Next Article in Journal
Artificial Vision Systems for Mobility Impairment Detection: Integrating Synthetic Data, Ethical Considerations, and Real-World Applications
Next Article in Special Issue
Comparing Strategies for Optimal Pumps as Turbines Selection in Pressurised Irrigation Networks Using Particle Swarm Optimisation: Application in Canal del Zújar Irrigation District, Spain
Previous Article in Journal
NCC—An Efficient Deep Learning Architecture for Non-Coding RNA Classification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Sm-Doped CuO–SnO2:FSprayed Thin Film: An Eco-Friendly Dual-Function Solution for the Buffer Layer and an Effective Photocatalyst for Ampicillin Degradation

1
LR99ES13 Laboratoire de Physique de La Matière Condensée (LPMC), Département de Physique, Faculté des Sciences de Tunis, Université Tunis El Manar, Tunis 2092, Tunisia
2
Department of Electrical Engineering, College of Engineering and Computers, Al-Lith, Umm Al-Qura University, Makkah 21955, Saudi Arabia
3
Nanotechnology and Advanced Materials Program, Kuwait Institute for Scientific Research, Safat 13109, Kuwait
4
ITODYS, Paris Université, CNRS UMR-7086, 75205 Paris, France
5
Faculty of Materials Science and Engineering, National University of Science and Technology POLITEHNICA Bucharest, 060042 Bucharest, Romania
6
American Romanian Academy of Arts and Sciences, Citrus Heights, CA 95616, USA
*
Author to whom correspondence should be addressed.
Technologies 2025, 13(5), 197; https://doi.org/10.3390/technologies13050197
Submission received: 21 March 2025 / Revised: 23 April 2025 / Accepted: 25 April 2025 / Published: 13 May 2025
(This article belongs to the Special Issue Sustainable Water and Environmental Technologies of Global Relevance)

Abstract

Synthesis and characterization of undoped and samarium-doped CuO–SnO2:F thin films using the spray pyrolysis technique are presented. The effect of the samarium doping level on the physical properties of these films was thoroughly analyzed. X-ray diffraction patterns proved the successful synthesis of pure CuO–SnO2:F thin films, free from detectable impurities. The smallest crystallite size was observed in 6% Sm-doped CuO–SnO2:F thin films. The 6% Sm-doped CuO–SnO2films demonstrated an increasedsurface area of 40.6 m2/g, highlighting improved textural properties, which was further validated by XPS analysis.The bandgap energy was found to increase from 1.90 eV for undoped CuO–SnO2:F to 2.52 eV for 4% Sm-doped CuO–SnO2:F, before decreasing to 2.03 eV for 6% Sm-doped CuO–SnO2:F thin films. Photoluminescence spectra revealed various emission peaks, suggesting a quenching effect. A numerical simulation of a new solar cell based on FTO/ZnO/Sm–CuO–SnO2:F/X/Mo was carried out using Silvaco Atlas software, where X represented the absorber layer CIGS, CdTe, and CZTS. The results showed that the solar cell with CIGS as the absorber layer achieved the highest efficiency of 15.98. Additionally, the thin films demonstrated strong photocatalytic performance, with 6% Sm-doped CuO–SnO2:F showing 86% degradation of ampicillin after two hours. This comprehensive investigation provided valuable insights into the synthesis, properties, and potential applications of Sm-doped CuO–SnO2 thin films, particularly for solar energy and pharmaceutical applications.

1. Introduction

Global challenges, such as the depletion of fossil fuels and the increasing levels of environmental pollution, have intensified the demand for renewable, clean, and sustainable energy sources. Among these, solar energy stands out as a promising alternative.
The technological evolution of solar cells can be categorized into three main development stages: silicon, thin film, and organic solar cells [1,2,3]. Currently, thin film solar cells have achieved commercial maturity, with global production exceeding 30 GW (gigawatts). However, researchers face the ongoing challenge of producing solar cells that are both cost-effective and environmentally friendly, while also achieving high efficiency [4].
Among the various solar cell structures, those utilizing cadmium sulfide (CdS) as the buffer layer have garnered considerable attention due to their long-term reliability and consumer confidence [5,6]. Yet, research in this area has plateaued, primarily due to the toxic nature and environmental concerns associated with CdS [7,8]. As a result, attention is now focused on using cost-effective, non-toxic thin film materials. Materials such as In2S3 and CuO–SnO2 thin films have emerged as promising candidates, owing to their suitable physical properties and low environmental impact [9,10,11]. In this study, we explore the optimization of CuO–SnO2 thin films through the introduction of samarium, aiming to enhance device efficiency.
Our choice of samarium as adoping material was based on previous research conducted in our laboratory on titanium dioxide (TiO2) thin film by Naffouti et al. [12]. The results showed that samarium doping demonstrated significant improvements in the physical properties of TiO2, which led to a better performance in a variety of optoelectronic applications [12].
In addition to energy applications, photocatalysis is another process that uses light energy to purify and disinfect water by degrading organic pollutants such as antibiotics. While antibiotics are essential in modern medicine for treating bacterial infections and saving lives [13,14,15], their widespread use in both human and veterinary medicine has resulted in their unintended presence in water sources, creating significant risks to both human health and the environment [16]. Ampicillin, a commonly used broad-spectrum antibiotic, is frequently found in wastewater due to its extensive application. The presence of antibiotics in water systems creates a dual threat: it can promote the development of antibiotic-resistant bacteria, complicating treatment options and threatening public health, while also disrupting aquatic ecosystems and threatening biodiversity [17,18,19,20,21].
Addressing this challenge requires innovative methods for efficiently removing antibiotics from wastewater. Heterogeneous photocatalysis has emerged as a promising technology for the degradation of organic pollutants, including antibiotics, in water [22]. However, the resistance of certain antibiotics, such as ampicillin, to complete degradation remains a significant barrier [23].
This study investigates the potential of Sm-doped CuO–SnO2-based thin films for solar energy and photocatalysis applications, providing a thorough characterization of their structure, morphology, and properties.

