel cell. In this part of the experimental analytical approach, the structural analysis of Sr-doped Ce/Mn nanocomposites has been performed, as depicted in Fig. 2.

XRD graph of 3% Sr-doped, 5% Sr-doped, and 7% Sr-doped with CeMnO3.
Fig. 2.
XRD graph of 3% Sr-doped, 5% Sr-doped, and 7% Sr-doped with CeMnO3.

The crystal structure of CeMnO3 nanoparticles was identified by XRD patterns, which have been presented in Fig. 2. Graphs of 3% Sr-doped, 5% Sr-doped, and 7% Sr-doped nanomaterials synthesized by the sol-gel method, namely SCM-3, SCM-5, and SCM-7, respectively. Well-matched XRD peaks of the Cubic crystal system, Fm-3m Space group confirmed using JCPDS# 00-034-0294 and Miller indices (111), (200), (220), (311), (400), and (331), respectively. Some other peaks, 29.77 and 30.92 of the Orthorhombic Phase of Mn8O16, confirmed by using JCPDS# 96-210-5816 and miller indices (040) and (026), respectively (Jianshen et al., 2024). XRD pattern confirmed the successful synthesis material and shows a decreasing trend in intensities by increasing doping concentration. In addition, a decreasing trend in crystallite size was also recorded by adopting 3% Sr-doped, 5% Sr-doped, and 7% Sr-doped nanomaterials with Ce/Mn composites. Crystallite size of Sr-doped Ce/Mn composites has been calculated by using the Scherrer formula, and the value of the shape factor is 0.90. Values of Crystallite size, Miller indices (hkl), and d-spacing at the high intensity peak have been shown in Table 1. One diffraction peak of CeO2 at (111) existence of CeO2 in CeMnO3 nanoparticles. After doping Sr, the intensities become weaker and broader. Hu et al. (2019) reported that some secondary phase at * marked peaks of Mn2O3 are visible. According to J. Wang et al (2024), Weaker and border peaks were observed low crystallinity caused by smaller crystal size and more structural defects (Yue et al., 2019). Barelli et al. (2022) observed this behavior in SEM results.

Table 1. XRD graph of 3% Sr-doped, 5% Sr-doped, and 7% Sr-doped.
No. Angel (2θ) FWHM (β) Crystal size (D) Dislocation density (δ)
SCM-3 28.54 0.2952 27.77 0.0013
SCM-3 29.77 0.4329 18.99 0.0028
SCM-3 30.92 0.3936 20.94 0.0023
SCM-3 33.10 0.3936 21.06 0.0023
SCM-3 47.49 0.6297 13.78 0.0053
SCM-3 56.36 0.4723 19.08 0.0027
Crystal size 20.27
SCM-5 28.53 0.3149 26.03 0.0015
SCM-5 33.09 0.6298 13.16 0.0058
SCM-5 47.28 0.7085 12.24 0.0067
SCM-5 56.27 0.6298 14.30 0.0049
Crystal size 16.43
SCM-7 28.53 0.3542 23.14 0.0019
SCM-7 33.09 0.7084 11.70 0.0073
SCM-7 47.28 0.6297 13.77 0.0053
SCM-7 56.27 0.3936 22.89 0.0019
Crystal size 17.88

3.2 Analysis of surface morphology

The surface morphology of synthesized SrxCe1-xMny (y=1) samples with varying strontium (Sr) concentrations (x=0.03, 0.05, 0.07) was investigated using SEM at magnifications of 25,000× and 50,000×, as presented in Fig. 3. For the sample with x=0.03 (images 1a and 1b), the microstructure exhibits agglomerated particles with relatively large grain clusters and irregular morphologies. The observed particles appear densely packed, indicating moderate porosity. A heterogeneous distribution of particle sizes is evident, with average grain dimensions in the submicron range. The compact morphology may hinder electrolyte penetration but suggests good particle connectivity, beneficial for electrical conductivity.

