Next Article in Journal
Kinetic Analysis of Construction, Renovation, and Demolition (CRD) Wood Pyrolysis Using Model-Fitting and Model-Free Methods via Thermogravimetric Analysis
Previous Article in Journal
A System Thinking Approach to Circular-Based Strategies for Deep Energy Renovation: A Systematic Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Strontium-Doped Tin Oxide Nanofibers for Enhanced Visible Light Photocatalysis

Advanced Manufacturing Alliance, Energy and Resources Institute, Faculty of Science and Technology, Charles Darwin University, Darwin, NT 0909, Australia
*
Author to whom correspondence should be addressed.
Energies 2025, 18(10), 2495; https://doi.org/10.3390/en18102495
Submission received: 30 March 2025 / Revised: 30 April 2025 / Accepted: 6 May 2025 / Published: 12 May 2025
(This article belongs to the Section D1: Advanced Energy Materials)

Abstract

:
This study investigates the photocatalytic degradation of methylene blue (MB) using strontium-doped SnO2 nanofibers synthesized via electrospinning. The 1% Sr-doped SnO2 nanofibers exhibited remarkable photocatalytic activity, achieving 84.74% MB degradation under visible light irradiation, substantially outperforming both undoped SnO2 nanofibers (61%) and the same catalyst under UV light (69%) under identical experimental conditions. Comprehensive electrochemical investigations revealed that Sr doping fundamentally transformed interfacial charge transfer kinetics, with 1% Sr-doped nanofibers exhibiting a remarkable three-fold decrease in charge transfer resistance (404 Ω compared to 1350 Ω for undoped samples), a dramatic enhancement in charge carrier density (5.17 × 1022 versus 9.24 × 1019 for undoped samples), and an approximately eight-fold increase in diffusion coefficient (8.78 × 10−10 versus 1.13 × 10−10 cm2s−1). These electrochemical improvements were corroborated by comprehensive structural characterization, which demonstrated that strategic Sr incorporation induced beneficial oxygen vacancies, reduced crystallite size, increased microstrain, and enhanced dislocation density, collectively contributing to superior surface reactivity and accelerated photocatalytic mechanisms. This work establishes a quantitative correlation between electrochemical characteristics and photocatalytic activity in Sr-doped SnO2 nanofibers, revealing the fundamental mechanisms that transform the SnO2 nanostructure from UV-dependent to efficient visible light-driven catalysts for organic pollutant degradation.

1. Introduction

Water pollution represents one of the most pressing environmental challenges of our time, with industrial effluents containing organic pollutants, which pose significant threats to aquatic ecosystems and human health. Among various treatment approaches, photocatalytic degradation has emerged as a promising advanced oxidation process, offering complete mineralization of pollutants into harmless end products like carbon dioxide and water [1]. This process harnesses the power of semiconductor materials that, when activated by light, generate reactive species capable of degrading organic contaminants without producing secondary waste streams.
Semiconductor-based photocatalysts function through a fundamental mechanism wherein light absorption promotes electrons from the valence band to the conduction band, creating electron-hole pairs. These charge carriers migrate to the catalyst surface where they initiate redox reactions with adsorbed water and oxygen molecules, generating highly reactive hydroxyl (•OH) and superoxide (•O2−) radicals [2]. These radicals subsequently attack and decompose organic pollutants through successive oxidation reactions. The efficiency of this process depends critically on three factors: light absorption capacity, charge carrier separation and transport, and surface catalytic activity.
Among various semiconductor photocatalysts, tin oxide (SnO2) has attracted considerable attention due to its exceptional chemical stability, non-toxicity, earth abundance, and relatively low cost [3,4]. With a wide bandgap of approximately 3.6 eV, high electron mobility (100–200 cm2V−1s−1), and excellent resistance to photocorrosion, SnO2 shares many advantageous properties with the widely studied titanium dioxide (TiO2) [5]. However, despite these favorable characteristics, the practical application of SnO2 in photocatalysis faces two significant challenges. First, its wide bandgap restricts light absorption primarily to the ultraviolet region, which constitutes only about 5% of the solar spectrum, severely limiting its efficiency under natural sunlight [4,6]. Second, the rapid recombination of photogenerated charge carriers diminishes quantum efficiency and overall photocatalytic performance [7].
Various strategies have been explored to enhance the photocatalytic performance of SnO2. These include forming heterojunctions with other semiconductors, such as ZnO [8] and TiO2 [9,10], which promote charge separation through interfacial electron transfer. While heterojunction formation has demonstrated improved photocatalytic activity, these composite systems often involve complex synthesis procedures and may introduce new interfaces that can act as recombination centers. Alternatively, doping with various elements offers a more direct approach to modify the electronic structure of SnO2, enhancing both light absorption and charge carrier dynamics.
Doping introduces impurity levels within the bandgap or creates oxygen vacancies, effectively narrowing the bandgap and extending light absorption into the visible region [11,12]. Additionally, dopants can act as trap sites for either electrons or holes, facilitating charge separation and reducing recombination rates. Previous studies have investigated SnO2 doping with various elements, including magnesium [13], niobium [14], potassium [15], and several transition metals, reporting enhanced photocatalytic activity. Among potential dopants, strontium (Sr) emerges as a particularly promising candidate due to its unique properties [7]. Strontium is a promising dopant for SnO2 due to its high stability and larger ionic radius (Sr2+ = 1.12 Å) than Sn4+ (0.71 Å) except potassium (K+ = 1.33 Å). But K+ has a charge of +1, which may create a charge imbalance in the SnO2 lattice, whereas Sr2+ can more effectively compensate for the charge imbalance by substituting a divalent ion (Sr2+) for a tetravalent Sn4+ in the SnO2 lattice [7]. Because of the larger ionic radius of Sr2+ and the compensating charge imbalance in Sn4+, Sr is a promising candidate when compared to other doping materials for SnO2.
Strontium, with its larger ionic radius compared to tin, introduces significant lattice distortion when incorporated into the SnO2 crystal structure, creating oxygen vacancies to maintain charge neutrality [7]. These oxygen vacancies serve dual functions: They create mid-gap energy levels, effectively reducing the bandgap and enhancing visible light absorption, while also acting as trapping sites that facilitate charge separation. Additionally, the divalent nature of Sr2+ (compared to tetravalent Sn4+) creates a charge imbalance that enhances electronic conductivity, potentially improving charge transport properties. Despite these potential advantages, the application of Sr-doped SnO2 nanostructures in photocatalysis remains relatively unexplored.
Beyond compositional modifications, the morphological engineering of photocatalysts represents another effective strategy to enhance performance. Nanostructured materials, with their high surface-to-volume ratios, provide abundant active sites for photocatalytic reactions while reducing bulk recombination through shortened charge diffusion distances. Among various nanostructures, one-dimensional (1D) nanofibers offer distinct advantages for photocatalytic applications [16]. Unlike nanoparticles, which tend to aggregate in aqueous media, nanofibers maintain their dispersibility while providing interconnected pathways for electron transport. Their high aspect ratio and nanoscale diameter result in significantly enhanced surface area and reduced average crystallite size compared to conventional nanoparticles [16].
The electrospinning technique represents a versatile and scalable approach for fabricating nanofibers with controlled composition and morphology. This method involves applying a high voltage to a polymer solution, creating a charged jet that undergoes elongation and solvent evaporation, resulting in continuous nanofibers [17]. Subsequent calcination removes the polymer matrix, leaving behind ceramic nanofibers with well-defined structural properties. Despite the advantages of nanofibers for photocatalysis, previous research on SnO2-based systems has primarily focused on nanoparticle morphologies or composite nanofibers (e.g., TiO2/SnO2, ZnO/SnO2), with limited exploration of pure SnO2 nanofibers for photocatalytic applications.
The present study addresses these research gaps by investigating Sr-doped SnO2 nanofibers for enhanced visible light photocatalysis. Through a systematic approach, we explore how Sr doping transforms the structural, optical, and electrochemical properties of SnO2 nanofibers and correlate these properties with photocatalytic performance. The study introduces several novel elements: the preservation of pristine nanofibrous morphology through optimized electrospinning, enhanced visible light activity through strategic Sr doping, and comprehensive electrochemical characterization establishing direct links between charge carrier dynamics and photocatalytic efficiency.
Methylene blue (MB), a common industrial dye, serves as a model pollutant to evaluate the photocatalytic performance of our synthesized materials [18]. As a cationic thiazine dye widely used in textiles, paper, and pharmaceuticals, MB represents a class of recalcitrant organic pollutants that persist in industrial effluents [18,19]. Its well-defined optical properties, with a characteristic absorption maximum at 664 nm, facilitate quantitative analysis of degradation efficiency, making it an ideal probe molecule for assessing photocatalytic activity.
Through a comprehensive investigation of Sr-doped SnO2 nanofibers with varying dopant concentrations (0, 1, 3, and 5%), this study demonstrates that 1% Sr-doped SnO2 nanofibers achieve 84.74% MB degradation under visible light illumination within 180 min, significantly outperforming both undoped SnO2 nanofibers (61%) and the same catalyst under UV light conditions (69%). This enhanced performance is attributed to the synergistic effects of modified electronic structure, improved charge transfer properties, and optimized nanofibrous morphology, as revealed through detailed electrochemical analyses. The findings provide valuable insights into the design principles for high-performance visible light photocatalysts, with potential applications extending beyond dye degradation to various environmental remediation processes.

2. Experimental Methods

2.1. Materials and Reagents

Tin (II) chloride dehydrate (SnCl2·2H2O), tin (IV) oxide (SnO2) (purity > 99.99%, trace metal basis, average particle size of 100 nm), polyvinylpyrrolidone (PVP), ethanol, N, N-dimethylformamide (DMF), strontium nitrate (Sr (NO3)2), methylene blue (MB) (0.05 wt.% in H2O) were purchased from Sigma-Aldrich (Merk, Melbourne, Australia). All chemicals were used without further purification.

2.2. Fabrication of Strontium-Doped SnO2 Nanofibers

SnO2 nanofibers were prepared by electrospinning technology. Briefly, 0.56 gm of SnCl2·2H2O was added to 5.5 mL of DMF and 5.5 mL of ethanol. Then, 0.76 gm of PVP was mixed with the solution. For Sr doping, Sr (NO3)2 was added to the as-prepared SnO2 precursor solution at concentrations of 0.1 g (1 wt.%), 0.3 g (3 wt.%), and 0.5 g (5 wt.%) relative to the total solution weight. These concentrations are referred to throughout the manuscript as 1% Sr-doped, 3% Sr-doped, and 5% Sr-doped SnO2, respectively.
After that, this final solution was stirred at 60 °C for 2 h and then stirred overnight at room temperature. The electrospinning solution was placed in a plastic syringe with a conductive needle (18 G) at a feed rate of 0.5 mL/h. The applied voltage was 17 kV and the distance between the needle tip and the collector wrapped in aluminum foil paper was set to 18 cm. The as-spun nanofiber (NF) was calcined in an oven at 600 °C for 2 h and SnO2 nanofibers were finally achieved after the removal of polymer and solvents.

2.3. Material Characterization

The samples were characterized by various characterization tools. The morphology of NFs was examined by scanning electron microscopy on a ThermoFisher Phenom™ XL G2 Desktop SEM. ImageJ software (v1.54g) was used to measure the average diameter of the fiber. The crystal structure and phase identification were investigated by X-ray diffraction (XRD) analysis using Cu Kα radiation with the wavelength of 1.54178 Å, in the range of 2θ = 20 − 90° on a Malvern Panalytical powder XRD machine.

2.4. Electrochemical Measurements

Electrochemical measurements were conducted using a Gamry Potentiostat IFC1010-27215 (Gamry Instruments, Warminster, PA, USA) in a three-electrode configuration with 1 M KOH, an alkaline medium as the electrolyte. Alkaline electrolytes, such as KOH, are well-established in supercapacitor research for their ability to enhance ion adsorption/desorption at the electrode surface, leading to improved charge storage. This is particularly relevant for pseudocapacitive materials like SnO2, where charge storage is not solely based on double-layer formation [20]. The intercalation and deintercalation of alkaline electrolyte ions can significantly influence the pseudocapacitive behavior of SnO2 nanofibers, potentially improving their performance. Among various alkaline electrolytes, KOH is widely favored in pseudo-capacitance applications of SnO2 [21,22] due to its high ionic conductivity of K+ [23], which accelerates electrochemical kinetics with higher stability and conductivity in an electrochemical environment, dissociating K+ and OH completely in the aqueous solution.
An Ag/AgCl electrode served as the reference, glassy carbon as the working electrode, and platinum as the counter electrode. The working electrode had an active area of 0.071 cm2. The binder solution was prepared as 5 wt.% of PVDF in 5 mL of DMF solution [24] and was stirred overnight at room temperature. The active material on the working electrode was prepared by dispersing 0.1 mg of SnO2 nanofibers in 5 µL of PVDF solution. A paste was then prepared, and a thin layer was deposited on the working electrode’s active area. The electrode was subsequently dried in an oven at 50 °C for 15 min.
The electrochemical study includes cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS), a Mott–Schottky (M-S) plot, linear sweep voltammetry (LSV), and chronoamperometry (CA). In our study, we will use CV to (1) determine the accumulated charge storage area and current density to discuss charge transfer kinetics and(2) observe redox potentials to understand the charge transfer process, We will also calculate specific capacitance following Equation (1) [25].
C + i v d v / 2 μ m V
where C is capacitance, i is current, v is the potential of the CV curve, m is the mass of active material (0.01 mg), V is the potential window (0.8 V), μ is the scan rate (50 mV/s), and i v d v is the integral of current (i) with respect to voltage (v). This represents the area under the CV curve.
Electrochemical Impedance Spectroscopy yields Nyquist plots of real and imaginary impedance, enabling the derivation of an equivalent circuit model. Analysis of this model will allow us to determine key electrochemical parameters: (1) charge transfer resistance to evaluate charge transfer kinetics, (2) double layer capacitance indicative of the charge transfer surface area, (3) series resistance representing the ohmic resistance, (4) diffusion-induced Warburg impedance, and (5) recombination resistance to understand ion diffusion at the electrode–electrolyte interface.
A Mott–Schottky plot, derived from capacitance measurements as a function of applied potential, is employed to determine a semiconductor’s doping concentration and flat band potential. The slope of this plot is directly related to doping concentration, while the x-intercept reveals the flat band potential [26]. This technique, which involves measuring the impedance of an electrochemical interface under varying DC voltage and a fixed AC signal, utilizes the Mott–Schottky equation to analyze the space charge region at semiconductor–electrolyte interfaces. The equation is
C−2 + (2/εε0)(1/NA + 1/ND)(VVfbkT/q)
where C is the capacitance of the space charge region; ε is the dielectric permittivity of the semiconductor; ε0 is the permittivity of free space; NA is the acceptor doping concentration; ND is the donor doping concentration; V is the applied potential; Vfb is the flat band potential; k is the Boltzmann constant; T is the temperature; and q is the elementary charge.
From chronoamperometry, we can calculate the diffusion coefficient by fitting the current transient curve with the Cotrell equation. From the diffusion coefficient, we can investigate the movement speed (faster or slower) of charge carriers to identify more efficient charge transport materials. The Cottrell plot is calculated using Equation (3) [27].
i + n F A c D ( π t ) 1 2
Here, i represents the current, n is the total number of electrons (4) transferred in the reaction, F is Faraday’s constant (96,485 C/mol), c is the initial concentration (6.64 × 10−5) of electroactive material, D is the diffusion coefficient, A is the active area of the electrode (0.071 cm2), and t is time.
Linear Sweep Voltammetry is a potentiodynamic technique where the working electrode’s potential is linearly swept over time at a constant scan rate, while the resulting current is measured. This method is employed to technically investigate electrochemical behavior by determining redox potential, studying reaction mechanisms and kinetics, enabling quantitative analysis, assessing material properties like stability and electrocatalytic activity, and facilitating initial screening and characterization of electrochemical systems through the analysis of the resulting current-potential voltammograms. The Tafel plot is derived from LSV, where the potential values are referenced to the reversible hydrogen electrode (RHE) through an RHE calibration. Equation (4) [28] is used to convert potential from Ag/AgCl to RHE.
ERHE + EAg/AgCl + 0.059pH + E0Ag/AgCl
(E0Ag/AgCl = +0.197 V)
The overpotential (η) was calculated [29] at different current densities using Equation (5) as
η = ERHE − 1.23