2. Experimental Details

2.1. Thin Film Synthesis and Characterization

Undoped and samarium (Sm)-doped CuO–SnO2 mixed oxide thin films were successfully deposited onto glass substrates using the spray pyrolysis technique. All chemical precursors—including copper(II) chloride (CuCl2), tin(IV) chloride (SnCl4), samarium(III) chloride (SmCl3), and methanol—were sourced from Sigma-Aldrich, with a purity exceeding 99%, ensuring the fabrication of high-quality thin films.
Glass substrates (2.5 × 2.5 cm2) were first cleaned in an ultrasonic bath for 15 min to remove surface contaminants. Two individual precursor solutions were then prepared: CuCl2 was dissolved in distilled water to serve as the copper source, while SnCl4 was dissolved in methanol to serveas the tin precursor. Both precursor concentrations were fixed at 0.2 mol·L−1. To enhance the electrical conductivity of the resulting SnO2films, 5 g of fluoride were added to the tin precursor solution.
The copper and tin precursor solutions were subsequently mixed at a molar ratio of [Cu]/[Cu + Sn] = 0.75. The resulting solution was subjected to vigorous magnetic stirring for 20 min to ensure complete homogenization, an essential step in achieving uniform film composition and optimal film quality.
The film deposition process was performed by spraying the prepared solution onto preheated glass substrates. A nozzle was positioned approximately 28 cm above the substrate, with a spray flow rate maintained at ~10 mL/min. The substrate temperature was held constant at 350 °C to facilitate efficient film adhesion and growth.
For the fabrication of Sm-doped CuO–SnO2:F thin films, the same procedure was followed, with the addition of SmCl3 to the precursor solution. Samarium was introduced in varying molar concentrations (0 to 6 mol%, in 2 mol% increments) to study the effect of dopant levels on the structural and functional properties of thin films.
To investigate the physical properties of Sm-doped CuO–SnO2 thin films, several characterization techniques were employed. Structural analysis was conducted using X-ray diffraction (XRD) with a monochromatic diffractometer type X’pert PRO with CuKα radiations (1.5418 Å) and Raman spectroscopy using a Jobin Yvon T64000 spectrometer with a 488 nm argon laser excitation at room temperature. Surface morphology was analyzed with a Cari Zeiss GmbH T3447 scanning electron microscope (SEM). Additionally, transmission electron microscopy (TEM) was conducted using a JOEL 2010 device to further analyze the structural and morphological features of the thin films. X-ray photoelectron spectroscopy (XPS) was conducted usinga using a THERMO-VG ESCALAB 250 X-ray Photoelectron Spectroscopy (XPS) system (Thermo Fisher Scientific, Waltham, MA, USA). Specific surface areas(SBET) were calculated using a multipoint Brunauer–Emmett–Teller (BET) method, utilizing adsorption data within the relative pressure P/P0 range of 0.05–0.25. Optical properties were measured using a PerkinElmer Lambda 950 UV/Vis/NIR spectrophotometer in the wavelength range of 200 to 2000 nm at room temperature. Photoluminescence properties were studied with a PerkinElmer LS55 system using an excitation wavelength of 300 nm.

2.2. Numerical Model Description

The development of advanced solar cell architecture requires a comprehensive understanding of material behavior, which is often complex and expensive to investigate through experimental methods alone. Simulation tools such as Silvaco TCAD offer a practical and cost-effective alternative by enabling accurate modeling of semiconductor device performance under diverse operating conditions. In this study, Silvaco TCAD was employed to simulate the proposed solar cell structure using input parameters derived from experimentally measured electrical and optical properties.
The simulated device incorporated molybdenum and fluorine-doped tin oxide (FTO) as ohmic contacts, a Sm–CuO–SnO2:F composite as the buffer layer, ZnO as the absorber, and material X as the optical window. Device performance was evaluated under standard AM 1.5 solar illumination at an intensity of 0.1 W/cm2, utilizing advanced numerical techniques such as the Newton method. This simulation-driven approach facilitates efficient design optimization and performance evaluation, significantly accelerating the development of next-generation solar cell technologies.

2.3. Pharmaceutical Application (Antibiotic Degradation)

Investigation into the photocatalytic behaviors of undoped and Sm-doped CuO–SnO2:F thin films started with their immersion in a beaker containing 50 mL ampicillin solution. Initially, thin films were allowed to sit in darkness for 30 min to establish an equilibrium between adsorption and desorption processes. Subsequently, they were exposed to sunlight to initiate the photocatalytic reaction. The setup allowed the study of thin films interacting with dye molecules in both dark and light, offering insights into their photocatalytic efficiency.

3. Results

3.1. Structural Properties

The crystallographic properties of undoped and Sm-doped CuO–SnO2 mixed oxides thin films were assessed using X-ray diffraction (Figure 1) within the scan range (2θ) of 20° to 70°.
First, it is evidentthat the XRD patterns showed well-resolved peaks, indicating high-quality crystallites in all the samples. Additionally, all the thin layers displayed a polycrystalline nature, as shown by the presence of prominent CuO peaks (–111), (–202), (202), and (–113), which correspond to the monoclinic crystal structure of copper oxide (JCPDS No. 89-5895) [24,25,26]. These were accompanied by SnO2 peaks at (110), (200), and (211), which are specificto the tetragonal structure (JCPDS No. 79-2205) [27].
Notably, all the diffraction peaks clearly corresponded to the monoclinic CuO and tetragonal SnO2:F phases, with no additional peaks observed, confirming the successful synthesis of CuO–SnO2:F compounds. The absence of solid solution phases suggests the formation of a CuO–SnO2 nanocomposite structure, potentially forming p–n nanojunctions [28]. Moreover, the widening of the diffraction peaks, especially SnO2 (110), indicates the presence of exceptionally fine crystallite sizes within the films, further shedding light on their structural characteristics.
The XRD data provided the average crystallite size D of the films using the Scherrer equation [29,30]:
D = k λ β c o s ( θ )
where k is the constant (k = 0.9), λ is the wavelength of the X-ray beam (λ = 0.15406 nm), β is the width at half height of the maximum peak (FWHM) of the diffraction peak, 2θ is the angle of the diffraction line, and θ is the Bragg angle corresponding to the diffraction peak. To further investigate the crystal structure and defects, the micro-strain, ε, and dislocation density δ were calculated according to the following equations [31,32]:
ε = β c o s ( θ ) 4
δ = 1 D 2
Moreover, the lattice parameters were calculated using the MAUD software version 2.999.
The structural constants are represented in Table 1, which shows that the smallest crystallite size was obtained for the 6% Sm-doped CuO-SnO2:F thin film, indicating its potential for photocatalytic applications. Smaller crystallite sizes are beneficial as they reduce the recombination probability of electron–hole pairs; the shorter diffusion paths allow these charge carriers to reach the semiconductor surface more quickly, thereby enhancing photocatalytic activity [33]. Furthermore, Table 1 reveals that with increasing Sm doping, the dislocation density also increases. This increase in dislocation density can induce structural distortions, leading to a higher surface-to-volume ratio and the formation of internal electric fields. These fields facilitate charge separation and migration, reducing recombination rates and thus improving the overall efficiency of redox reactions on the catalyst surface. In addition, the lattice parameters were found to increase with higher Sm doping levels. This expansion was attributed to the substitution of Cu2+ ions (ionic radius ~0.73 Å) by larger Sm3+ ions (ionic radius ~1.08 Å), which introduced strain and distortion into the crystal lattice. As a result, the unit cell dimensions increased, confirming the successful incorporation of Sm3+ into the CuO lattice and its impact on the structural properties of the films [34].
Raman spectroscopy was employed to further analyze the structural and vibrational properties of the CuO–SnO2 thin films before and after Sm doping. The spectra (Figure 2) revealed two distinct peaks centered around 310 cm−1 and 350 cm−1, which correspond to characteristic vibrational modes of CuO and SnO2 phases. Additionally, a broad peak was observed around 1150 cm−1, likely associated with multi-phonon scattering or defect-related vibrational modes.
Upon Sm doping, Raman peaks exhibited a noticeable red shift, indicating a slight expansion in the lattice parameters. This shift is typically associated with a reduction in vibrational frequency due to the substitution of smaller Cu2+ ions with larger Sm3+ ions, which leads to longer and slightly weaker bonds within the lattice. This interpretation is consistent with the XRD results, which showed an increase in lattice constants upon Sm incorporation. Moreover, the intensity and sharpness of the Raman peaks improved significantly in the doped sample, suggesting enhanced crystallinity. The vibrational changes observed upon doping further support the idea of improved structural order and reduced disorder, in line with the formation of a more uniform and well-crystallized film structure [35].