1a, 1b.SEM images of prepared sample SrxCe1-xMny (x=0.03, y=1), 2a, 2b. SEM images of prepared sample SrxCe1-xMny (x=0.05, y=1), 3a, 3b. SEM images of the prepared sample SrxCe1-xMny (x=0.07, y=1) at different scales.
Fig. 3. (a)
1a, 1b.SEM images of prepared sample SrxCe1-xMny (x=0.03, y=1), 2a, 2b. SEM images of prepared sample SrxCe1-xMny (x=0.05, y=1), 3a, 3b. SEM images of the prepared sample SrxCe1-xMny (x=0.07, y=1) at different scales.
The conductivity behavior of prepared samples (SrxCe1-xMny, where x= 0.03, 0.05, and 0.07, y=1).
Fig. 3. (b)
The conductivity behavior of prepared samples (SrxCe1-xMny, where x= 0.03, 0.05, and 0.07, y=1).

As the Sr concentration increases to x=0.05 (images 2a and 2b), the morphology undergoes noticeable transformation. The particles become more refined and less agglomerated compared to the x=0.03 sample. Smaller grains are well distributed with slightly increased surface roughness, indicating a higher degree of surface area exposure. The reduction in agglomeration and increase in nano-structuring may enhance active surface sites, favoring applications in catalysis or electrochemical systems.

For their highest doping level, x=0.07 (image 3a and 3b), the sample exhibits a more porous structure with well-developed granular morphology. The particles are loosely packed and display a higher degree of porosity and fragmentation. Sharp-edged grains are observed alongside more spherical particles, possibly due to varied nucleation and growth mechanisms at higher Sr content. The morphology can significantly improve electrolyte diffusion and ion transport, indicating the potential for improved performance in energy storage or sensing applications.

Overall, the SEM analysis reveals that increasing SR contents in the SrxCe1-xMny system leads to reduced particle agglomeration and enhanced porosity. The morphological evolution from a compact porous nanostructure as x increases from 0.03 to 0.07 suggests tunable microstructural features that can be tailored for specific functional applications.

According to the results of the SEM analysis, the structure of the grains and their grain boundaries in the sample are not distinguishable and do not have a similar size across the surface. These boundaries may play a particular role in determining the nature of electrical conduction; certainly, the grain boundaries may impede the flow of charges within the material and, consequently, the properties of the material. Moreover, the percentage yield of the sample and the quantitative determination of the elements present in it also play an important role in seeing whether the synthesized material can be used for different purposes or not. Fig. 3(a) (H-1, H-2, H-3) highlights quantitative analysis on the sample, including the chemical content as well as the details of the elements in the sample. This verification is crucial for establishing the observed physical characteristics with the chemical makeup of the material and to fine-tune the synthesis process to obtain the state of the art. Altogether, the structural, elemental, and purity data present a holistic picture of the material to the researcher. These kinds of synergistic approaches are important to design ultra-coherent nano-composite materials that can be used within high-technology industries, giving a solid basis for further investigation of the subject (Knight et al., 2015; Pasierb et al., 2021; Lee et al., 2013; Mai et al., 2015; Sailaja et al., 2016; Stockford et al., 2015).

4.3 Analysis of current-voltage

The investigations investigations of the said properties have revealed that the conductivity of the synthesized samples enhances with the enhanced concentration of Ce and Mn dopants in the sample, and the resistivity of the synthesized sample decreases with time. As it was expected, the investigated materials are semiconductors as the resistivity of the films has decreased and the drift motion of the electric ions carriers increased. From the graph, it can be inferred that as dopant content rises, conductivity rises as well, which implies the possibility of use in circuitry. Verwey hopping mechanism describes electrical conduction in ferrites as the result of electron displacement between the ions of different valency being located in the lattice sites at random. Fig. 3(b) indicates that the high resistance range is 105-108 Ω; nevertheless, this may be because of the high conductivity of material (Perry et al. 2018). This high conductivity is due to the hopping of electrons between ferric (Ce3+) and ferrous (Mn2+) ions in B-sites. The observed behavior supports the hypothesis that the synthesized materials may possess suitable characteristics for electrical applications (Yarahmadi et al., 2021); (Lv et al., 2015). The electrical behavior of the compound SrxCe1-xMny (where x is 0. 03, 0. 05, 0. 07 and y is 1) can be well elucidated based resistivity (ρ) measurements with the specimen at room temperature as depicted in Fig. 4(a). The results indicate that all samples possess exactly opposite behaviors for conductivity as well as resistivity (Hung et al. 2019).