2.5. Photocatalytic Experiment

A 10 mg/L MB solution was prepared using deionized water. For photocatalytic degradation, 6 mg of SnO2 nanofibers (NFs) were dispersed in 30 mL of the MB solution (0.2 g/L) under continuous stirring in a 100 mL glass beaker. After the addition of varying percentages of Sr-doped NFs to the MB solution, the beakers were wrapped in aluminum foil and stirred for 60 min in the dark medium to allow for adsorption–desorption equilibrium. Subsequently, the solutions were exposed to UV light (Analytik Jena, Jena, Germany 254/302/365 nm, 8 Watt) irradiation in a photoreactor with a dimension of 38.5 × 19 × 19 cm (W × H × L). Two different light sources, (1) a UV lamp with λ m a x   + 365 nm and (2) a solar simulator lamp with 100 mW/cm2 (1 sun) were used in the experiment. The height between the lamp and the glass beaker was maintained at 10 cm. At predetermined time intervals for 180 min, aliquots of the solution were withdrawn using a pipette and analyzed using UV-Vis spectrophotometry (GENESYS, ThermoFisher, Waltham, MA, USA) to monitor the degradation of the MB absorption peak. The samples were then returned to the beaker for continued irradiation. Finally, degradation efficiency was recorded using the initial concentration and final concentration at different time intervals from the corresponding absorption peaks.

3. Results and Analysis

3.1. Morphological Evaluation of SnO2 NFs

The SEM micrographs in Figure 1 reveal the morphological evolution of SnO2 nanofibers as a function of Sr doping concentration. Figure 1a reveals an agglomerated particulate structure characteristic of the SnO2 nanoparticles (NP), which showcases distinctive morphology compared to the nanofiber (NF). The undoped SnO2 nanofibers (Figure 1b) exhibit a dense network of interconnected fibers with relatively uniform morphology. Upon introducing 1% Sr (Figure 1c), the fibers maintain their continuous structure while showing a slight increase in diameter, which is quantitatively confirmed in Table 1. The 1% Sr-doped sample presents well-defined, smooth fibers with consistent morphology throughout the network. As the Sr content increases to 3% (Figure 1d), the fiber network remains dense but with visibly thinner fibers compared to the 1% sample. A further increase in Sr doping to 5% (Figure 1e) results in a more pronounced reduction in fiber diameter along with the emergence of structural voids and discontinuities within the network.
Table 1 quantifies the diameter variations observed in the SEM images. The undoped SnO2 nanofibers display an average diameter of 259.04 ± 25.57 nm. With the incorporation of 1% Sr, the mean diameter increases to 288.69 ± 17.24 nm, representing an 11.4% expansion. This diameter increase can be attributed to the substitution of Sn4+ (0.71 Å) by the larger Sr2+ ions (1.12 Å), which induce lattice distortion. Notably, the standard deviation in diameter measurements decreases from ±25.57 nm to ±17.24 nm with 1% Sr doping, suggesting enhanced uniformity in fiber formation. As Sr doping increases to 3% and 5%, the average fiber diameter progressively decreases to 209.49 ± 15.29 nm and 148.72 ± 5.98 nm, respectively. This represents a 19.1% reduction at 3% and a substantial 42.6% reduction at 5% Sr doping compared to the undoped sample. The standard deviation at 5% Sr doping (±5.98 nm) is significantly lower than for other samples, indicating highly uniform but much thinner fibers. This non-monotonic trend in fiber diameter with increasing Sr content suggests complex structural reconfiguration mechanisms at higher doping concentrations. Smaller fiber diameters lead to a larger specific surface area [30]. Our SnO2 nanofibers showed a nonlinear relationship with Sr doping in fiber diameter. At 1% doping, the fiber diameter slightly increased to 288 nm from the pristine 259 nm, suggesting a plausible minor reduction in specific surface area. However, with 3% and 5% Sr doping, the fiber diameter decreased, indicating a corresponding increase in specific surface area.
The observed morphological changes with Sr doping concentration are likely driven by multiple competing factors. At low doping levels (1 wt.%), the incorporation of Sr2+ ions primarily causes lattice expansion [31], resulting in larger fiber diameters. However, at higher doping concentrations (3 wt.% and 5 wt.%), additional mechanisms such as increased lattice strain, enhanced densification during calcination [32], and the possible formation of secondary phases becomes dominant, leading to fiber shrinkage. The structural discontinuities observed in the 5 wt.% Sr-doped sample suggest that excessive Sr incorporation may compromise the structural integrity of the nanofiber network, potentially affecting the material’s functional properties in catalytic applications.
The EDS (energy-dispersive X-ray spectroscopy) for 1% Sr-doped SnO2 NF derived from SEM showcased successful incorporation of Sr2+ in Figure 2.

3.2. Structural Properties

The XRD patterns presented in Figure 3 provide critical insights into the crystalline structure of the synthesized Sr-doped SnO2 nanofibers and reveal how Sr incorporation influences the structural properties of these materials. The diffraction pattern of the manufactured SnO2 nanoparticles (Figure 3a) serves as a reference standard, displaying well-defined, sharp diffraction peaks at 2θ values of approximately 26.6° (110), 33.9° (101), 37.9° (200), 51.8° (211), 54.8° (220), 57.8° (310), 61.9° (112), 64.7° (301), 71.3° (202), and 78.7° (321). These peaks correspond to the tetragonal rutile structure of SnO2 (JCPDS card file no. 41-1445) [33,34], confirming the phase purity of the reference material. The pristine SnO2 nanofibers (Figure 3b) exhibit diffraction peaks at identical 2θ positions to the manufactured SnO2, verifying the successful synthesis of pure SnO2 nanofibers. However, the diffraction peaks of the pristine nanofibers appear broader and less intense compared to the manufactured SnO2, indicating smaller crystallite sizes in the electrospun nanofibers. This observation is consistent with the nanofibrous morphology observed in the SEM images (Figure 1a), where the electrospinning process and subsequent calcination result in nanostructured materials with reduced crystallite dimensions.
Upon Sr doping, significant changes in the XRD patterns become evident. The 1% Sr-doped sample (Figure 3c) maintains the primary diffraction peaks of rutile SnO2 but also exhibits additional peaks at approximately 21.58° (004), 23.53° (202), 30.39° (204), and 46.35° (332), suggesting the formation of a secondary phase [35], displayed in Figure 3f. These new peaks likely correspond to strontium–tin oxide compounds Sr3Sn5, with a tetragonal crystal structure matched from Highscore software from Malvern Panalytical (v 5.2), indicating successful incorporation of Sr into the SnO2 matrix through the formation of Sr-Sn compositions. The diffraction peaks in the 1% Sr-doped sample appear broader than those of pristine SnO2, indicating further reduction in crystallite size [36] with Sr incorporation.
As Sr doping increases to 3% (Figure 3d) and 5% (Figure 3e), there is a progressive reduction in peak intensity and an increase in peak broadening. This trend suggests a continuous decrease in crystallite size [36] with increasing Sr concentration, which complements the observed reduction in nanofiber diameter from 288.69 nm at 1% Sr to 148.72 nm at 5% Sr, as quantified in Table 1. The 5% Sr-doped sample (Figure 3e) shows significantly suppressed diffraction peaks for most crystallographic planes, with only a few dominant peaks remaining visible. This substantial peak suppression indicates severe lattice distortion [37] and possibly increased amorphization at high Sr doping levels.
The structural changes observed in the XRD patterns correlate well with the morphological evolution seen in the SEM images (Figure 1). The initial increase in fiber diameter with 1% Sr doping, followed by diameter reduction at higher doping levels, parallels the crystallographic changes observed in XRD. The introduction of Sr2+ ions, which are significantly larger than Sn4+ ions, creates lattice distortions and oxygen vacancies that fundamentally alter both the crystalline structure and fiber morphology. These structural modifications are expected to significantly influence the electronic properties and photocatalytic performance of the Sr-doped SnO2 nanofibers, particularly through the creation of oxygen vacancies that can enhance visible light absorption and charge carrier separation efficiency. The quantitative crystallographic parameters presented in Table 2 provide deeper insights into the structural modifications induced by Sr doping and corroborate the qualitative observations from the XRD patterns in Figure 3. Table 2 presents the crystal properties of undoped and doped SnO2 NFs. Crystallite size was derived from the Debye–Scherrer equation from Equation (6) [38].
Crystallite   size ,   D   =   k λ β c o s θ
where k is the shape factor, typically around 0.9; λ is X-ray wavelength, the wavelength of Cu is 0.154 nm; β is the peak full width of half maximum intensity (FWHM) in radians, and θ is Bragg’s diffraction angle.
Dislocation density, δ , is calculated using Equation (7) [39].
Dislocation   density ,   δ + 1 D 2
where D is the average crystallite size. The microstrain is extracted from the full width at half maximum β of (2θω) diffraction peaks from Equation (8) [40].
Microstrain ,   ε + β 4 tan θ
A comparison between manufactured SnO2 nanoparticles and electrospun nanofibers reveals striking differences in their crystalline properties, with the electrospinning process significantly altering the microstructural characteristics. The manufactured SnO2 nanoparticles exhibit a relatively large average crystallite size of 47.23 nm, corresponding to the sharp, well-defined diffraction peaks observed in Figure 3a. In contrast, the pristine SnO2 nanofibers show a dramatically reduced crystallite size of 12.71 nm, consistent with the peak broadening observed in Figure 3b. This substantial reduction (approximately 73%) in crystallite size can be attributed to the rapid solidification during electrospinning and controlled crystallization during thermal treatment [41], which inhibits extensive crystal growth.
The influence of Sr doping on crystallite size follows a non-monotonic trend. With 1% Sr doping, the crystallite size slightly decreases to 11.71 nm, suggesting that low concentrations of Sr2+ ions impede crystal growth by creating lattice distortions [42]. Surprisingly, at 3% Sr doping, the crystallite size increases substantially to 28.08 nm, which deviates from the general trend observed in the XRD patterns. This anomalous increase might result from Sr-induced recrystallization processes or the formation of larger Sr-rich domains within the SnO2 matrix [43]. However, at 5% Sr doping, the crystallite size decreases dramatically to 7.14 nm, representing a 44% reduction compared to the pristine nanofibers. This significant decrease aligns with the severe peak broadening and suppression observed in Figure 3e and correlates with the substantial reduction in nanofiber diameter (148.72 nm) noted in Table 1.
The dislocation density, which quantifies crystal lattice imperfections, shows an inverse relationship with crystallite size across most samples [44]. The commercial SnO2 exhibits the lowest dislocation density (1.9 × 1015 m−2), consistent with its larger crystallites and higher crystallinity. The pristine SnO2 nanofibers show a higher dislocation density (1.37 × 1016 m−2), which further increases to 1.78 × 1016 m−2 with 1% Sr doping. At 3% Sr doping, the dislocation density decreases to 3.36 × 1015 m−2, mirroring the unexpected increase in crystallite size. The 5% Sr-doped sample exhibits the highest dislocation density (3.53 × 1016 m−2), indicating severe lattice distortion and defect formation at high Sr concentrations [45].
Microstrain values, which reflect lattice distortion, follow a similar pattern [46]. The commercial SnO2 shows minimal strain (0.23%), while pristine nanofibers exhibit higher strain (0.77%) due to their nanostructured nature. Sr doping progressively increases lattice strain, with values of 1.02% at 1% Sr doping and 1.42% at 5% Sr doping. The 3% Sr-doped sample again deviates from this trend with a lower strain value (0.41%), consistent with its larger crystallite size and lower dislocation density.
From Table 3, showing the calculated lattice parameters from XRD at (110) and (101) peaks, it is observed that the length of the edges (a, b, and c) of the unit cell is slightly changed, which indicates subtle lattice distortion upon Sr doping. Also, the ionic radius of Sr2+ (1.12 Å) is larger than Sn4+ (0.71 Å), which is reflected in the total cell volume, highest with 1% Sr doping (74.76 Å3) compared to pristine SnO2 nanofiber (71.1 Å3) [47], as Sr2+ ions substitute Sn4+ ions, indicating Sr incorporation. This substitution leads to local strain and a tendency to increase the interatomic distances, resulting in a larger unit cell volume. Conversely, for 5% of Sr doping, the volume (67.06 Å) of the unit cell is significantly contracted from the pristine one. It can be assumed that, when maintaining charge neutrality when Sr2+ substitutes Sn4+, oxygen vacancies are likely created in the SnO2 lattice. These oxygen vacancies can lead to a contraction of the lattice by reducing the electrostatic repulsion between the remaining oxide ions and allowing the Sn/Sr cations to move closer [48]. Additionally, from Table 2, it is evident that the lattice micro-strain of 5% Sr doping (1.42%) is significantly higher than pristine SnO2 NF (0.77%), with the smallest average crystallite size (7.14 nm), which could contribute to the non-monotonic volume change [40,44]. The interplay between these factors—the initial lattice expansion due to Sr substitution and the subsequent lattice contraction due to increased strain and reduced crystallite size—likely results in the non-monotonic trend observed in the unit cell volume.
These crystallographic parameters complement the morphological observations from SEM analysis (Figure 1). The initial increase in fiber diameter with 1% Sr doping, followed by a decrease at higher doping levels, can now be understood in terms of competing microstructural processes. The introduction of Sr2+ ions creates oxygen vacancies and lattice distortions that influence both crystal growth and fiber formation during electrospinning and calcination. The anomalous behavior of the 3% Sr-doped sample suggests a critical concentration threshold where the microstructural evolution mechanism changes, potentially involving phase segregation or recrystallization processes.