3.2. Morphological Properties

The surface morphology of undoped and 6% Sm-doped CuO–SnO2:F thin films was thoroughly examined using scanning electron microscopy (SEM) (Figure 3a,b). The observation revealed distinctive rounded and spherical shapes in both undoped and Sm-doped samples. Moreover, the porosity of the films was significantly increased after doping. This enhanced porosity is beneficial in various applications, particularly in catalysis, as it provides a larger surface area for catalytic reactions. The increased porosity improves the accessibility of reactants to active sites, which may lead to better catalytic performance and efficiency [36,37].
The TEM results presented in Figure 3c,d confirmed the morphological features previously observed in the SEM analysis, revealing well-defined and uniformly distributed nanoparticles. The micrographs clearly showed the grain-like structure and compact arrangement consistent with the surface texture observed in the SEM images. Moreover, the crystallite sizes determined from the TEM images were approximately 25 nm and 17 nm for the undoped and Sm-doped CuO–SnO2:F thin films, respectively. These values closely match those estimated from the XRD data using the Scherrer equation, further validating the consistency between structural and morphological characterization.
Elemental mapping analysis was conducted to examine the spatial distribution of elements within the CuO–SnO2 thin films both before and after Smdoping (6%) (Figure 3e,f). In the undoped samples, the mapping clearly confirmed the uniform presence of Cu, Sn, and O across the entire substrate surface, indicating a well-distributed film composition. After Sm doping, the mapping results additionally revealed the successful incorporation of Sm, with a homogenous distribution similar to the primary elements. The even coverage of Cu, Sn, O, and Sm across the substrate confirms not only the uniform deposition of the film but also suggests strong interaction and integration of Sm within the CuO–SnO2 matrix. This elemental interaction is likely to play a role in modifying the structural and electronic properties of the film, as supported by the enhanced crystallinity and spectral shifts observed in complementary analyses.
Energy-dispersive X-ray spectroscopy (EDS) analysis was carried out to confirm the elemental composition of the undoped and 6% Sm-doped CuO–SnO2 thin films. For the undoped film, the EDS spectrum (Figure 3g,h) confirmed the presence of copper (Cu), tin (Sn), and oxygen (O) as expected. In the Sm-doped sample, the EDS spectrum (Figure 3c) revealed the additional presence of samarium (Sm), confirming the successful incorporation of Sm into the CuO–SnO2 matrix. The detection of samarium indicates that the doping process was effective. Furthermore, minor peaks corresponding to elements such as silicon (Si) and sodium (Na) were observed, likely originating from the glass substrate used during the film deposition process [38].

3.3. BET Surface Area

The Brunauer–Emmett–Teller (BET) surface area is a critical parameter that directly impacts the performance of materials in applications such as catalysis, adsorption, and energy storage. BET analysis is commonly used to determine the specific surface area of porous and nanostructured materials by measuring nitrogen adsorption–desorption isotherms. A higher surface area typically correlates with a higher number of active sites available for chemical reactions, which is particularly important in photocatalysis.
In our previous study, the BET surface area of the undoped CuO–SnO2 thin film was found to be 13.4 m2/g [25]. In the current work, the BET surface area of the 6% Sm-doped CuO–SnO2 thin film, shown in Figure 4, was significantly higher at 40.6 m2/g. This substantial increase can be attributed to structural changes induced by Sm doping, such as smaller grain sizes, enhanced porosity, or modified crystallite boundaries—all contributing to a more developed surface morphology [39].
The increase in surface area provides a larger number of active sites for photocatalytic reactions, leading to improved degradation of pollutants. Moreover, a higher surface area enhances light absorption and facilitates more effective interaction between the photocatalyst and the reactant molecules in the solution. Consequently, the photocatalytic efficiency of the Sm-doped CuO–SnO2 thin film is significantly improved compared to the undoped version [40].

3.4. XPS Studies

To investigate the effects of Sm doping on the CuO–SnO2 thin films, X-ray photoelectron spectroscopy (XPS) analysis was performed on both the undoped and 6% Sm-doped samples. In both cases (Figure 5), characteristic peaks for Sn 3d (at 486 and 495 eV), Cu 2p (at 934 and 944 eV), and O 1s (at 529 eV) were clearly observed, confirming the presence of the main constituent elements. In the Sm-doped sample (6%), an additional peak corresponding to Sm 3d appeared around 1083 eV, which was not present in the undoped film—confirming the successful incorporation of Sm into the lattice [41].
Furthermore, the O 1s peak in the doped film showed broadening and asymmetry at higher binding energies, indicating the formation of oxygen vacancies or other Sm-induced defect states. These defects can create localized band tail states near the conduction or valence bands, which are known to enhance charge separation and prolong carrier lifetimes—beneficial for applications such as photocatalysis and optoelectronics [42].