(a) Plot (αhv)2 versus energy band gap (eV) of Sample (SrxCe1-xMny, where x= 0.03, y=1), (b) (SrxCe1-xMny, where x= 0.05, y=1), (c) (SrxCe1-xMny, where x= 0.07, y=1).
Fig. 4.
(a) Plot (αhv)2 versus energy band gap (eV) of Sample (SrxCe1-xMny, where x= 0.03, y=1), (b) (SrxCe1-xMny, where x= 0.05, y=1), (c) (SrxCe1-xMny, where x= 0.07, y=1).

4.4 UV-Visible spectroscopy

Energy gap (eV), transmittance, absorbance, and reflectance are the primary study areas of current experimental analysis. The optical properties of synthesized nanomaterials can be explained via UV-visible spectra from the range (200 nm to 1100 nm Absorption wavelength) as well as the Tauc plot (optical bandgap of synthesized nanomaterials plotting absorption coefficient (α) against photon energy (hν). Although optical research may take many different forms of analysis, UV-visible spectroscopy is one of the important, intriguing techniques for studying functional properties. The energy band gap and particle size of the material have an inverse relationship with the absorption wavelength. The given formula is utilized to calculate the absorption coefficient α and energy band gap (Ali et al., 2022).

α hv = A ( h ν  E g ) n E g = hc / λ

Where n is an index that can take on the values of 1/2, 3/2, 2, or 3, based on the kind of transition of electrons causing the reflection, and Eg is the energy gap, constant A, which varies for various transitions. Direct transition assessed by the all-energy gap. Figs. 4(a-c) shows the UV spectra, which were measured between 1.62-1.98 eV. At x=0.70 and y=1, the absorption spectra with Ce and Mn substitution have a minimum band gap of 1.62 eV, which is appropriate for better optical qualities. The synthesized materials’ band gap decreased steadily as their composition grew, which is characteristic of semiconductor materials. Particle size also impacts optical qualities, with smaller particles often having superior optical properties (Yarahmadi et al., 2021). The absorbance measurements indicate sunlight absorption, making these materials appropriate for solar cell as fuel cell applications. Specifically, the composition SrxCe1-xMny (x=0.07, y=1) is ideal for Solar cell as fuel cell production and energy storage devices. It is clear from the plots that if the molar concentration of Sr increases from 0.03 to 0.07, y=1, the energy bandgap value decreases from 1.98 to 1.62, as shown in Figs. 4(a and c) (Bhushan et al. 2010).

From Table 2. As long as the dopant concentration of Srx increases from 0.03, 0.05, and 0.07 molarity ratios, SrxCe1-x Mny (x=0.03, 0.05, 0.07, y=1), the energy bandgap decreases from 1.98 to 1.62 finally. This demonstrates that if the minimal ratio of Sr increases, materials become increasingly conductive (Yarahmadi et al., 2021).

Table 2. Band gap data of SrxCe1-xMny;(x=0.03,0.05,0.07, y=1)
Sample ID Band gap (eV)
Srx Ce1-x Mny; (x=0.03, y=1) 1.98
Srx Ce1-x Mny; (x=0.05, y=1) 1.88
Srx Ce1-x Mny; (x=0.07, y=1) 1.62