3.3. Electrochemical Studies

3.3.1. Cyclic Voltammetry (CV)

The electrochemical characteristics of the Sr-doped SnO2 nanofibers were investigated using cyclic voltammetry. Figure 4 presents the CV curves of pristine and Sr-doped SnO2 nanofibers recorded at a scan rate of 50 mV/s, while Table 4 summarizes the quantitative electrochemical parameters derived from these measurements.
The CV profiles exhibit quasi-rectangular shapes with distinctive redox humps, characteristic of both electric double-layer capacitance and pseudocapacitive behavior [30,49]. This dual charge storage mechanism arises from the combined effects of ion adsorption at the electrode–electrolyte interface and surface redox reactions involving oxygen vacancies in the SnO2 lattice [50]. The enclosed area within the CV loops, which represents the total charge storage capacity, varies significantly with Sr doping concentration, indicating a profound influence of Sr incorporation on the electrochemical performance. Pseudo-capacitance is a Faradaic energy storage mechanism arising from rapid, reversible redox reactions occurring at or near the electrode surface. This process involves electrosorption/electrodesorption coupled with charge transfer, but critically, it proceeds without any bulk phase transformation during charge/discharge cycles [51,52]. The pseudo-capacitance (Cϕ) is defined by the change in charge (Q) with respect to the potential (V), specifically dQ/dV. Unlike battery processes involving bulk chemical changes, pseudo-capacitance in our system with SnO2 nanofibers and a KOH electrolyte is attributed to fast surface redox reactions, potentially affecting the intercalation of alkali metal cations (specifically K+ from the KOH electrolyte) into the SnO2 structure. This leads to a charge storage behavior where the electrode potential exhibits a linear dependence on the accumulated charge and is proportional to the surface coverage of electroactive ions, forming a reversible functionalized molecular layer. The possible reaction is given below [51]:
SnO2 + K+ + e ↔ SnOOK
The potential of the electrode has a linear dependence on the charge [52]. From Figure S1, the linear relation between the scan rate of CV and peak current is explicitly observed for 1% Sr-doped SnO2 NF. The relationship between peak current (ip) and scan rate (v) can be described by the following power law equation [51].
ip + avb
log (ip) + log (a) + blog (v)
where a is the constant and b is the slope value. The b value is close to 1 (0.7), indicating a faster surface-controlled redox reaction than a diffusion-induced reaction from the bulk of the material with dominant capacitive behavior.
The current response of commercial SnO2 nanoparticles (NP) is almost identical to 5% Sr-doped SnO2 NF. From Table 4, the commercial SnO2 NP exhibits the lowest specific capacitance (22.5 F/g) and total accumulated charge (1.8 × 10−5 C) compared to SnO2 NFs (doped or undoped), which explicitly supports the superiority of the nanofibrous morphology of SnO2. The pristine SnO2 nanofibers display a relatively modest current response with a total charge of 4.35 × 10−5 C and a specific capacitance of 54.4 F/g. The specific capacitance was calculated using Equation (1) described in Section 2.4. Upon 1% Sr doping, there is a remarkable enhancement in electrochemical activity [53,54], manifested by the substantially larger CV loop area. This is quantitatively confirmed in Table 4, which shows that the 1% Sr-doped sample exhibits a total charge of 8.45 × 10−5 C, representing a 94% increase compared to the pristine sample. Correspondingly, the specific capacitance nearly doubles to 106 F/g, indicating significantly improved charge storage capability. As Sr doping increases to 3%, the CV curve area moderately decreases, with the total charge reducing to 6.80 × 10−5 C and specific capacitance to 85 F/g. This trend continues with 5% Sr doping, where the electrochemical performance deteriorates further, with total charge (3.76 × 10−5 C) and specific capacitance (47.1 F/g) values falling below those of the pristine SnO2 nanofibers. This non-monotonic dependence of electrochemical properties on Sr content parallels the trends observed in the morphological and crystallographic analyses.
The superior electrochemical performance of the 1% Sr-doped sample can be correlated with its unique structural and morphological characteristics previously identified. The slight increase in fiber diameter (288.69 nm) compared to pristine SnO2 (259.04 nm), as shown in Table 1, combined with reduced crystallite size (11.71 nm vs. 12.71 nm) and increased dislocation density (1.78 × 1016 vs. 1.37 × 1016 m−2) from Table 2, creates an optimal balance of surface area and defect concentration. The increased microstrain (1.02% vs. 0.77%) and the formation of secondary phases observed in the XRD pattern (Figure 3c) suggest the presence of abundant oxygen vacancies at 1% Sr doping, which serve as active sites for charge transfer and storage.
Conversely, the deterioration in electrochemical performance at higher Sr doping levels (3% and 5%) can be attributed to the compromised structural integrity observed in both the SEM images (Figure 1d) and crystallographic data. The excessive lattice distortion and defect concentration at 5% Sr doping, evidenced by the highest microstrain (1.42%) and dislocation density (3.53 × 1016 m−2), likely impedes charge transport within the nanofiber network. Additionally, the significant reduction in fiber diameter (148.72 nm) and the emergence of structural discontinuities may reduce the effective electrode-electrolyte contact area and disrupt electron conduction pathways. These electrochemical findings establish a clear structure–property relationship, demonstrating that moderate Sr doping (1%) optimizes the balance between creating beneficial defects that enhance charge storage capacity and maintaining the structural integrity necessary for efficient charge transport.

3.3.2. Electrochemical Impedance Spectroscopy (EIS)

Electrochemical impedance spectroscopy measurements were conducted to further elucidate the charge transfer dynamics and interfacial properties of the Sr-doped SnO2 nanofibers. Figure 5 presents the Nyquist plots for the different samples, while Table 5 provides the quantitative parameters derived from fitting these plots to an equivalent Randle circuit model [55].
The Nyquist plots exhibit linear regions in the low-frequency domain, characteristic of diffusion-controlled processes, and semicircular arcs in the high-frequency region, representing charge transfer resistance at the electrode–electrolyte interface. A notable trend is immediately apparent: the 1% Sr-doped sample displays the smallest semicircle diameter, indicating significantly enhanced charge transfer kinetics [56] compared to other samples. The fitted equivalent circuit parameters in Table 5 provide quantitative confirmation of these observations. Commercial SnO2 NP exhibits the charge transfer resistance of 859 Ohm-cm2, lower than pristine SnO2 NF but higher than 1% (404 Ohm-cm2) and 3% (450 Ohm-cm2) Sr-doped SnO2 NF. But the double-layer capacitance of SnO2 NP (7.06   ×   10 5 F) is the lowest among all of the conditions, supporting the specific capacitance derived from V (Figure 4). The diffusion-induced impedance, the Warburg impedance (3   ×   10 4 ohm-cm2) is also higher than in other nanofibrous morphologies. These values indicate the limitations of nanoparticles compared to nanofibers of SnO2 in light energy harvesting. The charge transfer resistance (Rct) of the pristine SnO2 nanofibers is remarkably high at 1350 Ω-cm2, indicating substantial impedance to electron transfer. Upon 1% Sr doping, this resistance drastically decreases to 404 Ω-cm2, representing a 70% reduction. This significant improvement in charge transfer kinetics aligns with the enhanced current response observed in the CV measurements (Figure 4) and the increased total charge and specific capacitance values (Table 4).
The Warburg impedance (ZW), which characterizes ion diffusion resistance in the electrode material, follows a similar trend. The 1% Sr-doped sample exhibits the lowest ZW value (3.84 × 103 Ω-cm2), approximately 64% lower than the pristine SnO2 (1.06 × 104 Ω-cm2). This indicates a greatly enhanced ion diffusion capability, which is critical for efficient electrochemical processes, including photocatalysis. The double-layer capacitance (CPEdl) increases substantially from 1.42 × 10−4 F for pristine SnO2 to 3.8 × 10−4 F for the 1% Sr-doped sample, suggesting a higher electrochemically active surface area. The solution resistance (Rs) also shows improvement with moderate Sr doping, decreasing from 3.73 Ω-cm2 for pristine SnO2 to 2.87 Ω-cm2 and 1.73 Ω-cm2 for 1% and 3% Sr-doped samples, respectively.
At higher Sr doping levels, the electrochemical parameters show varying trends. The 3% Sr-doped sample maintains improved performance compared to pristine SnO2 but does not match the optimal properties of the 1% Sr-doped sample. The 5% Sr-doped nanofibers exhibit complex behavior: while Rct (450 Ω-cm2) and ZW (5.71 × 103 Ω-cm2) remain lower than the pristine sample, other parameters such as CPEdl (6.97 × 10−5 F) deteriorate significantly, and Rs increases to 3.58 Ω-cm2.
These EIS results provide deeper insight into the electrochemical behavior observed in the CV analysis (Figure 5 and Table 5). The substantially reduced charge transfer resistance and Warburg impedance of the 1% Sr-doped sample explain its superior charge storage capability [57]. The increased CPEdl value further suggests enhanced surface reactivity [58], which is essential for photocatalytic applications. The trends in electrochemical parameters correlate remarkably well with the previously observed morphological and structural properties. The optimal performance of the 1% Sr-doped sample can be attributed to its unique combination of increased fiber diameter (288.69 nm, Table 1), moderately reduced crystallite size (11.71 nm, Table 2), and increased defect concentration (dislocation density 1.78 × 1016 m−2, Table 2). The increased microstrain (1.02%) and the formation of secondary phases observed in XRD (Figure 3c) likely create abundant oxygen vacancies that serve as active sites for charge transfer and storage, explaining the dramatically reduced Rct and ZW values.
The deterioration in electrochemical performance at 5% Sr doping, despite having the smallest crystallite size (7.14 nm) and highest defect concentration, can be attributed to excessive lattice distortion [59] and structural discontinuities observed in SEM (Figure 1e) and XRD (Figure 3e). These structural deficiencies likely create barriers to electron transport within the nanofiber network, offsetting the potential benefits of increased surface area and defect sites. The analysis reveals that Sr doping (1%) creates an optimal balance between beneficial defect creation and structural integrity preservation, significantly enhancing charge transfer kinetics and ion diffusion.

3.3.3. Mott–Schottky (MS) Plot

To gain deeper insights into the semiconductor properties and electronic band structure of the Sr-doped SnO2 nanofibers, Mott–Schottky (MS) analysis was performed. The Mott–Schottky (MS) analysis is a valuable technique for investigating the electronic characteristics of semiconductors. It enables the direct determination of flat band potential and charge carrier density through capacitance-voltage measurements [26]. Figure 6 presents the MS plots for different Sr doping concentrations, while Table 6 summarizes the key parameters extracted from these measurements. From the slope of the MS plot, we can determine the charge carrier density. The parameters are extracted following Equation (11) [60].
N D   + 2 ε 0 ε r e × s l o p e
where ND represents charge carrier density, ε 0 is the dielectric permittivity of air, and ε r corresponds to relative permittivity (for SnO2, 18 is considered). The MS plots display positive slopes for all samples, confirming the n-type semiconductor nature of the SnO2 nanofibers, where electrons serve as the majority charge carriers. The flat band potential (Vfb), which approximates the position of the conduction band edge, shows a systematic shift toward more negative values with increasing Sr doping concentration, moving from −0.45 V for pristine SnO2 to −0.53 V, −0.68 V, and −0.62 V for 1%, 3%, and 5% Sr-doped samples, respectively. This negative shift indicates a favorable energetic band alignment [61] for enhanced reduction reactions, which is beneficial for photocatalytic applications.
The most remarkable finding from the MS analysis is the extraordinary enhancement in charge carrier density (ND) observed with Sr doping. Commercial SnO2 NP exhibits a flat band potential of − 0.48 V, more negative than pristine SnO2 NF (−0.45 V). The charge carrier density of SnO2 NP ( 1.20   ×   10 22   c m 3 ) (Table 6) is 130-fold larger than pristine SnO2 NF ( 9.24   ×   10 19   c m −3). This discrepancy can be explained by the density of surface states (Figure 6) and some secondary phases from XRD peaks of SnO2 NP (Figure 3). The density of surface states for SnO2 NP is higher than pristine SnO2 NF. Also, from XRD peaks, some secondary phases are formed, which might create any defect state or trap centers trapping charge carriers, which increases the charge carrier density.
The pristine SnO2 nanofibers exhibit a carrier density of 9.24 × 1019 cm−3, which increases dramatically by approximately three orders of magnitude to 5.17 × 1022 cm−3 for the 1% Sr-doped sample. This represents an unprecedented 560-fold increase in charge carrier concentration, which explains the superior electrochemical performance [62] observed in CV (Figure 4, Table 4) and EIS (Figure 5, Table 5) measurements. The slope of the linear region in the MS plots, which is inversely proportional to carrier density, decreases substantially from 8.4 × 1010 for pristine SnO2 to 1.5 × 108 for 1% Sr-doped SnO2, confirming the dramatic increase in carrier concentration. At higher Sr concentrations, the carrier density decreases progressively to 3.23 × 1020 cm−3 for 3% Sr doping and further to 2.51 × 1019 cm−3 for 5% Sr doping, approaching the value of the pristine sample. This non-monotonic trend in carrier density correlates with the observed changes in charge transfer resistance (Rct) in the EIS analysis (Table 5), where Rct increases for 3% Sr doping before decreasing again at 5% Sr doping.
The electronic properties revealed by MS analysis provide a fundamental explanation for the superior electrochemical and photocatalytic performance of the 1% Sr-doped SnO2 nanofibers. The exceptionally high carrier density, combined with the negative shift in flat band potential, creates optimal conditions for efficient charge separation and transfer during photocatalytic reactions. The increased carrier concentration can be attributed to the oxygen vacancies [7,37,63] generated by the substitution of Sn4+ with Sr2+ ions, which act as electron donors in the SnO2 lattice. The density of surface states [64,65], indicated by the circled regions in the negative potential domain of Figure 6, also shows variations with Sr doping concentration, with the 1% Sr-doped sample exhibiting distinct surface states that may serve as catalytic sites. The decrease in carrier density at higher Sr concentrations (3% and 5%) may result from several competing mechanisms. Excessive Sr doping could lead to the formation of charge-compensating defects that trap carriers [66,67], clusters of Sr ions that create potential barriers for charge transport or increased structural disorder that reduces carrier mobility. These effects would counteract the beneficial impact of oxygen vacancies on carrier concentration.