3.5. Optical Properties

Designing and evaluating optoelectronic devices require an understanding of the optical properties of materials. In this study, we analyzed the optical transmission and reflection spectra of CuO–SnO2:F coupled oxide thin films across a broad wavelength range of 250–2000 nm to assess the impact of samarium (Sm) incorporation on optical performance and material quality. Figure 6a presents the optical transmission behavior of the undoped and Sm-doped CuO–SnO2:F thin films at various Sm doping levels. The undoped films exhibited high absorbance, particularly in the UV and visible regions (250–600 nm), where transmission approached zero percent, indicating strong light absorption. As the Sm doping level increased, the intrinsic absorption shifted to blue. The transmission reached a peak of 4% before shifting back to higher wavelengths at the 6% doping level, at which point the behavior resembled that of the undoped CuO–SnO2:F thin films.
The initial shift in intrinsic absorption towards lower wavelengths with increasing Sm doping can be attributed to the Burstein–Moss effect. In this case, the addition of Sm ions increases the free carrier concentration, pushing the Fermi level deeper into the conduction band and effectively widening the optical bandgap [43]. This results in a blue shift, as higher-energy photons are required for electron excitation. However, at the 6% Sm doping level, the absorption shifted back to higher wavelengths. This shift was due to the higher dislocation density, which often leads to greater structural disorder, as reflected in the structural parameters summarized in Table 1. From this, we deduced that the highest dislocation density occurred at the 6% Sm doping level. Excessive Sm incorporation introduced defects, band tail states, impurity scattering, and localized energy states near the conduction band, which reduced the bandgap and enabled the absorption of lower-energy photons, causing a red shift.
Additionally, a notable increase in transmittance was observed with rising Sm concentration, particularly in the visible region, where transmission reached a maximum of 70% for the optimal doping level. However, at the 6% Sm doping level, transmittance started to decrease, dropping to 30%. This suggests that excessive doping may induce structural or electronic changes that negatively impact optical transparency.
Figure 6b shows reflection spectra in which all the thin films display a similar pattern: decreased reflection in the UV region and a plateau in the visible range. This indicates consistent reflection properties across various samarium doping levels, particularly within the visible spectrum.
The absorbance spectra (Figure 6c) of the undoped and samarium (Sm)-doped CuO–SnO2 thin films revealed an interesting trend: absorbance decreased up to the doping level of 4%, after which it began to increase again at 6%. This behavior suggests that Sm doping initially reduces the material’s absorption of incident light, possibly due to enhanced transparency or a reduction in light scattering effects caused by the doping process. However, beyond the 4% doping level, there is a reversal in this trend, where absorbance starts to increase. This increase could be attributed to various causes, such as modifications in the electronic structure or the introduction of doping-induced defects within the material.
We calculated the energy bandgap by analyzing the transmittance and reflection spectra using the Tauc formula, which is expressed by the following equation [44,45]:
α h v = A h v E g n
where A is the constant related to the material, h is Planck’s constant, n corresponds to ½ for direct transitions and 2 for indirect transitions, and α was calculated using the following equation [46,47]:
α = 1 e ln ( 1 R ) 2 T
The optical bandgap values are provided in Table 2.
As shown in Figure 6d, the band gap increases rapidly with the increase in Sm incorporation, reaching a maximum value of 2.52 eV for thin films fabricated with a doping level of 4%. However, beyond this point, the band gap begins to decrease, reaching 2.04 eV at a doping level equal to 6%, which makes it suitable for use as the buffer layer in solar cell structures [12].
Finally, to complete our optical analysis, we examined the optical parameters of the thin films, including the refractive index n (λ), the extinction coefficient k (λ), the real (εr) and imaginary (εi) parts of the dielectric constant (ε), and the volume energy loss function (VELF). All these parameters were determined using the following equations [48,49]:
n = n + ik
n = 1 + R 1 R + 4 R 1 R 2 k 2
ε = εr + iεi
εr = n2k2
εi = 2nk
V E L F = I m 1 ԑ r + i ԑ i = ԑ i ԑ r 2 + ԑ i 2
In the investigation of optical parameters of the undoped and Sm-doped CuO–SnO2:F thin films, doping with samarium resulted in a stable refractive index around 1.6, contrasting with the higher value of 1.8 observed in the undoped thin films (Figure 6e). This consistency in refractive index suggests that introducing Sm ions does not significantly alter the material’s optical transparency or its ability to propagate light. Moreover, the extinction coefficient (k) decreased and maintained near-zero values in the infrared (IR) region for the Sm-doped films (Figure 6f), indicating reduced light absorption and contributing to enhanced transparency. Additionally, both the real and imaginary parts of the dielectric constant followed similar trends to the refractive index and extinction coefficient, respectively (Figure 6g,h).
The volume energy loss function (VELF), which quantifies energy dissipation mechanisms within the material, is crucial for analyzing optical and electronic properties. Figure 6i shows that the VELF reached its minimum at the 6% Sm doping level in the CuO–SnO2 thin films, indicating reduced plasmon energy losses and lower electronic damping. This is highly beneficial for optoelectronic applications, as a lower VELF enhances charge carrier mobility, dielectric stability, and overall energy efficiency, making 6% Sm-doped CuO–SnO2 an excellent candidate as a green buffer layer for high-performance solar cells.
Figure 7 presents the photoluminescence spectra, providing an important understanding of the emission behavior of these films. PL spectra of the undoped and Sm-doped CuO-SnO2:F revealed three distinct peaks that provide insights into the electronic and defect states within the material.
The observed peaks at 395 nm, 410 nm, and 490 nm can be attributed to different phenomena. The peak at 395 nm is likely due to the near-band-edge emission, which is associated with the recombination of excitons (electron–hole pairs) close to the conduction band edge of SnO2. The 410 nm peak may result from shallow defect states or donor–acceptor pair recombination, indicating the presence of defects in the crystalline film, specifically missing or extra oxygen atoms within the SnO2 matrix. Finally, the 490 nm peak is generally attributed to deep-level defect states, which could be associated with complex defect centers or impurities within the CuO-SnO2 mixed oxide [50]. The decrease in photoluminescence peak intensity after doping with samarium can be attributed to several factors. Samarium ions introduce non-radiative recombination centers in the host material. These centers capture charge carriers (electrons and holes) that would otherwise recombine radiatively to emit light, thus reducing the overall PL intensity. Additionally, samarium ions increase scattering of the emitted photons, further diminishing the observed PL signal. Doping can also alter the crystal structure or introduce defects, both of which can adversely affect the material’s electronic properties and reduce the efficiency of the radiative recombination processes that are responsible for photoluminescence [51].