4.5 Dielectric analysis

The dielectric characteristics of Ce3+ substituted SrxCe1-xMny; (x=0.03, 0.05, 0.07, y=1) at room temperature were determined with the frequency ranges of 4 Hz to 8 MHz (Mazumder et al., 2019). From the results presented in Figs. 5 and 6, it may be concluded that for all samples, the dielectric constant decreases with increasing frequency and starts to level off at frequencies above 10 kHz. At low frequencies, the behavior of the dielectric is dependent on the space charge and dipolar polarization. In the Ce3+ substituted Sr samples, oxygen and Ce vacancies cause the formation of space charges. In lower frequencies, dipoles can orient themselves along the direction of the electric field, whereas in higher frequencies, phase retardation takes place, and dipoles are unable to follow the oscillations of the electric field. In this case, the obtained results are in good agreement with the descriptions of the dipole relaxation in the literature (Li et al., 2016). On the same note, the positive dielectric constant value recorded in the samples may be due to free oscillating electrons that create a plasmonic state that resembles that of metal plasma oscillations when an electric field is applied. Furthermore, Figs. 5 and 6 demonstrate that the dielectric constant is maximum for the sample of SrxCe1-xMny (x=0.07, y=1) at low frequencies. The tangent loss (tan δ) first drops as frequency increases up to 10 kHz, after which it rises. The minimum loss is recorded for SrxCe1-xMny (where x=0.05 and y=1). This behavior supports the concept proposed by Koop and the two-layer model of Maxwell-Wagner for magnetization, electrical conduction, and electron and grain responses at different frequencies (Bellino et al., 2016; Tahir et al., 2023; Munir et al., 2023). Similar studies for biomedical applications have been conducted and successfully applied (Ahmed et al., 2025; Badshah et al., 2025).

(a) Dielectric constant of prepared sample (Srx Ce1-x Mny where x = 0.03, y=1, (b) Dielectric constant of prepared sample (Srx Ce1-x Mny where x = 0.05.
Fig. 5.
(a) Dielectric constant of prepared sample (Srx Ce1-x Mny where x = 0.03, y=1, (b) Dielectric constant of prepared sample (Srx Ce1-x Mny where x = 0.05.
Dielectric loss and tangent loss of prepared samples (SrxCe1-xMny, where x=0.03, 0.05, 0.07, y=1).
Fig. 6.
Dielectric loss and tangent loss of prepared samples (SrxCe1-xMny, where x=0.03, 0.05, 0.07, y=1).

4.6 Electrochemical behavior

At normal temperature, the electrochemical study was conducted with a scan rate of 1 mVsec-1. The computed values have been compiled in Table 2, and the cyclic voltammetry (CV) has been visually depicted in Fig. 7. As the applied potential difference (V) increased, variations in the form of the CV curves were seen. Using a slurry approach, the as-prepared H-1, H-2, and H-3 samples were put onto nickel foam electrodes for use in energy storage applications. A potentiostat with three electrodes, the functioning electrode, the counter electrode, and the reference electrode (Ag/AgCl) was used to investigate electrochemical studies. To calculate specific capacitance (Csp), use the formula Csp=I×Δt/m×3600, where I(A) is the current discharge, Δt is the discharge period, and m(g) is the mass of the active specimen. The specific capacitance of all synthetic samples was determined to be (382.39, 436.97, and 500.312) F/g. The Csp value was greatest for sample H-3 and lowest for sample H-1. A similar tendency was previously noted by several researchers. They have all been displayed as computed values in Table 3. The substantial dependency of this outcome may be explained by an increase in electrochemical activity. The electrochemical function of the electrode can be developed by storing more electrolytic ions during the intercalation process, which is made feasible by this kind of porous nanostructure. Therefore, the unique pseudocapacitive property is ascribed to the intrinsic electrochemical activity of transition metal sites, which might be greatly enhanced by generating anion vacancies, thus offering a possible material response to future energy storage requirements (Elanthamilan & Wang, 2023; Dastgir et al., 2025; Li et al., 2024).

CV curves of prepared samples SrxCe1-xMny, (where x=0.03, 0.05, 0.07 and y=1).
Fig. 7.
CV curves of prepared samples SrxCe1-xMny, (where x=0.03, 0.05, 0.07 and y=1).
Table 3. CV data of prepared samples (SrxCe1-xMny, where x= 0.03, 0.05, and 0.07, y=1)
Sample ID Scan rate mV sec-1 Potential window PW(V) Area of loop (Ampere× Volt) Mass (g) Specific capacitance Csp (F/g)
H-1 1 0.8 0.03671 0.06 382.39
H-2 1 0.8 0.04195 0.06 436.97
H-3 1 0.8 0.04803 0.06 500.31