3.3.4. Linear Sweep Voltammetry (LSV)

Linear sweep voltammetry and Tafel analysis were conducted to further evaluate the electrocatalytic activity and reaction kinetics of the Sr-doped SnO2 nanofibers. The LSV curves in Figure 7a show a clear influence of Sr doping on the current response of the SnO2 nanofibers. The 1% Sr-doped sample exhibits the highest current density across the entire potential range, followed by the 3% Sr-doped sample. Both the pristine and 5% Sr-doped samples display significantly lower current densities, with the 5% Sr-doped sample showing the poorest performance. This trend directly correlates with the electrochemical properties observed in CV (Figure 4) and EIS (Figure 5) measurements, reinforcing the superior electron transfer capabilities of the 1% Sr-doped nanofibers [68].
The Tafel plots in Figure 7b provide critical insights into the reaction kinetics and catalytic efficiency of the samples. The Tafel slope [69], which represents the voltage increase required to enhance the current density by one order of magnitude, is a key indicator of electrocatalytic activity, with lower values indicating more efficient catalysis. The Tafel plot is derived from LSV, where the potential values are referenced to the reversible hydrogen electrode (RHE) through an RHE calibration. Equation (4) [28] is used to convert potential from Ag/AgCl to RHE, and the overpotential (η) was calculated [29] at different current densities using Equation (5) as described in Section 2.4.
The overpotential values of 0, 1, 3, and 5% of Sr-doped SnO2 NFs are −2.63, −2.43, −2.46, and −3.03 V at the current density of 10 mA/cm2. 1% Linear regions of the Tafel plots were analyzed using the Tafel equation: η + b log (j) + a, where η represents overpotential, j denotes current density, and b is the Tafel slope. The extracted slope values are presented in Table 7. The ascending order of the Tafel slope is 1% < 5% < 3% < 0%-Sr doped SnO2 NF.
The current-voltage characteristics in Figure 7a demonstrate that the 1% Sr-doped SnO2 nanofibers exhibit the highest current response at any given voltage, followed by 3% Sr-doped samples. This suggests enhanced electrical conductivity and charge transfer properties in the low-concentration Sr-doped samples. The 5% Sr-doped nanofibers, however, show the lowest current response, indicating that excessive Sr doping may introduce defects or phase segregation that impede charge transport.
The Tafel plots in Figure 7b and the derived Tafel slopes in Table 7 provide crucial insights into the reaction kinetics. The commercial SnO2 NP exhibits the highest Tafel slope (2.02 V/dec), suggesting sluggish electron transfer kinetics. In contrast, all nanofiber samples demonstrate significantly lower Tafel slopes, with the 1% Sr-doped sample showing the lowest value (0.47 V/dec), followed by 5% Sr-doped (0.49 V/dec), 3% Sr-doped (0.62 V/dec), and pristine SnO2 NF (0.84 V/dec).
The dramatic reduction in the Tafel slope from commercial NPs to nanofibers, particularly with Sr doping, indicates that the nanofiber morphology and Sr incorporation facilitate faster electron transfer kinetics. A lower Tafel slope corresponds to a higher exchange current density and thus more favorable reaction kinetics [70,71].
As Sr doping increases to 3%, the Tafel slopes increase slightly to 0.62 V/dec and for 5%, reduce to 0.49 V/dec, respectively. These values remain lower than the pristine sample but are higher than the optimal 1% Sr-doped nanofibers. This trend suggests that excessive Sr doping, while still beneficial compared to undoped SnO2, does not maintain the optimal catalytic efficiency achieved at 1% doping concentration.
The electrocatalytic performance observed in the LSV and Tafel analyses can be directly linked to the electronic and structural properties previously characterized. The exceptionally high charge carrier density of the 1% Sr-doped sample (5.17 × 1022 cm−3, Table 6) provides abundant mobile electrons for catalytic reactions, contributing to its superior current response and lower Tafel slope. The reduced charge transfer resistance (404 Ω-cm2, Table 5) further facilitates efficient electron transfer at the electrode–electrolyte interface. The slightly higher Tafel slopes for 3% and 5% Sr-doped samples correlate with their lower carrier densities (3.23 × 1020 and 2.51 × 1019 cm−3, respectively) and higher charge transfer resistances compared to the 1% Sr-doped sample. The anomalous behavior of the 3% Sr-doped sample, particularly its increased crystallite size (28.08 nm, Table 2) and charge transfer resistance (1000 Ω-cm2, Table 5), aligns with its intermediate performance in the LSV and Tafel analyses. The poor current response of the 5% Sr-doped sample, despite having a relatively low Tafel slope, suggests that while the intrinsic catalytic activity per active site may be reasonable, the overall performance is limited by other factors such as reduced electrochemically active surface area (lower CPEdl values, Table 5) and structural discontinuities observed in SEM (Figure 1d).

3.3.5. Chronoamperometry (CA)

To further investigate the charge transport dynamics and diffusion properties of the Sr-doped SnO2 nanofibers, chronoamperometry (CA) measurements were conducted. Figure 8a displays the current response over time following a potential step, while Figure 8b presents the corresponding Cottrell plots. Table 8 summarizes the diffusion coefficients calculated using the Cottrell equation.
The CA curves in Figure 8a reveal significant differences in the initial current surge and subsequent decay behavior among the samples. The 1% Sr-doped nanofibers exhibit a remarkably higher initial current density (approximately 3.2 mA/cm2) compared to other samples, indicating superior charge storage capacity and electrode kinetics. This initial current is approximately three times higher than that of the pristine SnO2 nanofibers (approximately 1.1 mA/cm2) and 45 times higher than commercial SnO2 NP (approximately 0.07 mA/cm2); it also significantly higher than both 3% and 5% Sr-doped samples, which show initial currents of approximately 1.2 mA/cm2 and 0.6 mA/cm2, respectively. The enhanced initial current response of the 1% Sr-doped sample aligns with its superior capacitive properties observed in CV measurements (Figure 3, Table 3).
The Cottrell plots in Figure 8b, which display current versus t−1/2, provide quantitative insights into the diffusion behavior [72] of the electroactive species within the nanofiber electrodes. The linear relationship between current and t−1/2 confirms that the electrochemical processes are primarily diffusion-controlled. The slope of the Cottrell plot is directly proportional to the diffusion coefficient, with steeper slopes indicating more rapid diffusion of electroactive species. The Cottrell plot is calculated using Equation (3) [27] discussed in Section 2.4.
The diffusion coefficient is a critical parameter for photocatalytic applications, as it directly influences the ability of photogenerated charge carriers to reach reactive sites on the catalyst surface before recombination [73,74]. A higher diffusion coefficient in the context of photocatalytic activity means the reactants (e.g., pollutants) can move quickly through the solution to the catalyst surface. The reaction rate will be higher because of the faster diffusion rate and improved mass transfer. Due to the reduced diffusion length between the catalyst and bulk pollutant, the limited mass transfer can be largely reduced, which may accelerate photocatalytic degradation [75].
As shown in Table 8, the 1% Sr-doped sample exhibits the steepest slope (3.04 × 10−5), approximately 2.8 times higher than that of pristine SnO2 (1.09 × 10−5). This translates to a significantly higher diffusion coefficient (D) of 8.78 × 10−10 cm2s−1 for the 1% Sr-doped sample compared to 1.13 × 10−10 cm2s−1 for pristine SnO2, representing an almost 8-fold enhancement and 24-fold higher than commercial SnO2 NP (3.62 × 10−11 cm2s−1). This substantial improvement in diffusion properties is a critical factor contributing to the superior electrochemical and photocatalytic performance of the 1% Sr-doped nanofibers. Interestingly, the 3% Sr-doped sample shows a moderate enhancement in diffusion coefficient (1.84 × 10−10 cm2s−1), approximately 63% higher than the pristine sample but significantly lower than the 1% Sr-doped sample. The 5% Sr-doped sample exhibits a diffusion coefficient (1.03 × 10−10 cm2s−1) slightly lower than the pristine sample, indicating compromised diffusion properties at high Sr concentrations.
The diffusion behavior observed in the CA measurements correlates well with the electrochemical impedance results (Figure 5, Table 5), particularly the Warburg impedance (ZW) values, which characterize diffusion resistance. The 1% Sr-doped sample showed the lowest ZW (3.84 × 103 Ω-cm2), consistent with its highest diffusion coefficient. Similarly, the higher ZW values for other samples align with their lower diffusion coefficients. The significantly enhanced diffusion properties of the 1% Sr-doped nanofibers suggest that photogenerated electrons and holes can more efficiently migrate to the surface to participate in redox reactions, explaining the superior photocatalytic performance observed for this sample.

3.4. Optical Properties

The UV-visible absorption technique was employed to analyze the optical characteristics of undoped and 1% Sr-doped SnO2 nanofibers (NFs) within the 300–800 nm range (Figure 9). The absorption spectra, influenced by factors such as oxygen vacancies, morphology, band gap, and impurity centers, exhibited a cut-off wavelength between 320 and 360 nm, indicative of electron photo-excitation from the valence to the conduction band [7]. Notably, the absorption peak displayed a subtle red shift (shift towards higher wavelengths) with increased Sr doping, resulting in a reduction in the estimated energy band gap, as determined from the Tauc plot [76]. Using the Tauc plot, the band gap values were calculated to be 3.77 eV for pristine and 3.28 eV for 1% Sr-doped SnO2 NFs. This band gap reduction is attributed to a shift in the conduction band minimum to lower energies [77] and the introduction of additional energy states within the SnO2 band gap due to oxygen vacancies formed to maintain charge neutrality upon Sr2+ incorporation into the SnO2 lattice.

3.5. Photocatalytic Degradation

3.5.1. UV Light as a Light Source

Following the comprehensive characterization of morphological, structural, and electrochemical properties, the functional performance of the Sr-doped SnO2 nanofibers was evaluated through photocatalytic degradation of methylene blue (MB) dye under UV light irradiation. Figure 10 presents the time-dependent absorption spectra of MB in the presence of different doping levels of SnO2 NF photocatalysts, while Table 9 summarizes the degradation efficiencies achieved after 180 min of UV irradiation.
The absorption spectra in Figure 10 display the characteristic peaks of MB with maximum absorption at approximately 664 nm. The progressive decrease in absorption intensity with increasing irradiation time indicates the gradual degradation of the dye molecules. The rate and extent of this degradation vary significantly among the different catalysts, revealing their relative photocatalytic activities. As shown in Table 9, the commercial SnO2 nanoparticles exhibit poor photocatalytic performance, achieving only 16% MB degradation after 180 min of UV irradiation. In contrast, the pristine SnO2 nanofibers demonstrate substantially higher activity with 61% degradation efficiency. This nearly four-fold enhancement highlights the inherent advantages of nanofiber morphology, which provides higher surface area, enhanced light harvesting, and improved charge transport properties compared to conventional nanoparticles. The 1% Sr-doped SnO2 nanofibers exhibit the highest photocatalytic activity among all samples, achieving 69% MB degradation after 180 min. This represents a 13% improvement over pristine SnO2 nanofibers and reinforces the significant beneficial impact of moderate Sr doping on catalytic performance. The enhanced photocatalytic activity of the 1% Sr-doped sample aligns perfectly with its superior electrochemical properties observed in previous analyses, particularly its exceptionally high-charge carrier density (5.17 × 1022 cm−3, Table 6), low-charge transfer resistance (404 Ω-cm2, Table 5), and enhanced diffusion coefficient (8.78 × 10−10 cm2s−1, Table 8).
Interestingly, the photocatalytic activity exhibits a non-monotonic relationship with Sr doping concentration. The 3% Sr-doped nanofibers show reduced performance (58% degradation) compared to both 1% Sr-doped and pristine samples, while the 5% Sr-doped nanofibers demonstrate intermediate activity (63% degradation). This complex trend mirrors the variations observed in the electrochemical properties, particularly the charge transfer resistance (Table 5) and carrier density (Table 6), which also showed non-monotonic dependence on Sr content. The degradation profiles visible in Figure 10 reveal additional insights into the photocatalytic mechanism. All samples show more rapid degradation during the initial 60 min, followed by slower decay in the later stages. The 1% Sr-doped sample (Figure 10c) exhibits particularly effective degradation in the early phase, consistent with its enhanced charge transfer kinetics and lower Tafel slope (0.47 V/dec, Table 7).
Figure 11 provides a comprehensive view of the photocatalytic degradation kinetics of methylene blue (MB) using various SnO2-based photocatalysts. Relative concentration profiles (Ct/C0) in Figure 11a and the corresponding degradation efficiency curves in Figure 11b offer valuable insights into both the rate and extent of the photocatalytic process over the 180 min irradiation period. As illustrated in Figure 11a, the commercial SnO2 nanoparticles demonstrate minimal photocatalytic activity, with the Ct/C0 value decreasing only slightly from 1.0 to approximately 0.84 after 180 min, corresponding to a mere 16% degradation efficiency. This poor performance can be attributed to the large crystallite size (47.23 nm, from Table 2) and limited surface area of commercial nanoparticles, resulting in fewer active sites for catalytic reactions and inefficient charge separation. In contrast, all the electrospun SnO2 nanofibers exhibit significantly enhanced photocatalytic performance. The pristine SnO2 nanofibers show a steady decrease in relative concentration to approximately 0.39 after 180 min, achieving 61% degradation efficiency. This substantial improvement over commercial nanoparticles highlights the inherent advantages of the nanofiber morphology, including higher surface-to-volume ratio, enhanced light harvesting, and improved charge transport along the one-dimensional fiber structure.
The 1% Sr-doped SnO2 nanofibers display the most remarkable photocatalytic activity, with the relative concentration decreasing more rapidly and reaching the lowest final value of approximately 0.31 (Figure 11a), corresponding to 69% degradation efficiency after 180 min. Notably, the degradation curve for the 1% Sr-doped sample shows a distinct kinetic profile compared to other samples, with a rapid initial decrease in the first 45 min, followed by a steady decline throughout the experimental duration. This suggests that Sr doping not only enhances the overall catalytic efficiency but also accelerates the initial reaction rate. The degradation kinetics of the 3% and 5% Sr-doped samples exhibit intermediate performance between the pristine and 1% Sr-doped nanofibers. The 3% Sr-doped sample achieves approximately 58% degradation, while the 5% Sr-doped sample reaches approximately 63% degradation after 180 min. Interestingly, both samples show similar degradation profiles to the 1% Sr-doped sample during the first 60 min but diverge in the later stages, indicating that while all Sr-doped samples benefit from enhanced initial reaction rates, only the 1% Sr-doped sample maintains superior performance throughout the entire process.
The degradation efficiency curves in Figure 11b reveal distinct phases in the photocatalytic process. A minimal degradation phase is observed during the first 30 min, likely representing the surface adsorption equilibrium period [1,18]. This is followed by a rapid degradation phase between 30 and 60 min, where significant differences between the catalysts become apparent. The 1% Sr-doped sample shows the steepest increase in degradation efficiency during this phase, reaching approximately 35% at 60 min compared to approximately 25% for other nanofiber samples. In the final phase (60–180 min), all samples show continued degradation but at varying rates, with the 1% Sr-doped sample maintaining the highest degradation rate.
From 40 to 120 min, a plateau is observed where very subtle degradation of MB occurs. But from 120 min to 180 min, a sharp MB degradation is evident. The plateau observed within 40 to 120 min can be described by the Langmuir adsorption isotherm model [78,79]. This model can be mathematically expressed as
q + Q C C + k e q 1
where q is the amount of adsorbate (MB dye) adsorbed per unit of area, Q is the amount of adsorbate adsorbed at surface saturation for a given concentration (C), and keq is the Langmuir equilibrium constant.
At the beginning of the experiment, the SnO2 surface has many vacant adsorption sites. Methylene blue molecules from the solution rapidly adsorb onto these sites, as predicted by the Langmuir isotherm model. The photocatalytic degradation proceeds at a reasonable rate because the dye molecules are in close contact with the photocatalyst surface, where the reactive species (electrons and holes) are generated. As the reaction proceeds and more dye molecules are adsorbed and potentially degraded, the concentration of methylene blue in the solution decreases [80]. According to the Langmuir isotherm [78], as the concentration (C) decreases, the fraction of surface coverage also decreases as q decreases, but not linearly. Initially, a decrease in C leads to a significant decrease in surface coverage. By 40 min, the SnO2 surface may be approaching saturation with methylene blue molecules covering the active sites of the catalytic surface [81,82]. Even though the photocatalyst is still active under light irradiation, there are fewer available adsorption sites for new dye molecules to bind and be degraded. The rate of degradation becomes limited by the availability of adsorbed dye molecules on the surface. The subsequent sharp degradation from 120 to 180 min might be due to the desorption or further degradation of MB, freeing up active sites for fresh dye molecules to adsorb and undergo photocatalysis.
A pseudo-first-order kinetic model is fitted following the equation of the simplified Langmuir–Hinshelwood model [7].
ln (C0/Ct) + kt
where Ct and C0 are the concentration of MB after and before the illumination time, respectively, and k is the rate constant. For each experiment, the plots of ln (C0/Ct) vs. irradiation time (t) for MB are presented in Figure 12, respectively, and their calculated rate constant values are mentioned in Table 10. From Table 10, it is clear that the 1% Sr-doped SnO2 has the highest degradation rate (0.00125 min−1) constant with correlation constant, an R2 value of 0.87, compared to the undoped (6.32 × 10−4 min−1, R2 = 0.83) and 3% (8.76 × 10−4 min−1, R2 = 0.98) and 5% Sr-doped SnO2 (5.65 × 10−4 min−1, R2 = 0.99). These findings suggest that the degradation process best fits a pseudo-first-order kinetics model. Even in a plateau, 1% Sr-doping exhibits a better reaction rate than other conditions. The high correlation constant, R2, shown in Table 10, showing MB dye degradation approaching unity, indicates that the catalytic degradation process is dominant on the surface [81] of the SnO2 photocatalyst.