3.6. Solar Cell Simulation

Solar cells offer a renewable alternative to fossil fuels. Among the conventional absorber layers used in thin-film solar cells, materials such as Cu (In,Ga)Se2 (CIGS), CdTe, and Cu2ZnSnS4 (CZTS) have gained significant attention due to their high efficiency and stability [52,53,54,55,56]. However, the CdS buffer layer in many of these solar cells raises environmental concerns due to the toxicity of cadmium. To address this issue, researchers have been exploring alternative buffer layers that maintain high efficiency while being environmentally friendly. Building on our previous studies, we explored the replacement of CdS with CuO–SnO2:F thin films, achieving an efficiency of 15.31% [25]. In this work, we further enhanced the performance by doping CuO–SnO2:F with 6% samarium (Sm), achieving key parameters of 0.72 V open-circuit voltage, 30 mA/cm2 short-circuit current, and an 86% fill factor, which resulted in an improved efficiency of 15.98%, as shown in Table 3. The structure and the mesh of our simulated solar cell are illustrated in Figure 6a.
The increase in solar cell efficiency from 15.31% to 15.98% after 6% Sm doping of CuO–SnO2:F is attributed to several key factors. Samarium doping enhances electrical conductivity by increasing charge carrier concentration, reducing series resistance, and improving electron transport [25]. It also optimizes the band alignment between the buffer and absorber layers, facilitating better charge separation and minimizing recombination losses. Additionally, Sm acts as a defect passivator, reducing trap states and extending carrier lifetimes, leading to higher open-circuit voltage (0.72 V) and fill factor (86%). The slight change in optical properties ensures high transparency, allowing more photons to reach the absorber layer, which maintains a strong short-circuit current density (30 mA/cm2) [57].
In addition to integrating Sm-doped CuO–SnO2:F as a Cd-free buffer layer in CIGS-based solar cells, we also evaluated its performance with other widely used absorber materials, namely CdTe and CZTS, to assess its compatibility and efficiency across different photovoltaic technologies, as shown in Figure 8b. The results are particularly promising, demonstrating that this buffer layer effectively supports multiple absorber layers while maintaining competitive efficiency levels.
For the CdTe-based solar cells, the recorded parameters included an open-circuit voltage (Voc) of 0.28 V, a short-circuit current density (Jsc) of 12.66 mA/cm2, a fill factor (FF) of 70%, and an overall efficiency of 9.36%. While the Voc was relatively low, these results indicate that the buffer layer still facilitated efficient charge transport, and the moderate FF suggests controlled recombination losses.
On the other hand, the CZTS-based solar cells showed significantly better results, with a Voc of 1.39 V, aJsc of 12.72 mA/cm2, an FF of 72%, and an improved efficiency of 13.22%. The notably higher Voc in CZTS indicates excellent band alignment between the Sm-doped CuO–SnO2:F buffer layer and the CZTS absorber, effectively reducing interface recombination and enhancing charge separation [58]. Additionally, the FF values of 70% for CdTe and 72% for CZTS indicate that this buffer layer contributes to stable charge extraction with minimal resistive losses.
These findings are particularly significant as they confirm that Sm-doped CuO–SnO2:F is not only an effective substitute for the toxic CdS, but also a highly adaptable buffer layer that can work efficiently with multiple absorber materials. This enables the creation of eco-friendly, high-performance thin-film solar cells [59].

3.7. Antibiotic Degradation: Ampicillin

Ampicillin (AMP), a commonly used antibiotic, persists in aquatic ecosystems and can have toxic effects when released into water [60,61]. Consequently, there has been a growing interest in developing advanced methods to effectively degrade such pharmaceutical pollutants. Among these methods, photocatalysis powered by solar light and using metal oxide thin films is the most promising due to its lowcost and eco-friendly nature to combat water contamination [62]. These materials are known for their chemical stability and ability to generate highly reactive species that can break down organic contaminants [63].
CuO–SnO2-based photocatalysts have demonstrated potential in the degradation of organic pollutants [24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64], but they have not been explored for ampicillin degradation in previous studies. Notably, this work represents the first attempt to investigate the effects of doping CuO–SnO2:F with samarium (Sm), a rare-earth element, to improve its photocatalytic performance for the degradation of ampicillin and achieve higher degradation efficiencies. Samarium can enhance the material’s photocatalytic activity in several ways. First, Sm ions can introduce new energy levels within the bandgap of the composite, which can act as electron trap sites, reducing the recombination of charge carriers (electrons and holes). By minimizing recombination, more charge carriers remain available to participate in the degradation process. Additionally, Sm doping can improve the absorption of light, especially in the visible region, due to the presence of f-electrons in Sm ions, which help absorb a broader spectrum of light [65]. This allows the photocatalyst to be activated more efficiently under visible light, which is more abundant and practical for real-world applications.
The enhancement in photocatalytic activity is clearly reflected in the experimental results presented in Figure 9a. As shown in Figure 9b, the absorption spectra of ampicillin showed a rapid decline in the intensity peak, indicating significant degradation of the antibiotic over time. The photocatalytic efficiency is quantified by the ratio of the initial concentration (C0) to the final concentration (C) of ampicillin after a certain time using the following equation [66]:
e f f e c i e n t y = C 0 C C 0 100
Notably, the 6% Sm-doped CuO-SnO2:F thin films exhibited remarkable performance, achieving 86% degradation efficiency after 2 h of sunlight irradiation (Figure 9b). This high level of efficiency can be credited to the effects of Sm doping, which not only enhanced the material’s light absorption, but also improved charge carrier separation, resulting in a more efficient degradation of ampicillin. Additionally, increased porosity provided more active sites for catalytic reactions, as shown in Figure 9.
We compared the efficiency of our mixed oxide thin film with earlier works demonstrating the highest photocatalytic efficiency for ampicillin degradation, as shown in Table 4.
A comparative analysis of various metal oxide-based photocatalysts reveals significant differences in degradation efficiency under sunlight irradiation over a 2 h period, depending on the material composition and deposition techniques. Pure ZnO, synthesized via hydrothermal methods, demonstrated a relatively modest degradation efficiency of 41% [67]. In contrast, a ZnO/ZnWO4 nanocomposite prepared using an activated carbon-supported method significantly improved the efficiency to 83% [68], highlighting the positive impact of composite formation and support materials. Similarly, an 8% bismuth-doped CuO–ZnO film fabricated via spray pyrolysis achieved a degradation efficiency of 77% [26], emphasizing the role of dopants in enhancing photocatalytic performance. Notably, our study on 6% samarium-doped CuO–SnO2, also synthesized using spray pyrolysis, achieved the highest degradation efficiency of 86%, demonstrating the synergistic effects of Sm doping and the CuO–SnO2 heterostructure in boosting photocatalytic activity under sunlight.
The degradation of ampicillin using the undoped and Sm-doped CuO-SnO2 photocatalysts followed the same mechanism facilitated by photocatalytic reactions (Figure 9c). Upon irradiation with light (hv), the CuO–SnO2-based photocatalyst absorbs energy, generating electron–hole pairs in the CuO and SnO2 components. This process is represented by the following equation [69]:
CuO-SnO2 + hv →CuO(e + h+)/SnO2
Here, the photogenerated electrons (e) and holes (h+) are separated, with the electrons migrating to the SnO2 component and the holes remaining in the CuO component. The subsequent reactions occur as follows:
CuO (e + h+)/SnO2 → CuO (h+)/SnO2(e)
The electrons on SnO2 reduce oxygen (O2) to form superoxide radicals (O2), while the holes on CuO oxidize water molecules to form hydroxyl radicals (OH):
SnO2(e) + O2 → O2 + SnO2
CuO (h+) + H2O → h+ + OH + CuO + H2
These reactive species, O2 and OH, are highly reactive and play a significant role in breaking down ampicillin. Superoxide radicals (O2) react with ampicillin to produce byproducts such as ammonia (NH3), sulfate ions (SO42−), carbon dioxide (CO2), and water (H2O):
O2 + AMP → Products (NH3 + SO42− + CO2 + H2O)
Similarly, the holes (h+) also interact with ampicillin, further breaking it down into similar byproducts:
h+ + AMP → Products (NH3 + SO42− + CO2 + H2O)
Thus, the photocatalytic degradation of ampicillin involves the generation of holes and reactive oxygen species that attack ampicillin molecules, leading to their complete mineralization into harmless byproducts such as ammonia, sulfate, carbon dioxide, and water. This mechanism describes how the CuO–SnO2:F composite enhances the photocatalytic process, making it effective for environmental remediation. As a result, the incorporation of rare-earth ions such asSmions into the mixed oxide structure is a powerful strategy for optimizing photocatalytic materials, improving their efficiency and sustainability for environmental remediation. This approach is particularly valuable for water treatment processes targeting persistent pollutants such as ampicillin, offering a promising solution for the degradation of pharmaceutical contaminants in aquatic ecosystems.
To assess the role of various reactive species in the catalytic reaction, we performed a series of scavenger tests using isopropanol, ascorbic acid, Na2SO4, and EDTA. These scavengers were used to selectively quench hydroxyl radicals (HO•), superoxide radicals (O2), hydrated electrons (e), and photogenerated holes (h+), respectively. The experimental results (Figure 9d) showed a significant decrease in degradation rates with the addition of each scavenger, indicating the active involvement of these species in the photocatalytic degradation process. This confirms that multiple radical-mediated pathways are operating, supporting the proposed photocatalytic mechanism and highlighting the complex interactions of reactive intermediates during the reaction.