5. Conclusions

The present work focuses on the production and analysis of Sr-doped Ce/Mn nanocomposites with a chemical composition of SrxCe1-xMny (x=0.03, 0.56, 0.07, and y=1). The composites were synthesized by a sol-gel process and subsequently subjected to heat treatment. Structural characterization was performed using XRD analysis, which revealed that the crystallite size for all samples was 54.08 nm, which is fairly satisfactory given that the size is less than 100 nm. Studies suggest that particle size has a crucial role in a variety of high-tech applications. Morphological study revealed distinct grain boundaries on the samples and uniform size throughout the surface. These grain boundaries inhibit electrical conductivity. In addition, the prepared material’s energy band gap was studied using UV-Visible spectroscopy. The synthesized material’s band gap fell gradually as its composition increased, as is typical of semiconductor materials. An electrochemical workstation and I-V measurements demonstrated that the conductivity of the synthesized samples increases with increasing concentrations of Ce and Mn dopants in the sample. Additionally, the specific capacitance of the electrolyte material in all prepared samples was found to be (382.39, 436.97, and 500.31) F/g. The Csp value was greatest in the sample having a greater concentration of dopant. Overall, the current characterization technique offers useful insights into the structural, optical, and electrical properties of the synthesized nanocomposites.

Acknowledgment

The authors extend their sincere appreciation to the Ongoing Research Funding program (ORF-2025-397), King Saud University, Riyadh, Saudi Arabia, for the financial support.

CRediT authorship contribution statement

Muhammad Fakhar-e-Alam: Writing review editing and format, Analysis. Muhammad Hashim: Writing-Format Analysis. Tanveer Hussain Bukhari: Writing review editing and Formal Analysis. Muhammad Sajid: Software Investigation. Ghulam Dastgir: Technical Discussion. Muhammad Atif: Formal Analysis, Proof reading. Muhammad Jehanzaib Aslam: Software investigation.

Declaration of competing interest

The authors declare that they have no competing financial interests or personal relationships that could have influenced the work presented in this paper.

Declaration of Generative AI and AI-assisted technologies in the writing process

The authors confirm that there was no use of Artificial Intelligence (AI)-Assisted Technology for assisting in the writing or editing of the manuscript and no images were manipulated using AI.