3.5.2. Visible Light as a Light Source

While the UV light photocatalytic experiments demonstrated the significant benefits of the nanofiber morphology and Sr doping, the performance under visible light (solar simulator) provides even more compelling evidence for the transformative impact of strategic Sr doping on SnO2 photocatalysts. Figure 13 presents the time-dependent absorption spectra of methylene blue (MB) degradation under simulated solar irradiation, while Table 11 summarizes the degradation efficiencies achieved after 180 min.
The most striking observation from Table 11 is the remarkable enhancement in photocatalytic activity for the 1% Sr-doped SnO2 nanofibers under visible light, achieving an impressive 84.74% MB degradation after 180 min. This represents a dramatic improvement compared to its performance under UV light (69%, Table 10), suggesting that Sr doping fundamentally alters the light absorption properties of SnO2, extending its photocatalytic activity into the visible region of the spectrum. This finding is particularly significant considering that pristine SnO2 NF is known to have a wide bandgap (>3 eV) that limits its activation to UV light [83]. Interestingly, the commercial SnO2 nanoparticles also show enhanced performance under visible light (70% degradation) compared to UV light (16% degradation), though the improvement mechanism likely differs from that of the Sr-doped nanofibers. The pristine SnO2 nanofibers maintain similar performance under both UV and visible light (61% degradation), suggesting minimal bandgap modification in the absence of dopants.
The absorption spectra in Figure 13 reveal distinct degradation kinetics for each photocatalyst. The commercial SnO2 (Figure 13a) shows a significant decrease in absorption intensity between 0 and 60 min, with slower degradation thereafter. The pristine SnO2 nanofibers (Figure 13b) exhibit more uniform degradation throughout the 180 min. Most notably, the 1% Sr-doped sample (Figure 13c) demonstrates exceptionally rapid degradation within the first 15 min, followed by continued progressive degradation throughout the experiment.
The observed visible light photocatalytic activity of commercial SnO2 NP surpasses what is typically expected, despite its high bulk purity. This enhanced performance can be attributed to multiple factors. Surface hydroxyl groups and inherent oxygen vacancies may create defect levels within the band gap, enabling some lower-energy photon absorption [84]. The key takeaway from the XRD analysis (Figure 2) reveals the presence of secondary crystalline phases (112), (024) alongside primary peaks of SnO2. These additional phases, potentially possessing narrower band gaps and the presence of secondary crystalline phases in the commercial SnO2, as indicated by XRD, likely form heterojunctions with SnO2. These interfaces can create a more favorable energetic landscape for charge separation and transfer upon visible light excitation of the secondary phases, ultimately leading to the observed enhanced photocatalytic performance under visible light [85]. Electrochemical analysis from the M-S plot (Figure 6) also revealed that the density of surface states of commercial SnO2 is slightly higher than the pristine nanofibrous morphology of SnO2, indicating trapping sites or mid-gap states [86] potentially introducing energy levels within the energy band gap, and narrowing the energy band gap.

3.5.3. Photocatalytic Mechanism

SnO2 acts as an efficient photosensitizer because of its ability to absorb UV light (λ < 380 nm) due to its wider energy bandgap (3.2–3.6 eV) [87,88]. Previously, many research groups established SnO2 as a photosensitizer for successful photocatalytic degradation [7,63,87,89,90]. The established effectiveness of tin dioxide (SnO2) as a photosensitizer for photocatalytic degradation stems from its combination of high optical transparency, infrared reflectivity, and robust chemical and thermal stability [91]. As a prominent photocatalyst, SnO2 shares a rutile-type crystal structure with TiO2 [88]. In its unit cell, each Sn (IV) ion is octahedrally coordinated by six oxygen ions, while each oxygen is bonded to three Sn (IV) ions, forming a (6, 3) network [88,92]. This octahedral arrangement is considered crucial for high photocatalytic activity [93] as it enhances the mobility of photogenerated electron-hole pairs, increasing their likelihood of participating in surface reactions. Furthermore, the inherent defects within the SnO2 structure contribute to its electrical conductivity by narrowing the energy band gap [94], thus improving its semiconducting properties. The introduction of additional oxygen anions [95] facilitates the oxidation of Sn2+ to Sn4+, maintaining charge neutrality within the material. Figure 14 illustrates a schematic diagram of the oxidation process taking place on the SnO2 crystal surface. Figure 15 explicitly describes the reaction steps of Sr-doped SnO2 NF as a photocatalyst on MB dye degradation.
The enhanced photocatalytic performance of Sr-doped SnO2 can be attributed to several factors. Firstly, Sr incorporation can effectively suppress charge carrier recombination by trapping electrons from the SnO2 conduction band. Secondly, the lattice mismatch between Sr2+ (1.12 Å) and Sn4+ (0.71 Å) ions promotes the formation of oxygen vacancies, which act as charge separation centers. Finally, the reduced crystallite size of Sr-doped SnO2 results in a larger surface area, facilitating enhanced adsorption of organic pollutants and consequently improving photocatalytic activity [96].
Upon UV light irradiation, photogenerated electron-hole pairs are generated within the SnO2 lattice. In undoped SnO2, rapid electron-hole pair recombination limits their participation in photocatalytic reactions. Sr doping effectively mitigates this recombination by trapping electrons, enabling their efficient separation. Furthermore, Sr incorporation induces the formation of oxygen vacancies, which act as charge separation centers. Concurrently, photogenerated holes in the valence band of SnO2 oxidize water molecules to form hydroxyl radicals, while trapped electrons react with oxygen to generate superoxide radicals, which subsequently undergo multi-electron reduction pathways to form additional hydroxyl radicals [90,97]. Oxygen vacancies are easily formed in SnO2 compared to Sn vacancies due to their lower formation energy [98]. These oxygen vacancies can exhibit a dual role: trapping charge carriers in the valence band or transferring trapped electrons to the conduction band under UV irradiation [63]. The generated hydroxyl or superoxide radicals effectively oxidize organic pollutants, ultimately converting them into non-toxic substances such as CO2, O2, or H2O [99]. The step-by-step reactions associated with this degradation mechanism of MB dye with Sr-doped SnO2 are provided below in Figure 15.

4. Conclusions

This comprehensive study demonstrates that strategic strontium doping of SnO2 nanofibers via electrospinning technology creates highly effective photocatalysts for environmental remediation applications. The investigation reveals that precise control of dopant concentration is crucial for optimizing photocatalytic performance, with 1% Sr-doped SnO2 nanofibers exhibiting exceptional degradation efficiency for methylene blue under both UV (69%) and visible light (84.74%) conditions. The superior photocatalytic activity of the 1% Sr-doped sample stems from synergistic effects across multiple scales. At the morphological level, Sr doping at 1% concentration maintains an ideal nanofiber network with optimal fiber diameter (288.69 nm), providing efficient one-dimensional electron transport pathways while preserving structural integrity. Crystallographically, moderate Sr doping reduces crystallite size (11.71 nm) and increases microstrain (1.02%) and dislocation density (1.78 × 1016 m−2), creating abundant oxygen vacancies that serve as electron traps to enhance charge separation. From UV absorbance spectra, the energy bandgap of 1% Sr-doped SnO2 is reduced to 3.28 eV from 3.77 eV of the undoped one, reflecting successful incorporation of Sr.
The most remarkable impact of Sr doping manifests in the electronic properties, where 1% Sr incorporation produces an extraordinary three-order-of-magnitude increase in charge carrier density (5.17 × 1022 cm−3) compared to pristine SnO2 nanofibers (9.24 × 1019 cm−3). This dramatic enhancement in carrier concentration, coupled with significantly reduced charge transfer resistance (404 Ω vs. 1350 Ω for undoped) and an approximately eight-fold increase in the diffusion coefficient (8.78 × 10−10 vs. 1.13 × 10−10 cm2s−1), creates an ideal electronic environment for efficient photocatalytic reactions. The most significant finding emerges from the visible light experiments, where 1% Sr-doped SnO2 nanofibers achieve remarkable 84.74% methylene blue degradation, representing a 39% improvement over pristine SnO2 nanofibers. This substantial enhancement under visible light demonstrates that strategic Sr doping effectively transforms SnO2 from a UV-dependent to a visible light-driven photocatalyst, greatly expanding its practical utility for solar-powered environmental applications.
Beyond photocatalysis, the significantly improved electrochemical properties of the Sr-doped SnO2 nanofibers suggest their potential application in other fields such as sensing, energy storage, and photoelectrochemical water splitting. The substantial capacitance values (106 F/g for 1% Sr-doped vs. 54.4 F/g for pristine) indicate promising capabilities for supercapacitor applications. The established structure–property–function relationships provide a framework for further optimization of metal-doped semiconductor nanofibers for sustainable environmental and energy applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en18102495/s1. Figure S1: Linear relation between logarithmic value of peak current and scan rate derived from cyclic voltammetry.

Author Contributions

Conceptualization, N.K.E.; Methodology, P.B. and N.K.E.; Validation, P.B., T.T., K.K. and N.K.E.; Formal analysis, P.B., T.T., K.K. and N.K.E.; Investigation, P.B., T.T. and N.K.E.; Resources, K.K.; Data curation, P.B., T.T. and N.K.E.; Writing—original draft, P.B. and T.T.; Writing—review & editing, K.K. and N.K.E.; Supervision, K.K. and N.K.E.; Project administration, N.K.E.; Funding acquisition, N.K.E. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are available from the authors upon request.