4. Conclusions

In conclusion, this study successfully fabricated undoped and samarium-doped CuO–SnO2:F thin films using the spray pyrolysis technique, advancing the field significantly. The detailed investigation of samarium doping levels revealed notable enhancements in the physical properties of the thin films. Numerical simulation of a novel solar cell structure incorporating 6% Sm–CuO–SnO2 thin films as a green buffer layer demonstrated the highest efficiency of 15.98% when paired with a CIGS absorber layer. Additionally, the 6% Sm–CuO–SnO2 thin films exhibited a high degradation efficiency of 86% for ampicillin after two hours of exposure, highlighting their dual potential functions in both environmental remediation and solar cell applications.

Author Contributions

Conceptualization, N.T.K.; software modeling, B.Y.; photocatalysis tests, S.G.D.; experimental work by spray pyrolysis, G.C.; synthesis by spray pyrolysis, M.H.; characterization, G.C., B.A., and M.H.; investigation, G.C.; film analysis, M.H.; writing—original draft preparation, G.C.; writing—editing and figure preparation, B.A.; writing—review and editing, R.V.; supervision, R.V. and N.T.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was funded by Umm Al-Qura University, Saudi Arabia, under grant number 25UQU4350568GSSR01.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

The authors extend their appreciation to Umm Al-Qura University, Saudi Arabia, for funding this research through grant No. 25UQU4350568GSSR01.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
TLAThree-letter acronym
LDLinear dichroism