References

  1. , , , , , . Structural characterization and dielectric properties of ceria–titania nanocomposites in low ceria region. Mater Res Express. 2017;4:125016. https://doi.org/10.1088/2053-1591/aa9c51
    [Google Scholar]
  2. , , , , , , , , , . Synthesis and characterization of Zn(1-x-y)MnxCoyO NPs for liver cancer treatment. CPD 2025:31. https://doi.org/10.2174/0113816128330548250206101727
    [Google Scholar]
  3. , , , , , , . A brief description of high temperature solid oxide fuel cell’s operation, materials, design, fabrication technologies and performance. Appl Sci. 2016;2076-3417:6.
    [Google Scholar]
  4. , , , , , , , , , , . Tailoring the optical and magnetic properties of La-BaM hexaferrites by Ni substitution. Chinese Phys. B. 2022;31:027502. https://doi.org/10.1088/1674-1056/ac1412
    [Google Scholar]
  5. , , , , , , . Structural properties of Sm-doped ceria electrolytes at the fuel cell operating temperatures. Solid State Ionics. 2018;315:85-91. https://doi.org/10.1016/j.ssi.2017.12.009
    [Google Scholar]
  6. , , . Processing and conduction behavior of nanocrystalline Gd-doped and rare earth co-doped ceria electrolytes. Electrochimica Acta. 2016;209:541-550. https://doi.org/10.1016/j.electacta.2016.05.118
    [Google Scholar]
  7. , , , , . Coherent control of surface plasmon polaritons excitation via tunneling-induced transparency in quantum dots. Opt & Laser Tech. 2025;182:112078. https://doi.org/10.1016/j.optlastec.2024.112078
    [Google Scholar]
  8. , , , . Solid oxide fuel cell systems in hydrogen-based energy storage applications: Performance assessment in case of anode recirculation. J Energy Storage. 2022;54:105257. https://doi.org/10.1016/j.est.2022.105257
    [Google Scholar]
  9. , , , , , , . Barium non‐stoichiometry role on the properties of Ba1+ xCe0. 65Zr0. 20Y0. 15O3–δ proton conductors for IT‐SOFCs. Fuel Cells. 2008;8:360-368.
    [Google Scholar]
  10. , , . Enhanced ionic conductivity in nanostructured, heavily doped ceria ceramics. Adv Funct Materials. 2006;16:107-113. https://doi.org/10.1002/adfm.200500186
    [Google Scholar]
  11. , , , , . Sr induced modification of structural, optical and magnetic properties in Bi1−xSrxFeO3 (x= 0, 0.01, 0.03, 0.05 and 0.07) multiferroic nanoparticles. Solid State Sci. 2010;12:1063-1069.
    [Google Scholar]
  12. , , , , , , . Phase Transformations in the CeO2–Sm2O3 System: a multiscale powder diffraction investigation. Inorg Chem. 2018;57:879-891.
    [Google Scholar]
  13. , . Ionic conductivity in multiply substituted ceria-based electrolytes. Solid State Ionics. 2018;316:9-19. https://doi.org/10.1016/j.ssi.2017.12.013
    [Google Scholar]
  14. , , , , . Structural and electrochemical analysis of enhanced ionic conductivity and oxygen permeability in La0.60Bi0.15 Cr0.05Fe0.20O3-δ nanocomposites for high-performance ceramic fuel cells. Ceram Int. 2025;51:37179-37188. https://doi.org/10.1016/j.ceramint.2025.05.426
    [Google Scholar]
  15. , . Properties and applications of perovskite proton conductors. Mat Res. 2010;13:385-394. https://doi.org/10.1590/s1516-14392010000300018
    [Google Scholar]
  16. , , , . High performance metal-supported solid oxide fuel cells with infiltrated electrodes. J Power Sources. 2019;410:91-98. https://doi.org/10.1016/j.jpowsour.2018.11.004
    [Google Scholar]
  17. , . Construction of 3D flower-like SnS particles anchored Ni-metal-organic framework composite: A novel positive electrode material for high-performance asymmetric supercapacitors. J Energy Storage. 2023;71:108144. https://doi.org/10.1016/j.est.2023.108144
    [Google Scholar]
  18. , , , , . Solid oxide fuel cells: Materials properties and performance. CRC press; .
  19. , . Electrochemical and morphological characterizations of La2-xNiO4+d (x = 0.01, 0.02, 0.03 and 0.05) as new cathodes materials for IT-SOFC. Mater Res Bull. 2016;83:268-274. https://doi.org/10.1016/j.materresbull.2016.06.009