Acknowledgments

Pranta Barua gratefully acknowledges the support of the Australian Government Research Training Program (RTP) Scholarship, which has funded his PhD research and candidature.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Khan, S.; Noor, T.; Iqbal, N.; Yaqoob, L. Photocatalytic Dye Degradation from Textile Wastewater: A Review. ACS Omega 2024, 9, 21751–21767. [Google Scholar] [CrossRef]
  2. Pathak, D.; Sharma, A.; Sharma, D.P.; Kumar, V. A review on electrospun nanofibers for photocatalysis: Upcoming technology for energy and environmental remediation applications. Appl. Surf. Sci. Adv. 2023, 18, 100471. [Google Scholar] [CrossRef]
  3. Pantò, F.; Dahrouch, Z.; Saha, A.; Patanè, S.; Santangelo, S.; Triolo, C. Photocatalytic degradation of methylene blue dye by porous zinc oxide nanofibers prepared via electrospinning: When defects become merits. Appl. Surf. Sci. 2021, 557, 149830. [Google Scholar] [CrossRef]
  4. Dhanalakshmi, M.; Saravanakumar, K.; Lakshmi Prabavathi, S.; Abinaya, M.; Muthuraj, V. Fabrication of novel surface plasmon resonance induced visible light driven iridium decorated SnO2 nanorods for degradation of organic contaminants. J. Alloys Compd. 2018, 763, 512–524. [Google Scholar] [CrossRef]
  5. Shabna, S.; Dhas, S.S.J.; Biju, C.S. Potential progress in SnO2 nanostructures for enhancing photocatalytic degradation of organic pollutants. Catal. Commun. 2023, 177, 106642. [Google Scholar] [CrossRef]
  6. Sun, C.; Yang, J.; Xu, M.; Cui, Y.; Ren, W.; Zhang, J.; Zhao, H.; Liang, B. Recent intensification strategies of SnO2-based photocatalysts: A review. Chem. Eng. J. 2022, 427, 131564. [Google Scholar] [CrossRef]
  7. Ahmed, A.; Naseem Siddique, M.; Alam, U.; Ali, T.; Tripathi, P. Improved photocatalytic activity of Sr doped SnO2 nanoparticles: A role of oxygen vacancy. Appl. Surf. Sci. 2019, 463, 976–985. [Google Scholar] [CrossRef]
  8. Uddin, M.T.; Hoque, M.E.; Chandra Bhoumick, M. Facile one-pot synthesis of heterostructure SnO2/ZnO photocatalyst for enhanced photocatalytic degradation of organic dye. RSC Adv. 2020, 10, 23554–23565. [Google Scholar] [CrossRef]
  9. Rajput, R.B.; Jamble, S.N.; Kale, R.B. A review on TiO2/SnO2 heterostructures as a photocatalyst for the degradation of dyes and organic pollutants. J. Environ. Manag. 2022, 307, 114533. [Google Scholar] [CrossRef]
  10. Zheng, Y.; Luo, C.; Liu, L.; Yang, Z.; Ren, S.; Cai, Y.; Xiong, J. Synthesis of hierarchical TiO2/SnO2 photocatalysts with different morphologies and their application for photocatalytic reduction of Cr(VI). Mater. Lett. 2016, 181, 169–172. [Google Scholar] [CrossRef]
  11. Talinungsang; Purkayastha, D.D.; Krishna, M.G. Dopant controlled photoinduced hydrophilicity and photocatalytic activity of SnO2 thin films. Appl. Surf. Sci. 2018, 447, 724–731. [Google Scholar] [CrossRef]
  12. Thangaraju, B. Structural and electrical studies on highly conducting spray deposited fluorine and antimony doped SnO2 thin films from SnCl2 precursor. Thin Solid Film. 2002, 402, 71–78. [Google Scholar] [CrossRef]
  13. Padmaja, B.; Dhanapandian, S.; Ashokkumar, K. Hydrothermally derived Mg doped tin oxide nanostructures for photocatalytic and supercapacitor applications. Mater. Sci. Eng. B 2023, 297, 116699. [Google Scholar] [CrossRef]
  14. Zulfiqar, M. Development of niobium doped tin oxide nanostructure via hydrothermal route for photocatalytic degradation of methylene blue and antimicrobial study. Ceram. Int. 2023, 49, 31718–31726. [Google Scholar] [CrossRef]
  15. Naz, F.; Nabi, G.A.K.; Nawaz, A.; Ali, S.; Siddique, M. A Novel Approach for the Photocatalytic Degradation of Binary Dyes Mixture Using SnO2 Nanoparticles as a Catalyst. J. Clust. Sci. 2023, 34, 2047–2066. [Google Scholar] [CrossRef]
  16. Ren, Y.; Wang, S.; Liu, R.; Dai, J.; Liu, X.; Yu, J. A novel route towards well-dispersed short nanofibers and nanoparticles via electrospinning. RSC Adv. 2016, 6, 30139–30147. [Google Scholar] [CrossRef]
  17. Ji, D.; Lin, Y.; Guo, X.; Ramasubramanian, B.; Wang, R.; Radacsi, N.; Jose, R.; Qin, X.; Ramakrishna, S. Electrospinning of nanofibres. Nat. Rev. Methods Primers 2024, 4, 1. [Google Scholar] [CrossRef]
  18. Khan, I.; Saeed, K.; Zekker, I.; Zhang, B.; Hendi, A.H.; Ahmad, A.; Ahmad, S.; Zada, N.; Ahmad, H.; Shah, L.A.; et al. Review on Methylene Blue: Its Properties, Uses, Toxicity and Photodegradation. Water 2022, 14, 242. [Google Scholar] [CrossRef]
  19. Hynes, N.R.J.; Kumar, J.S.; Kamyab, H.; Sujana, J.A.J.; Al-Khashman, O.A.; Kuslu, Y.; Ene, A.; Suresh Kumar, B. Modern enabling techniques and adsorbents based dye removal with sustainability concerns in textile industrial sector -A comprehensive review. J. Clean. Prod. 2020, 272, 122636. [Google Scholar] [CrossRef]
  20. Muzaffar, A.; Ahamed, M.B.; Hussain, C.M. Chapter 11—Electrolyte materials for supercapacitors. In Smart Supercapacitors; Hussain, C.M., Ahamed, M.B., Eds.; Elsevier: Amsterdam, The Netherlands, 2023; pp. 227–254. [Google Scholar] [CrossRef]
  21. Bonu, V.; Gupta, B.; Chandra, S.; Das, A.; Dhara, S.; Tyagi, A.K. Electrochemical supercapacitor performance of SnO2 quantum dots. Electrochim. Acta 2016, 203, 230–237. [Google Scholar] [CrossRef]
  22. Velmurugan, V.; Srinivasarao, U.; Ramachandran, R.; Saranya, M.; Grace, A.N. Synthesis of tin oxide/graphene (SnO2/G) nanocomposite and its electrochemical properties for supercapacitor applications. Mater. Res. Bull. 2016, 84, 145–151. [Google Scholar] [CrossRef]
  23. Han, C.; Li, W.; Liu, H.K.; Dou, S.; Wang, J. Principals and strategies for constructing a highly reversible zinc metal anode in aqueous batteries. Nano Energy 2020, 74, 104880. [Google Scholar] [CrossRef]
  24. Qian, J.; Wiener, C.G.; Zhu, Y.; Vogt, B.D. Swelling and plasticization of polymeric binders by Li-containing carbonate electrolytes using quartz crystal microbalance with dissipation. Polymer 2018, 143, 237–244. [Google Scholar] [CrossRef]
  25. Zhang, L.; Gong, H. Improvement in flexibility and volumetric performance for supercapacitor application and the effect of Ni–Fe ratio on electrode behaviour. J. Mater. Chem. A 2015, 3, 7607–7615. [Google Scholar] [CrossRef]
  26. Almora, O.; Aranda, C.; Mas-Marzá, E.; Garcia-Belmonte, G. On Mott-Schottky analysis interpretation of capacitance measurements in organometal perovskite solar cells. Appl. Phys. Lett. 2016, 109, 173903. [Google Scholar] [CrossRef]
  27. Ngamchuea, K.; Eloul, S.; Tschulik, K.; Compton, R.G. Planar diffusion to macro disc electrodes—What electrode size is required for the Cottrell and Randles-Sevcik equations to apply quantitatively? J. Solid State Electrochem. 2014, 18, 3251–3257. [Google Scholar] [CrossRef]
  28. Wang, L.; Lee, C.-Y.; Schmuki, P. Solar water splitting: Preserving the beneficial small feature size in porous α-Fe2O3 photoelectrodes during annealing. J. Mater. Chem. A 2013, 1, 212–215. [Google Scholar] [CrossRef]
  29. Hou, Y.; Qin, Z.; Han, X.; Liu, Y.; Zhang, W.; Cao, X.; Cao, Y.; Lang, J.-P.; Gu, H. Thermal evaporation-driven fabrication of Ru/RuO2 nanoparticles onto nickel foam for efficient overall water splitting. Nanoscale 2024, 16, 6662–6668. [Google Scholar] [CrossRef]
  30. Kim, H.; Kim, M.-C.; Kim, S.-b.; Kim, Y.-S.; Choi, J.-H.; Park, K.-W. Porous SnO2 nanostructure with a high specific surface area for improved electrochemical performance. RSC Adv. 2020, 10, 10519–10525. [Google Scholar] [CrossRef]
  31. Wu, B.; Zheng, H.; Wu, Y.-Q.; Huang, Z.; Thong, H.-C.; Tao, H.; Ma, J.; Zhao, C.; Xu, Z.; Liu, Y.-X.; et al. Origin of ultrahigh-performance barium titanate-based piezoelectrics: Stannum-induced intrinsic and extrinsic contributions. Nat. Commun. 2024, 15, 7700. [Google Scholar] [CrossRef]
  32. Li, J.; Luo, R.; Yan, Y. Densification kinetics and matrix microstructure of carbon fiber/carbon nanofiber/pyrocarbon composites prepared by electrophoresis and thermal gradient chemical vapor infiltration. Carbon 2011, 49, 242–248. [Google Scholar] [CrossRef]
  33. Ali Baig, A.B.; Rathinam, V.; Palaninathan, J. Photodegradation activity of yttrium-doped SnO2 nanoparticles against methylene blue dye and antibacterial effects. Appl. Water Sci. 2020, 10, 76. [Google Scholar] [CrossRef]
  34. Tammina, S.K.; Mandal, B.K.; Kadiyala, N.K. Photocatalytic degradation of methylene blue dye by nonconventional synthesized SnO2 nanoparticles. Environ. Nanotechnol. Monit. Manag. 2018, 10, 339–350. [Google Scholar] [CrossRef]
  35. Jia, T.; Sun, C.; Shi, N.; Yu, D.; Long, F.; Hu, J.; Wang, J.; Dong, B.; Li, J.; Fu, F.; et al. Efficient Oxygen Vacancy Defect Engineering for Enhancing Visible-Light Photocatalytic Performance over SnO2−x Ultrafine Nanocrystals. Nanomaterials 2022, 12, 3342. [Google Scholar] [CrossRef]
  36. Badar, N.; Yusoff, H.M.; Elong, K.; Kamarulzaman, N. Crystallite size reduction of Cr doped Al2O3 materials via optimized high-energy ball milling method. Adv. Powder Technol. 2023, 34, 104102. [Google Scholar] [CrossRef]
  37. Ferdous Anik, M.J.; Mim, S.R.; Swapno, S.S.; Munira, S.; Roy, O.; Billah, M.M. Vacancy induced enhanced photocatalytic activity of nitrogen doped CuO NPs synthesized by Co-precipitation method. Heliyon 2024, 10, e27613. [Google Scholar] [CrossRef] [PubMed]
  38. Samal, A.; Pouthika, K.; Rajesh, A.; Mohana Roopan, S.; Madhumitha, G. Photocatalytic degradation and kinetic investigations of ZnO-SnO2 heterostructures for treatment of methyl violet using non-conventional approach. Inorg. Chem. Commun. 2024, 159, 111809. [Google Scholar] [CrossRef]
  39. Sutapa, I.W.; Wahid Wahab, A.; Taba, P.; Nafie, N.L. Dislocation, crystallite size distribution and lattice strain of magnesium oxide nanoparticles. J. Phys. Conf. Ser. 2018, 979, 012021. [Google Scholar] [CrossRef]
  40. Yon, V.; Rochat, N.; Charles, M.; Nolot, E.; Gergaud, P. X-Ray Diffraction Microstrain Analysis for Extraction of Threading Dislocation Density of GaN Films Grown on Silicon, Sapphire, and SiC Substrates. Phys. Status Solidi (b) 2020, 257, 1900579. [Google Scholar] [CrossRef]
  41. Nguyen, T.D.; Roh, S.; Nguyen, M.T.N.; Lee, J.S. Structural Control of Nanofibers According to Electrospinning Process Conditions and Their Applications. Micromachines 2023, 14, 2022. [Google Scholar] [CrossRef]
  42. Mezyen, M.; El Fidha, G.; Bitri, N.; Harrathi, F.; Ly, I.; Llobet, E. Visible light activated SnO2:Dy thin films for the photocatalytic degradation of methylene blue. RSC Adv. 2023, 13, 31151–31166. [Google Scholar] [CrossRef]
  43. Borisenko, E.B.; Gnesin, B.A. Low-temperature static recrystallization in strontium doped potassium chloride. Scr. Mater. 2001, 44, 923–927. [Google Scholar] [CrossRef]
  44. Agrawal, H.; Patra, B.K.; Altantzis, T.; De Backer, A.; Garnett, E.C. Quantifying Strain and Dislocation Density at Nanocube Interfaces after Assembly and Epitaxy. ACS Appl. Mater. Interfaces 2020, 12, 8788–8794. [Google Scholar] [CrossRef]
  45. Wang, C.; Liu, H.; Wang, G.; Ye, F.; Tao, S.; Li, X.; Qiu, P.; Huang, Z. Ni (II) doping induced lattice distortion in Zn3In2S6/BiOBr-OVs for boosting photocatalytic removal of antibiotics and Cr (VI) performance. Sep. Purif. Technol. 2023, 324, 124457. [Google Scholar] [CrossRef]
  46. Udayabhaskar, R.; Karthikeyan, B. Role of micro-strain and defects on band-gap, fluorescence in near white light emitting Sr doped ZnO nanorods. J. Appl. Phys. 2014, 116, 094310. [Google Scholar] [CrossRef]
  47. Yu, Y.; He, T.; Guo, L.; Yang, Y.; Guo, L.; Tang, Y.; Cao, Y. Efficient visible-light photocatalytic degradation system assisted by conventional Pd catalysis. Sci. Rep. 2015, 5, 9561. [Google Scholar] [CrossRef] [PubMed]
  48. Lal, M.; Sharma, P.; Ram, C. Synthesis and photocatalytic potential of Nd-doped TiO2 under UV and solar light irradiation using a sol-gel ultrasonication method. Results Mater. 2022, 15, 100308. [Google Scholar] [CrossRef]