References

  1. Zhang, H.; Ji, X.; Yao, H.; Fan, Q.; Yu, B.; Li, J. Review on efficiency improvement effort of perovskite solar cell. Sol. Energy 2022, 233, 421–434. [Google Scholar] [CrossRef]
  2. Romeo, A.; Artegiani, E. CdTe-Based Thin Film Solar Cells: Past, Present and Future. Energies 2021, 14, 1684. [Google Scholar] [CrossRef]
  3. Ngqoloda, S.; Ngwenya, T.; Raphulu, M. Recent Advances on the Deposition of Thin Film Solar Cells; IntechOpen: London, UK, 2025. [Google Scholar]
  4. Ramanujam, J.; Bishop, D.M.; Todorov, T.K.; Gunawan, O.; Rath, J.; Nekovei, R.; Artegiani, E.; Romeo, A. Flexible CIGS, CdTe and a-Si: H based thin film solar cells: A review. Prog. Mater. Sci. 2020, 110, 100619. [Google Scholar] [CrossRef]
  5. Kumar, V.; Prasad, R.; Chaure, N.B.; Singh, U.P. Advancement in Copper Indium Gallium Sel-enide (CIGS)-Based Thin-Film Solar Cells. In Recent Advances in Thin Film Photovoltaics; Springer: Singapore, 2022; pp. 5–39. [Google Scholar]
  6. Tobbeche, S.; Kalache, S.; Elbar, M.; Kateb, M.N.; Serdouk, M.R. Improvement of the CIGS solar cell performance: Structure based on a ZnS buffer layer. Opt. Quantum Electron. 2019, 51, 1–13. [Google Scholar] [CrossRef]
  7. Kanchan, K.; Sahu, A.; Yadav, S. The Improved Performance with Reduction in Toxicity in CIGS Solar Cell Using Ultra-Thin BaSi2 BSF Layer. J. Nano-Electron. Phys. 2023, 15, 02025. [Google Scholar] [CrossRef]
  8. Nkuissi, H.J.T.; Konan, F.K.; Hartiti, B.; Ndjaka, J.-M. Toxic materials used in thin film photovoltaics and their impacts on environment. In Reliability and Ecological Aspects of Photovoltaic Modules; IntechOpen: London, UK, 2020. [Google Scholar]
  9. Soni, P.; Raghuwanshi, M.; Wuerz, R.; Berghoff, B.; Knoch, J.; Raabe, D.; Cojocaru-Mirédin, O. Sputtering as a viable route for In2S3 buffer layer deposition in high efficiency Cu (In, Ga) Se2 solar cells. Energy Sci. Eng. 2019, 7, 478–487. [Google Scholar] [CrossRef]
  10. Charrada, G.; Hajji, M.; Ajili, M.; Garcia-Loureiro, A.; Kamoun, N.T. Unlocking the potential of CIGS solar cells: Harnessing CZTS as a second absorber layer for enhanced performance and sustainability. J. Opt. 2024, 1–10. [Google Scholar] [CrossRef]
  11. Chen, S.; Chen, Y.; Aziz, H.S.; Zhang, H.; Li, Z.; Chen, Y.; Zeng, Y.; Zheng, Z.; Hu, J.; Su, Z.; et al. A Cd-Free Electron Transport Layer Simultaneously Enhances Charge Carrier Separation and Transfer in Sb2Se3 Photocathodes for Efficient Solar Hydrogen Production. Adv. Funct. Mater. 2024, 35, 2420912. [Google Scholar] [CrossRef]
  12. Naffouti, W.; Ben Nasr, T.; Battaglini, N.; Ammar, S.; Turki-Kamoun, N. Effect of samarium doping on the physical properties of chemically sprayed titanium dioxide thin films. J. Renew. Sustain. Energy 2015, 7, 063132. [Google Scholar] [CrossRef]
  13. Chahkandi, M.; Zargazi, M. New water based EPD thin BiVO4 film: Effective photocatalytic degradation of Amoxicillin antibiotics. J. Hazard. Mater. 2020, 389, 121850. [Google Scholar] [CrossRef]
  14. Pant, B.; Park, M.; Park, S.-J. Recent Advances in TiO2 Films Prepared by Sol-gel Methods for Photocatalytic Degradation of Organic Pollutants and Antibacterial Activities. Coatings 2019, 9, 613. [Google Scholar] [CrossRef]
  15. Tiwari, D.; Lee, S.-M.; Kim, D.-J. Photocatalytic degradation of amoxicillin and tetracycline by template synthesized nano-structured Ce3+@TiO2 thin film catalyst. Environ. Res. 2022, 210, 112914. [Google Scholar]
  16. Denissen, J.; Reyneke, B.; Waso-Reyneke, M.; Havenga, B.; Barnard, T.; Khan, S.; Khan, W. Prevalence of ESKAPE pathogens in the environment: Antibiotic resistance status, community-acquired infection and risk to human health. Int. J. Hyg. Environ. Health 2022, 244, 114006. [Google Scholar] [CrossRef] [PubMed]
  17. Kyuchukova, R. Antibiotic residues and human health hazard-review. Bulg. J. Agric. Sci. 2020, 26, 3. [Google Scholar]
  18. Feng, L.; Cheng, Y.; Zhang, Y.; Li, Z.; Yu, Y.; Feng, L.; Zhang, S.; Xu, L. Distribution and human health risk assessment of antibiotic residues in large-scale drinking water sources in Chongqing area of the Yangtze River. Environ. Res. 2020, 185, 109386. [Google Scholar] [CrossRef]
  19. Rahman, M.; Alam, M.-U.; Luies, S.K.; Kamal, A.; Ferdous, S.; Lin, A.; Sharior, F.; Khan, R.; Rahman, Z.; Parvez, S.M.; et al. Contamination of Fresh Produce with Antibiotic-Resistant Bacteria and Associated Risks to Human Health: A Scoping Review. Int. J. Environ. Res. Public Health 2022, 19, 360. [Google Scholar] [CrossRef]
  20. Leonard, A.F.; Morris, D.; Schmitt, H.; Gaze, W.H. Natural recreational waters and the risk that exposure to antibiotic resistant bacteria poses to human health. Curr. Opin. Microbiol. 2021, 65, 40–46. [Google Scholar] [CrossRef]
  21. Huemer, M.; Shambat, S.M.; Brugger, S.D.; Zinkernagel, A.S. Antibiotic resistance and persistence—Implications for human health and treatment perspectives. Embo Rep. 2020, 21, e51034. [Google Scholar] [CrossRef]
  22. Li, S.; Wu, Y.; Zheng, H.; Li, H.; Zheng, Y.; Nan, J.; Ma, J.; Nagarajan, D.; Chang, J.-S. Antibiotics degradation by advanced oxidation process (AOPs): Recent advances in ecotoxicity and antibiotic-resistance genes induction of degradation products. Chemosphere 2022, 311, 136977. [Google Scholar] [CrossRef]
  23. Montoya-Rodríguez, D.M.; Serna-Galvis, E.A.; Ferraro, F.; Torres-Palma, R.A. Degradation of the emerging concern pollutant ampicillin in aqueous media by nonchemical advanced oxidation processes-Parameters effect, removal of antimicrobial activity and pollutant treatment in hydrolyzed urine. J. Environ. Manag. 2020, 261, 110224. [Google Scholar] [CrossRef]
  24. Charrada, G.; Ajili, M.; Jebbari, N.; Kamoun, N.T. Investigation on physical properties of CuO and SnO2: F mixed oxide sprayed thin films for photocatalytic application: Coupling effect between oxides. J. Mater. Sci. Mater. Electron. 2024, 35, 685. [Google Scholar] [CrossRef]