    [Google Scholar]
  20. , , , , , , , , , . Effects of manganese oxides on the activity and stability of Ni-Ce0.8Sm0.2O1.9 anode for solid oxide fuel cells with methanol as the fuel. Catal Today. 2019;330:222-227. https://doi.org/10.1016/j.cattod.2018.01.014
    [Google Scholar]
  21. , , , , , , . A high performance composite cathode with enhanced CO2 resistance for low and intermediate-temperature solid oxide fuel cells. J Power Sources. 2018;405:124-131. https://doi.org/10.1016/j.jpowsour.2018.10.025
    [Google Scholar]
  22. , , , , , . Fabrication of nanoscale yttria stabilized zirconia for solid oxide fuel cell. Int J Hydrogen Energy. 2017;42:29949-29959. https://doi.org/10.1016/j.ijhydene.2017.08.143
    [Google Scholar]
  23. , , , , . High performance of protonic solid oxide fuel cell with BaCo0.7Fe0.22Sc0.08O3−δ electrode. Int J Hydrogen Energy. 2017;42:25021-25025. https://doi.org/10.1016/j.ijhydene.2017.08.107
    [Google Scholar]
  24. , , , , . NiMo-ceria-zirconia-based anode for solid oxide fuel cells operating on gasoline surrogate. Appl Catal B: Environ. 2019;242:31-39. https://doi.org/10.1016/j.apcatb.2018.09.095
    [Google Scholar]
  25. , , , , , , , , . Facile synthesis and electrochemical properties of perovskite‐type CeMnO3 nanofibers. ChemistrySelect. 2019;4:11903-11912. https://doi.org/10.1002/slct.201903426
    [Google Scholar]
  26. , , , , . Phase stability and conductivity of Ba1−ySryCe1−xYxO3 solid oxide fuel cell electrolyte. J Power Sources. 2019;193:155-159.
    [Google Scholar]
  27. , , , . Microstructures and grain boundary conductivity of BaZr1−xYxO3 (x = 0.05, 0.10, 0.15) ceramics. Solid State Ionics. 2007;178:691-695. https://doi.org/10.1016/j.ssi.2007.02.019
    [Google Scholar]
  28. , , , . Metal alloy hybrid nanoparticles with enhanced catalytic activities in fuel cell applications. J Solid State Chem. 2019;270:295-303. https://doi.org/10.1016/j.jssc.2018.11.014
    [Google Scholar]
  29. , . The crystal structures of some doped and undoped alkaline earth cerate perovskites. Mater Res Bull. 1995;30:347-356. https://doi.org/10.1016/0025-5408(95)00009-7
    [Google Scholar]
  30. , , , , , . Strontium doping effect on phase homogeneity and conductivity of Ba1−xSrxCe0.6Zr0.2Y0.2O3 proton-conducting oxides. Int J Hydrogen Energy. 2013;38:11097-11103.
    [Google Scholar]
  31. , , , , , . Efficient low-temperature catalytic combustion of trichloroethylene over flower-like mesoporous Mn-doped CeO2 microspheres. Appl Catal B: Environ. 2011;102:475-483. https://doi.org/10.1016/j.apcatb.2010.12.029
    [Google Scholar]
  32. , , , , , . Influence of lightly Sm-substitution on crystal structure, magnetic and dielectric properties of BiFeO3 ceramics. J Alloys Compd. 2016;682:672-678. https://doi.org/10.1016/j.jallcom.2016.05.023
    [Google Scholar]
  33. , , , , , , , , . Effect of Sn addition on improving the stability of Ni-Ce0.8Sm0.2O1.9 anode material for solid oxide fuel cells fed with dry CH4. Catal Today. 2019;330:209-216. https://doi.org/10.1016/j.cattod.2018.04.030
    [Google Scholar]
  34. , , , , , , . Reduced graphene oxide-wrapped ZnS–SnS2 heterojunction bimetallic hollow cubic boxes as high-magnification and long lifespan supercapacitor anode materials. Nanoscale; .
  35. , , , , , , . The next generation solid acid fuel cell electrodes: Stable, high performance with minimized catalyst loading. J Mater Chem A. 2017;5:15021-15025. https://doi.org/10.1039/c7ta03690f
    [Google Scholar]
  36. , , , , , . Chemical stability of doped BaCeO3-BaZrO3 solid solutions in different atmospheres. J Rare Earths. 2008;26:505-510. https://doi.org/10.1016/s1002-0721(08)60127-1
    [Google Scholar]
  37. , , , , . Sintering, chemical stability and electrical conductivity of the perovskite proton conductors BaCe0.45Zr0.45M0.1O3 (M= In, Y, Gd, Sm) J Alloys Compd. 2015;467:376-382.
    [Google Scholar]