  49. Staaf, L.G.H.; Lundgren, P.; Enoksson, P. Present and future supercapacitor carbon electrode materials for improved energy storage used in intelligent wireless sensor systems. Nano Energy 2014, 9, 128–141. [Google Scholar] [CrossRef]
  50. Morka, T.D.; Ujihara, M. Enhanced Performance of WO3/SnO2 Nanocomposite Electrodes with Redox-Active Electrolytes for Supercapacitors. Int. J. Mol. Sci. 2023, 24, 6045. [Google Scholar] [CrossRef]
  51. Liu, J.; Wang, J.; Xu, C.; Jiang, H.; Li, C.; Zhang, L.; Lin, J.; Shen, Z.X. Advanced Energy Storage Devices: Basic Principles, Analytical Methods, and Rational Materials Design. Adv. Sci. 2018, 5, 1700322. [Google Scholar] [CrossRef]
  52. Gogotsi, Y.; Penner, R.M. Energy Storage in Nanomaterials—Capacitive, Pseudocapacitive, or Battery-like? ACS Nano 2018, 12, 2081–2083. [Google Scholar] [CrossRef] [PubMed]
  53. Surendra, B.S.; Veerabhdraswamy, M.; Anantharaju, K.S.; Nagaswarupa, H.P.; Prashantha, S.C. Green and chemical-engineered CuFe2O4: Characterization, cyclic voltammetry, photocatalytic and photoluminescent investigation for multifunctional applications. J. Nanostruct. Chem. 2018, 8, 45–59. [Google Scholar] [CrossRef]
  54. Sajid, M.M.; Khan, S.B.; Shad, N.A.; Amin, N.; Zhang, Z. Visible light assisted photocatalytic degradation of crystal violet dye and electrochemical detection of ascorbic acid using a BiVO4/FeVO4 heterojunction composite. RSC Adv. 2018, 8, 23489–23498. [Google Scholar] [CrossRef]
  55. Mainka, J.; Gao, W.; He, N.; Dillet, J.; Lottin, O. A General Equivalent Electrical Circuit Model for the characterization of MXene/graphene oxide hybrid-fiber supercapacitors by electrochemical impedance spectroscopy—Impact of fiber length. Electrochim. Acta 2022, 404, 139740. [Google Scholar] [CrossRef]
  56. De Pasquale, L.; Tavella, F.; Longo, V.; Favaro, M.; Perathoner, S.; Centi, G.; Ampelli, C.; Genovese, C. The Role of Substrate Surface Geometry in the Photo-Electrochemical Behaviour of Supported TiO2 Nanotube Arrays: A Study Using Electrochemical Impedance Spectroscopy (EIS). Molecules 2023, 28, 3378. [Google Scholar] [CrossRef] [PubMed]
  57. Yadav, S.; Jindal, J.; Mittal, A.; Sharma, S.; Kumari, K.; Kumar, N. Facile solution combustion synthesized, Li doped ZnO nanostructures for removal of abiotic contaminants. J. Phys. Chem. Solids 2021, 157, 110217. [Google Scholar] [CrossRef]
  58. Lee, J.-S.M.; Briggs, M.E.; Hu, C.-C.; Cooper, A.I. Controlling electric double-layer capacitance and pseudocapacitance in heteroatom-doped carbons derived from hypercrosslinked microporous polymers. Nano Energy 2018, 46, 277–289. [Google Scholar] [CrossRef]
  59. Huang, P.-R.; He, Y.; Cao, C.; Lu, Z.-H. Impact of lattice distortion and electron doping on α-MoO3 electronic structure. Sci. Rep. 2014, 4, 7131. [Google Scholar] [CrossRef]
  60. Sivula, K. Mott–Schottky Analysis of Photoelectrodes: Sanity Checks Are Needed. ACS Energy Lett. 2021, 6, 2549–2551. [Google Scholar] [CrossRef]
  61. Hankin, A.; Bedoya-Lora, F.E.; Alexander, J.C.; Regoutz, A.; Kelsall, G.H. Flat band potential determination: Avoiding the pitfalls. J. Mater. Chem. A 2019, 7, 26162–26176. [Google Scholar] [CrossRef]
  62. Gadisa, B.T.; Kassahun, S.K.; Appiah-Ntiamoah, R.; Kim, H. Tuning the charge carrier density and exciton pair separation in electrospun 1D ZnO-C composite nanofibers and its effect on photodegradation of emerging contaminants. J. Colloid Interface Sci. 2020, 570, 251–263. [Google Scholar] [CrossRef]
  63. Wu, H.; Wang, J.; Chen, R.; Yuan, C.; Zhang, J.; Zhang, Y.; Sheng, J.; Dong, F. Zn-doping mediated formation of oxygen vacancies in SnO2 with unique electronic structure for efficient and stable photocatalytic toluene degradation. Chin. J. Catal. 2021, 42, 1195–1204. [Google Scholar] [CrossRef]
  64. Klahr, B.; Gimenez, S.; Fabregat-Santiago, F.; Hamann, T.; Bisquert, J. Water Oxidation at Hematite Photoelectrodes: The Role of Surface States. J. Am. Chem. Soc. 2012, 134, 4294–4302. [Google Scholar] [CrossRef]
  65. Jacobsson, T.J.; Edvinsson, T. Photoelectrochemical Determination of the Absolute Band Edge Positions as a Function of Particle Size for ZnO Quantum Dots. J. Phys. Chem. C 2012, 116, 15692–15701. [Google Scholar] [CrossRef]
  66. Lasia, A. Semiconductors and Mott-Schottky Plots. In Electrochemical Impedance Spectroscopy and Its Applications; Lasia, A., Ed.; Springer: New York, NY, USA, 2014; pp. 251–255. [Google Scholar] [CrossRef]
  67. Khlyustova, A.; Sirotkin, N.; Kusova, T.; Kraev, A.; Titov, V.; Agafonov, A. Doped TiO2: The effect of doping elements on photocatalytic activity. Mater. Adv. 2020, 1, 1193–1201. [Google Scholar] [CrossRef]
  68. Liu, J.; Li, N.; Dong, Q.; Li, J.; Qin, C.; Wang, L. Tailoring electrical property of the low-temperature processed SnO2 for high-performance perovskite solar cells. Sci. China Mater. 2019, 62, 173–180. [Google Scholar] [CrossRef]
  69. van der Heijden, O.; Park, S.; Vos, R.E.; Eggebeen, J.J.J.; Koper, M.T.M. Tafel Slope Plot as a Tool to Analyze Electrocatalytic Reactions. ACS Energy Lett. 2024, 9, 1871–1879. [Google Scholar] [CrossRef]
  70. Chen, X.; Luo, X.; Liu, L.; Ping, J.; Sun, S. SnS2 Quantum Dots Decorated MoS2 Nanosheets Enabling Efficient Photocatalytic H2 Evolution in CO2 Saturated Water. Photocatal. Res. Potential 2024, 1, 10003. [Google Scholar] [CrossRef]
  71. Paul, P.C.; Shah, A.; Singh, L.R.; Mahato, M.; Mahato, D.K. Mn-doped SrTiO3 and SrTiO3–Fe2O3 composite perovskites: Photocatalytic dye degradation and energy storage application. J. Mater. Sci. Mater. Electron. 2023, 34, 1305. [Google Scholar] [CrossRef]
  72. Levi, M.D.; Markevich, E.; Aurbach, D. Comparison between Cottrell diffusion and moving boundary models for determination of the chemical diffusion coefficients in ion-insertion electrodes. Electrochim. Acta 2005, 51, 98–110. [Google Scholar] [CrossRef]
  73. Salvador, P. On the Nature of Photogenerated Radical Species Active in the Oxidative Degradation of Dissolved Pollutants with TiO2 Aqueous Suspensions:  A Revision in the Light of the Electronic Structure of Adsorbed Water. J. Phys. Chem. C 2007, 111, 17038–17043. [Google Scholar] [CrossRef]
  74. Mohamadpour, F.; Amani, A.M. Photocatalytic systems: Reactions, mechanism, and applications. RSC Adv. 2024, 14, 20609–20645. [Google Scholar] [CrossRef]
  75. Suhaimi, N.A.A.; Roslan, N.N.; Amirul, N.B.; Lau, H.L.H.; Hasman, A.A.; Nur, M.; Lim, J.W.; Usman, A. Unraveling the photocatalytic degradation kinetics and efficiency of methylene blue, rhodamine B, and auramine O in their ternary mixture: Diffusion and conformational insights. React. Kinet. Mech. Catal. 2024, 137, 3441–3462. [Google Scholar] [CrossRef]
  76. Tauc, J.; Menth, A. States in the gap. J. Non-Cryst. Solids 1972, 8–10, 569–585. [Google Scholar] [CrossRef]
  77. Liu, D.; Wu, G.; Luo, S.; Wang, G.; Wang, X.; Wang, X. Unlocking n-type semiconductivity in diamond: A breakthrough approach via surface metal doping. APL Mater. 2024, 12, 121118. [Google Scholar] [CrossRef]
  78. Latour, R.A. The langmuir isotherm: A commonly applied but misleading approach for the analysis of protein adsorption behavior. J. Biomed. Mater. Res. Part A 2015, 103, 949–958. [Google Scholar] [CrossRef]
  79. Hameed, B.H.; Din, A.T.M.; Ahmad, A.L. Adsorption of methylene blue onto bamboo-based activated carbon: Kinetics and equilibrium studies. J. Hazard. Mater. 2007, 141, 819–825. [Google Scholar] [CrossRef]
  80. Mussa, Z.H.; Al-Ameer, L.R.; Al-Qaim, F.F.; Deyab, I.F.; Kamyab, H.; Chelliapan, S. A comprehensive review on adsorption of methylene blue dye using leaf waste as a bio-sorbent: Isotherm adsorption, kinetics, and thermodynamics studies. Environ. Monit. Assess. 2023, 195, 940. [Google Scholar] [CrossRef]
  81. Jara, Y.S.; Mekiso, T.T.; Washe, A.P. Highly efficient catalytic degradation of organic dyes using iron nanoparticles synthesized with Vernonia Amygdalina leaf extract. Sci. Rep. 2024, 14, 6997. [Google Scholar] [CrossRef]
  82. Wei, Y.; Fu, Z.; Meng, Y.; Li, C.; Yin, F.; Wang, X.; Zhang, C.; Guo, L.; Sun, S. Photocatalytic degradation of methylene blue over MIL-100(Fe)/GO composites: A performance and kinetic study. Int. J. Coal Sci. Technol. 2024, 11, 42. [Google Scholar] [CrossRef]
  83. Wang, H.; Sun, F.; Zhang, Y.; Li, L.; Chen, H.; Wu, Q.; Yu, J.C. Photochemical growth of nanoporous SnO2 at the air–water interface and its high photocatalytic activity. J. Mater. Chem. 2010, 20, 5641–5645. [Google Scholar] [CrossRef]
  84. Batzill, M. Surface Science Studies of Gas Sensing Materials: SnO2. Sensors 2006, 6, 1345–1366. [Google Scholar] [CrossRef]
  85. Chen, S.; Qi, Y.; Li, C.; Domen, K.; Zhang, F. Surface Strategies for Particulate Photocatalysts toward Artificial Photosynthesis. Joule 2018, 2, 2260–2288. [Google Scholar] [CrossRef]
  86. Sachs, M.; Pastor, E.; Kafizas, A.; Durrant, J.R. Evaluation of Surface State Mediated Charge Recombination in Anatase and Rutile TiO2. J. Phys. Chem. Lett. 2016, 7, 3742–3746. [Google Scholar] [CrossRef]
  87. Bhawna; Sharma, R.; Kumar, S.; Kumar, R.; Sahu, P.K.; Kumari, V.; Mishra, A.K.; Kumar, V. Unlocking the Potential of N-Doped SnO2 for Sustainable Photocatalytic Degradation of Carcinogenic Dyes. Separations 2023, 10, 322. [Google Scholar] [CrossRef]
  88. Al-Hamdi, A.M.; Rinner, U.; Sillanpää, M. Tin dioxide as a photocatalyst for water treatment: A review. Process Saf. Environ. Prot. 2017, 107, 190–205. [Google Scholar] [CrossRef]
  89. Reddy, C.V.; Babu, B.; Shim, J. Synthesis of Cr-doped SnO2 quantum dots and its enhanced photocatalytic activity. Mater. Sci. Eng. B 2017, 223, 131–142. [Google Scholar] [CrossRef]
  90. Akkera, H.S.; Mann, V.; Varalakshmi, B.N.; Ploloju, M.; Kambhala, N.; Venkatesh, G. Effect of Sr-doped on physical and photoluminescence properties of SnO2 transparent conducting oxide thin films. J. Mater. Sci. Mater. Electron. 2023, 34, 1044. [Google Scholar] [CrossRef]
  91. Bouras, K.; Rehspringer, J.L.; Schmerber, G.; Rinnert, H.; Colis, S.; Ferblantier, G.; Balestrieri, M.; Ihiawakrim, D.; Dinia, A.; Slaoui, A. Optical and structural properties of Nd doped SnO2 powder fabricated by the sol–gel method. J. Mater. Chem. C 2014, 2, 8235–8243. [Google Scholar] [CrossRef]
  92. Entradas, T.; Cabrita, J.F.; Dalui, S.; Nunes, M.R.; Monteiro, O.C.; Silvestre, A.J. Synthesis of sub-5 nm Co-doped SnO2 nanoparticles and their structural, microstructural, optical and photocatalytic properties. Mater. Chem. Phys. 2014, 147, 563–571. [Google Scholar] [CrossRef]
  93. Khaenamkaew, P.; Manop, D.; Tanghengjaroen, C.; Palakawong Na Ayuthaya, W. Crystal Structure, Lattice Strain, Morphology, and Electrical Properties of SnO2 Nanoparticles Induced by Low Calcination Temperature. Adv. Mater. Sci. Eng. 2020, 2020, 3852421. [Google Scholar] [CrossRef]
  94. Zhang, G.; Xie, C.; Zhang, S.; Zhang, S.; Xiong, Y. Defect Chemistry of the Metal Cation Defects in the p- and n-Doped SnO2 Nanocrystalline Films. J. Phys. Chem. C 2014, 118, 18097–18109. [Google Scholar] [CrossRef]
  95. Zulfiqar; Yuan, Y.; Yang, J.; Wang, W.; Ye, Z.; Lu, J. Structural and optical properties of (Zn, Co) co-doped SnO2 nano particles. J. Mater. Sci. Mater. Electron. 2016, 27, 12119–12127. [Google Scholar] [CrossRef]
  96. Prakash, K.; Senthil Kumar, P.; Pandiaraj, S.; Saravanakumar, K.; Karuthapandian, S. Controllable synthesis of SnO2 photocatalyst with superior photocatalytic activity for the degradation of methylene blue dye solution. J. Exp. Nanosci. 2016, 11, 1138–1155. [Google Scholar] [CrossRef]
  97. Shao, M.; Liu, J.; Ding, W.; Wang, J.; Dong, F.; Zhang, J. Oxygen vacancy engineering of self-doped SnO2−x nanocrystals for ultrasensitive NO2 detection. J. Mater. Chem. C 2020, 8, 487–494. [Google Scholar] [CrossRef]
  98. Zhou, Z.H.; Min, Y.M.; Liu, X.X.; Ding, J.Q.; Guo, J.H.; Hu, F.R.; Liu, L.Z. Regulation of oxygen vacancy types on SnO2 (110) surface by external strain. AIP Adv. 2016, 6, 055102. [Google Scholar] [CrossRef]
  99. Dai, M.; Niu, Q.; Wu, S.; Lin, Y.; Biswas, J.K.; Yang, C. Hydroxyl radicals in ozone-based advanced oxidation of organic contaminants: A review. Environ. Chem. Lett. 2024, 22, 3059–3106. [Google Scholar] [CrossRef]
Figure 1. SEM images of SnO2 nanofibers with (a) commercial SnO2 NP with a scale size of 8 μm. (b) Pristine (c) 1%, (d), 3%, and (e) 5% of Sr doping with a scale size of 3 μm.
Figure 1. SEM images of SnO2 nanofibers with (a) commercial SnO2 NP with a scale size of 8 μm. (b) Pristine (c) 1%, (d), 3%, and (e) 5% of Sr doping with a scale size of 3 μm.
Energies 18 02495 g001
Figure 2. EDS mapping of 1% Sr-doped SnO2 NF (a) Table for atomic and weight concentration of Oxygen, Strontium and Tin (b) a graph representing element signature based on peak conditions considering relative amounts (c) Analyzed area of SEM image (pink color) with mapping of Sr, O and Sn with different colors.
Figure 2. EDS mapping of 1% Sr-doped SnO2 NF (a) Table for atomic and weight concentration of Oxygen, Strontium and Tin (b) a graph representing element signature based on peak conditions considering relative amounts (c) Analyzed area of SEM image (pink color) with mapping of Sr, O and Sn with different colors.
Energies 18 02495 g002
Figure 3. XRD patterns of (a) commercial SnO2 and (b) pristine (c) 1% Sr-doped, (d) 3% Sr-doped, (e) and 5% Sr-doped SnO2 NFs. (f) Additional phases (Sr3Sn5) of 1% Sr-doped SnO2 NF derived from Highscore XRD analysis software (v 5.2) after Rietveld refinement.
Figure 3. XRD patterns of (a) commercial SnO2 and (b) pristine (c) 1% Sr-doped, (d) 3% Sr-doped, (e) and 5% Sr-doped SnO2 NFs. (f) Additional phases (Sr3Sn5) of 1% Sr-doped SnO2 NF derived from Highscore XRD analysis software (v 5.2) after Rietveld refinement.
Energies 18 02495 g003
Figure 4. CV plots for commercial SnO2 NP and SnO2 nanofibers with 0, 1, 3, 5 wt.% of Sr doping.
Figure 4. CV plots for commercial SnO2 NP and SnO2 nanofibers with 0, 1, 3, 5 wt.% of Sr doping.
Energies 18 02495 g004
Figure 5. (a) Nyquist plot of EIS spectra of commercial SnO2 NP, 0%, 1%, 3%, and 5% Sr-doped SnO2 nanofiber. (b) Zoomed-in high-frequency region with equivalent circuit model (inset).
Figure 5. (a) Nyquist plot of EIS spectra of commercial SnO2 NP, 0%, 1%, 3%, and 5% Sr-doped SnO2 nanofiber. (b) Zoomed-in high-frequency region with equivalent circuit model (inset).
Energies 18 02495 g005
Figure 6. Mott–Schottky graph for (a) commercial SnO2 NP and SnO2 NFs with (b) pristine (c) 1% (d) 3% and (e) 5% of Sr doping. The dashed line intersecting X-axis indicates the extrapolation of the linear region of the M-S plot to determine the flat-band potential. The red circle indicates the density of surface states.
Figure 6. Mott–Schottky graph for (a) commercial SnO2 NP and SnO2 NFs with (b) pristine (c) 1% (d) 3% and (e) 5% of Sr doping. The dashed line intersecting X-axis indicates the extrapolation of the linear region of the M-S plot to determine the flat-band potential. The red circle indicates the density of surface states.
Energies 18 02495 g006
Figure 7. (a) Linear sweep voltammetry and (b) Tafel plot of commercial SnO2 NP and with different concentrations of Sr-doped SnO2 nanofibers.
Figure 7. (a) Linear sweep voltammetry and (b) Tafel plot of commercial SnO2 NP and with different concentrations of Sr-doped SnO2 nanofibers.
Energies 18 02495 g007
Figure 8. (a) Chronoamperometry measurement and (b) Cottrell plot of commercial SnO2 NP and Sr-doped (0, 1, 3, 5%) SnO2 nanofiber.
Figure 8. (a) Chronoamperometry measurement and (b) Cottrell plot of commercial SnO2 NP and Sr-doped (0, 1, 3, 5%) SnO2 nanofiber.
Energies 18 02495 g008
Figure 9. Tauc plot of pristine and 1% Sr-doped SnO2 NF derived from UV−Vis absorbance spectra (inset).
Figure 9. Tauc plot of pristine and 1% Sr-doped SnO2 NF derived from UV−Vis absorbance spectra (inset).
Energies 18 02495 g009
Figure 10. UV light driven-photocatalytic degradation: UV-Vis absorption of MB dye with (a) commercial SnO2 (b) 0, (c) 1, (d) 3, and (e) 5% of Sr-doped SnO2 NFs as photocatalyst at λ m a x + 664   n m .
Figure 10. UV light driven-photocatalytic degradation: UV-Vis absorption of MB dye with (a) commercial SnO2 (b) 0, (c) 1, (d) 3, and (e) 5% of Sr-doped SnO2 NFs as photocatalyst at λ m a x + 664   n m .
Energies 18 02495 g010
Figure 11. (a) Relative degradation with concentration and (b) degradation efficiency until 180 min of MB with Sr-doped and undoped SnO2 NF photocatalyst.
Figure 11. (a) Relative degradation with concentration and (b) degradation efficiency until 180 min of MB with Sr-doped and undoped SnO2 NF photocatalyst.
Energies 18 02495 g011
Figure 12. Degradation curves of ln (C0/Ct) vs. irradiation time of MB dye degradation with the presence of different samples of SnO2 NF. The red colored lines indicate the linear fitting of a pseudo-first-order kinetic model.
Figure 12. Degradation curves of ln (C0/Ct) vs. irradiation time of MB dye degradation with the presence of different samples of SnO2 NF. The red colored lines indicate the linear fitting of a pseudo-first-order kinetic model.
Energies 18 02495 g012
Figure 13. Solar simulator lamp-driven photocatalytic degradation: UV-Vis absorption spectra of MB solution with (a) commercial SnO2, (b) pristine SnO2 NF, and (c) 1% Sr-doped SnO2 NF as photocatalysts with different time intervals for 180 min.
Figure 13. Solar simulator lamp-driven photocatalytic degradation: UV-Vis absorption spectra of MB solution with (a) commercial SnO2, (b) pristine SnO2 NF, and (c) 1% Sr-doped SnO2 NF as photocatalysts with different time intervals for 180 min.
Energies 18 02495 g013
Figure 14. Schematic illustration of the oxidation process removing organic pollutants from wastewater [88] (Reprinted with permission).
Figure 14. Schematic illustration of the oxidation process removing organic pollutants from wastewater [88] (Reprinted with permission).
Energies 18 02495 g014
Figure 15. Photocatalytic reaction of SnO2 with MB dye degradation.
Figure 15. Photocatalytic reaction of SnO2 with MB dye degradation.
Energies 18 02495 g015
Table 1. The average diameter of Sr-doped SnO2 nanofibers.
Table 1. The average diameter of Sr-doped SnO2 nanofibers.
%Sr-Doped SnO2 NFAverage Diameter (nm)
Pristine 259.04   ±   25.57
1 288.69   ±   17.24
3 209.49   ±   15.29
5 148.72   ±   5.98
Table 2. Crystallographic data derived from XRD of manufactured SnO2, undoped, and doped SnO2 NFs.
Table 2. Crystallographic data derived from XRD of manufactured SnO2, undoped, and doped SnO2 NFs.
Sample NameAverage Crystallite Size (nm)Dislocation Density (m−2)Micro Strain (%)
Manufactured SnO247.231.9   ×   10 15 0.23
0%-Sr doped SnO2 NF12.711.37   ×   10 16 0.77
1%-Sr doped SnO2 NF11.711.78   ×   10 16 1.02
3%-Sr doped SnO2 NF28.083.36   ×   10 15 0.41
5%-Sr doped SnO2 NF7.143.53   ×   10 16 1.42
Table 3. Lattice parameters of undoped and doped SnO2 derived from XRD peaks (110) and (101).
Table 3. Lattice parameters of undoped and doped SnO2 derived from XRD peaks (110) and (101).
Sample Namea = b (Å)c (Å)Cell Volume (Å3)
Commercial SnO24.723.1670.4
Pristine SnO2 NF4.703.2171.1
1% Sr-doped SnO2 NF4.823.2174.76
3% Sr-doped SnO2 NF4.753.1972.06
5% Sr-doped SnO2 NF4.723.0167.06
Table 4. Total charge and specific capacitance calculated from CV.
Table 4. Total charge and specific capacitance calculated from CV.
% Sr Doping of SnO2 NFsTotal Charge, Q (C)Specific Capacitance (F/g)
Commercial SnO2 NP
Pristine SnO2 NF
1.80 ×   10 −5
4.35 ×   10 −5
22.5
54.4
1% Sr-doped8.45 ×   10 −5106
3% Sr-doped6.80 ×   10 −585
5% Sr-doped3.76 ×   10 −547.1
Table 5. Equivalent circuit parameters fitted with EIS plot for SnO2 NFs with Sr doping.
Table 5. Equivalent circuit parameters fitted with EIS plot for SnO2 NFs with Sr doping.
Equivalent Circuit ParametersCommercial SnO2 NPPristine SnO2 NF1 wt.% Sr-Doped3 wt.% Sr-Doped5 wt.% Sr-Doped
Rs (Ohm-cm2)3.83.732.871.733.58
CPEdl (F)7.06   ×   10 5 1.42   ×   10 4 3.8   ×   10 4 1.4   ×   10 4 6.97   ×   10 5
η dl0.940.970.840.890.93
Rct (Ohm-cm2)85913504041000450
ZW (Ohm-cm2)3   ×   10 4 1.06   ×   10 4 3.84   ×   10 3 1.13   ×   10 4 5.71   ×   10 3
RGC (Ohm-cm2)200200100100200
CPEGC (F)7.5   ×   10 4 7.33   ×   10 4 3.5   ×   10 4 6.4   ×   10 4 2.5   ×   10 3
η GC0.720.680.80.860.45
Table 6. Parameters obtained from the slope of MS graphs for Sr-doped SnO2 nanofibers.
Table 6. Parameters obtained from the slope of MS graphs for Sr-doped SnO2 nanofibers.
% of Sr DopingFlat Band Potential, VfbSlope Charge   Carrier   Density ,   N D   ( c m −3)
Commercial SnO2 NP−0.48 V 6.5   ×   10 8 1.2   ×   10 22
Pristine SnO2 NF−0.45 V8.4   ×   10 10 9.24   ×   10 19
1% Sr doping−0.53 V 1.5   ×   10 8 5.17   ×   10 22
3% Sr doping−0.68 V2.4   ×   10 10 3.23   ×   10 20
5% Sr doping−0.62 V3.09   ×   10 11 2.51   ×   10 19
Table 7. Tafel slope derived from Tafel plot of Sr-doped SnO2 NFs and commercial SnO2.
Table 7. Tafel slope derived from Tafel plot of Sr-doped SnO2 NFs and commercial SnO2.
SampleTafel Slope (V/dec)
Commercial SnO2 NP2.02
Pristine SnO2 NF0.84
1% Sr-doped0.47
3% Sr-doped0.62
5% Sr-doped0.49
Table 8. Diffusion coefficient calculated from Cottrell equation for undoped and doped SnO2 nanofibers.
Table 8. Diffusion coefficient calculated from Cottrell equation for undoped and doped SnO2 nanofibers.
% of Sr DopingSlope i/(t)0.5D, Diffusion Coefficient (cm2s−1)
Commercial SnO2 NP 6.17   ×   10 6 3.62   ×   10 11
Pristine SnO2 NF1.09   ×   10 5 1.13   ×   10 10
1% Sr-doped3.04   ×   10 5 8.78   ×   10 10
3% Sr-doped1.39   ×   10 5 1.84   ×   10 10
5% Sr-doped1.04   ×   10 5 1.03   ×   10 10
Table 9. UV light-driven photocatalytic degradation efficiency of MB with SnO2 NF photocatalyst.
Table 9. UV light-driven photocatalytic degradation efficiency of MB with SnO2 NF photocatalyst.
Sample% Degradation
Commercial SnO2 NP16
0% Sr-doped SnO2 NF61
1% Sr-doped SnO2 NF69
3% Sr-doped SnO2 NF58
5% Sr-doped SnO2 NF63
Table 10. Reaction rate and R2 value derived from the pseudo-first-order model for undoped and doped SnO2 NFs.
Table 10. Reaction rate and R2 value derived from the pseudo-first-order model for undoped and doped SnO2 NFs.
SampleReaction Rate, k (min−1), from 40 to 120 minR2
Pristine SnO2 NF6.32 × 10−40.83
1% Sr-doped0.001250.87
3% Sr-doped8.76 × 10−40.98
5% Sr-doped5.65 × 10−40.99
Table 11. Visible light-driven photocatalytic degradation efficiency of MB with SnO2 NF.
Table 11. Visible light-driven photocatalytic degradation efficiency of MB with SnO2 NF.
Sample NameCommercial SnO20% Sr-Doped SnO2 NF1% Sr-Doped SnO2 NF
% degradation70%61%84.74%
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

Barua, P.; Thai, T.; Krishnan, K.; Elumalai, N.K. Strontium-Doped Tin Oxide Nanofibers for Enhanced Visible Light Photocatalysis. Energies 2025, 18, 2495. https://doi.org/10.3390/en18102495

AMA Style

Barua P, Thai T, Krishnan K, Elumalai NK. Strontium-Doped Tin Oxide Nanofibers for Enhanced Visible Light Photocatalysis. Energies. 2025; 18(10):2495. https://doi.org/10.3390/en18102495

Chicago/Turabian Style

Barua, Pranta, Tan Thai, Kannoorpatti Krishnan, and Naveen Kumar Elumalai. 2025. "Strontium-Doped Tin Oxide Nanofibers for Enhanced Visible Light Photocatalysis" Energies 18, no. 10: 2495. https://doi.org/10.3390/en18102495

APA Style

Barua, P., Thai, T., Krishnan, K., & Elumalai, N. K. (2025). Strontium-Doped Tin Oxide Nanofibers for Enhanced Visible Light Photocatalysis. Energies, 18(10), 2495. https://doi.org/10.3390/en18102495

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

Article Metrics

Back to TopTop