  38. , , , , , , , . Shape-selective synthesis and oxygen storage behavior of ceria nanopolyhedra, nanorods, and nanocubes. J Phys Chem B. 2005;109:24380-24385. https://doi.org/10.1021/jp055584b
    [Google Scholar]
  39. , . Effect of Pb-doping on dielectric properties of BiFeO3 ceramics. J Alloys Compd. 2009;475:577-580. https://doi.org/10.1016/j.jallcom.2008.07.082
    [Google Scholar]
  40. , , , , , , . Study of electrode performance for nanosized La0.4Sr0.6Co0.8Fe0.2O3 SOFC Cathode. Electrochimica Acta Transactions. 2015;66:169-182.
    [Google Scholar]
  41. , , , , , . Synthesis and ionic conductivity of calcium-doped ceria relevant to solid oxide fuel cell applications. Mater Adv. 2022;3:8780-8791. https://doi.org/10.1039/d2ma00868h
    [Google Scholar]
  42. , , , , , , . Structural, optical and thermoelectric properties of (Al and Zn) doped Co9S8-NPs synthesized via co-precipitation method. J King Saud Univ Sci. 2023;35:102836. https://doi.org/10.1016/j.jksus.2023.102836
    [Google Scholar]
  43. , , , . Structural properties of Li2CO3-BaCO3 system derived from IR and Raman spectroscopy. J Mol Struct. 2021;596:151-156. https://doi.org/10.1016/S0022-2860(01)00703-7
    [Google Scholar]
  44. , , . Local electrical and dielectric properties of nanocrystalline yttria-stabilized zirconia. J Mater Sci. 2008;43:4684-4692. https://doi.org/10.1007/s10853-008-2553-x
    [Google Scholar]
  45. , , . Low temperature densification by lithium co-doping and its effect on ionic conductivity of samarium doped ceria electrolyte. Ceram Int. 2019;45:5819-5828. https://doi.org/10.1016/j.ceramint.2018.11.251
    [Google Scholar]
  46. , , , , , , . Copper local structure in spinel ferrites determined by X-ray absorption and mössbauer spectroscopy and their catalytic performance. Mater Res Bull. 2019;109:117-123. https://doi.org/10.1016/j.materresbull.2018.09.017
    [Google Scholar]
  47. , , , . Synthesis and characterization of Ba0.96Sr0.04Ce0.7Zr0.3O3 solid oxide fuel cell electrolyte by citrate EDTA complexing sol-gel process. Astron. 2016;71:1-10.
    [Google Scholar]
  48. , , , , , , , , , , , , , , . H2FC SUPERGEN: An overview of the hydrogen and fuel cell research across the UK. Int J Hydrogen Energy. 2015;40:5534-5543. https://doi.org/10.1016/j.ijhydene.2015.01.180
    [Google Scholar]
  49. , , , , . Investigation of optical, electrical and magnetic properties of hematite a-Fe2O3 nanoparticles via sol-gel and co-precipitation method. J King Saud Univ Sci. 2023;35:102695.
    [Google Scholar]
  50. , , , , , , , , . A facile method for preparing the CeMnO3 catalyst with high activity and stability of toluene oxidation: The critical role of small crystal size and Mn3+-Ov-Ce4+ sites. J Hazard Mater. 2024;470:134114. https://doi.org/10.1016/j.jhazmat.2024.134114
    [Google Scholar]
  51. , , , . Synthesis and properties of SmBaCo2−xNixO5 perovskite oxide for IT-SOFC cathodes. Ceramics International. 2016;42:1272-1280.
    [Google Scholar]
  52. , , , , , , . Proton shuttles in CeO2/CeO2− δ core–shell structure. ACS Energy Lett. 2019;4:2601-2607.
    [Google Scholar]
  53. , , , , , . Synthesis and characterization of Sr-doped ZnO nanoparticles for photocatalytic applications. J Alloys Compd. 2021;853:157000. https://doi.org/10.1016/j.jallcom.2020.157000
    [Google Scholar]
  54. , , , , . Facile synthesis of perovskite CeMnO3 nanofibers as an anode material for high performance lithium-ion batteries. RSC Adv. 2019;9:38271-38279. https://doi.org/10.1039/c9ra07660c
    [Google Scholar]
  55. , , , , , , , , , , . Strongly correlated perovskite fuel cells. Nature. 2016;534:231-234. https://doi.org/10.1038/nature17653
    [Google Scholar]
  56. , , . Semiconductor‐ionic materials could play an important role in advanced fuel‐to‐electricity conversion. Int J Energy Research. 2018;42:3413-3415. https://doi.org/10.1002/er.4105
    [Google Scholar]
Show Sections