  25. Charrada, G.; Ajili, M.; Jebbari, N.; Hajji, M.; Bernardini, S.; Aguir, K.; Kamoun, N.T. Investigation on thermal annealing effect on the physical properties of CuO-SnO2:F sprayed thin films for NO2 gas sensor and solar cell simulation. Mater. Lett. 2024, 367, 136666. [Google Scholar] [CrossRef]
  26. Hajji, M.; Jebbari, N.; Ajili, M.; Thebti, A.; Ouzari, H.-I.; Garcia-Loureiro, A.; Kamoun, N.T. Bismuth doping for enhanced physical and electrochemical properties of CuO–ZnO thin films for complete degradation of Rifampicin and other antibiotics alongside organic dyes. Opt. Mater. 2024, 157, 116048. [Google Scholar] [CrossRef]
  27. Charrada, G.; Ajili, M.; Jebbari, N.; Bernardini, S.; Aguir, K.; Kamoun, N.T. Improvement of ozone sensing parameters by CuO–SnO2: F mixed oxide sprayed thin films. J. Mater. Sci. Mater. Electron. 2024, 35, 1120. [Google Scholar] [CrossRef]
  28. Hajji, M.; Dabbabi, S.; Ajili, M.; Jebbari, N.; Loureiro, A.G.; Kamoun, N.T. Investigations on physical properties of CuO–ZnO couple oxide sprayed thin films for environmental applications (ozone gas sensing and MB degradation). J. Mater. Sci. Mater. Electron. 2024, 35, 663. [Google Scholar] [CrossRef]
  29. Souli, M.; Reghima, M.; Secu, M.; Bartha, C.; Enculescu, M.; Mejri, A.; Kamoun-Turki, N.; Badica, P. Physical properties investigation of samarium doped calcium sulfate thin films under high gamma irradiations for space photovoltaic and dosimetric applications. Superlattices Microstruct. 2019, 126, 103–119. [Google Scholar] [CrossRef]
  30. Kamoun, O.; Gassoumi, A.; Shkir, M.; Gorji, N.E.; Turki-Kamoun, N. Synthesis and Characterization of Highly Photocatalytic Active Ce and Cu Co-Doped Novel Spray Pyrolysis Developed MoO3 Films for Photocatalytic Degradation of Eosin-Y Dye. Coatings 2022, 12, 823. [Google Scholar] [CrossRef]
  31. Yahmadi, B.; Kamoun, O.; Alhalaili, B.; Alleg, S.; Vidu, R.; Turki, N.K. Physical Investigations of (Co, Mn) Co-Doped ZnO Nanocrystalline Films. Nanomaterials 2020, 10, 1507. [Google Scholar] [CrossRef]
  32. Zaidi, B.; Hajji, M.; Bouarroudj, T.; Akhtar, M.S.; Alam Saeed, M.; Charrada, G.; Hadjoudja, B.; Chouial, B.; Jebbari, N.; Kamoun-Turki, N. Synthesis, Characterization, and Photocatalytic Performance of (Eu, Ni) Co-Doped ZnO Thin Films for Environmental Applications. J. Nano Res. 2024, 86, 77–88. [Google Scholar] [CrossRef]
  33. Hajji, M.; Jebbari, N.; Ajili, M.; Garcia-Loureiro, A.; Vidu, R.; Kamoun, N. Advanced oxidation processes (AOPs): Navigating from MOS to COS and X-COS systems-Unraveling strategies, tackling challenges, and pioneering advances. Opt. Mater. 2024, 152, 115399. [Google Scholar] [CrossRef]
  34. Kumar, N.; Limbu, S.; Sharma, S.; Chand, P.; Yadav, P.; Chauhan, R.N. Impact of Sm3+ on Cu2O: Co and subsequent attributes in structural, optical and electrochemical aspects. Opt. Mater. 2025, 162, 116874. [Google Scholar] [CrossRef]
  35. Srinivasan, K.; Rukkumani, V.; Radhika, V.; Saravanakumar, M.; Kavitha, S. Characterization of CuO–SnO2 composite nano powder by hydrothermal method for solar cell. In Machine Learning and the Internet of Things in Solar Power Generation; CRC Press: Boca Raton, FL, USA, 2023; pp. 123–140. [Google Scholar]
  36. Hedau, B.; Ha, T.-J. Indium-doped iron-coordinated covalent organic framework as an efficient bifunctional oxygen electrocatalyst for energy applications. J. Alloy. Compd. 2024, 1010, 178134. [Google Scholar] [CrossRef]
  37. Gupta, J.; Shaik, H.; Kumar, K.N. A review on the prominence of porosity in tungsten oxide thin films for electrochromism. Ionics 2021, 27, 2307–2334. [Google Scholar] [CrossRef]
  38. Faisal, A.D.; Kalef, W.K.; Salim, E.T.; Alsultany, F.H. Synthesis of CuO/SnO2 NPs on quartz substrate for temperature sensors application. J. Ovonic Res. 2022, 18, 205–212. [Google Scholar] [CrossRef]
  39. Berwal, P.; Rani, S.; Sihag, S.; Singh, P.; Bulla, M.; Jatrana, A.; Kumar, A.; Kumar, A.; Kumar, V. Fabrication of NiO based thin film for high-performance NO2 gas sensors at low concentrations. Phys. B Condens. Matter 2024, 685, 416023. [Google Scholar] [CrossRef]
  40. Hu, Y.; Wang, F.; Yang, Z.; Tang, C.Y. Modeling nanovoid-enhanced water permeance of thin film composite membranes. J. Membr. Sci. 2023, 675, 121555. [Google Scholar] [CrossRef]
  41. Li, K.; Wang, Y.; Wang, S.; Zhu, B.; Zhang, S.; Huang, W.; Wu, S. A comparative study of CuO/TiO2-SnO2, CuO/TiO2 and CuO/SnO2 catalysts for low-temperature CO oxidation. J. Nat. Gas Chem. 2009, 18, 449–452. [Google Scholar] [CrossRef]
  42. Wang, W.; Cao, J.; Wang, S.; Zhang, R.; Zhang, Y. CuO–SnO2 sensor for room-temperature CO detection: Experiments and DFT calculations. Sens. Actuators B Chem. 2024, 420, 136427. [Google Scholar] [CrossRef]
  43. Ivchenko, V. Theory of Burstein-Moss effect in semiconductors with anisotropic energy bands. Phys. Scr. 2024, 99, 035952. [Google Scholar] [CrossRef]
  44. Jrad, A.; Naouai, M.; Ammar, S.; Turki-Kamoun, N. Investigation of molybdenum dopant effect on ZnS thin films: Chemical composition, structural, morphological, optical and luminescence surveys. Mater. Sci. Semi-Conduct. Process. 2021, 130, 105825. [Google Scholar] [CrossRef]
  45. Akkari, A.; Guasch, C.; Castagne, M.; Kamoun-Turki, N. Optical study of zinc blend SnS and cubic In2S3:Al thin films prepared by chemical bath deposition. J. Mater. Sci. 2011, 46, 6285–6292. [Google Scholar] [CrossRef]
  46. Mandal, P.; Singh, U.P.; Roy, S. Optical performance of europium-doped β gallium oxide PVD thin films. J. Mater. Sci. Mater. Electron. 2021, 32, 3958–3965. [Google Scholar] [CrossRef]
  47. Charrada, G.; Ajili, M.; Bernardini, S.; Aguir, K.; Kamoun, N.T. Dual-Functional Green Facile synthesis of graphene dopedCuO-SnO2:F sprayed thin film as an efficient photocatalyst and ammonia gas sensor at low concentration. Ceram. Int. 2025, in press. [Google Scholar] [CrossRef]
  48. Panachikkool, M.; Aparna, E.T.; Asaithambi, P.; Pandiyarajan, T. Design and numerical simulation of CuBi2O4 solar cells with graphene quantum dots as hole transport layer under ideal and non-ideal conditions. Sci. Rep. 2025, 15, 100. [Google Scholar] [CrossRef]
  49. Ajili, M.; Ayed, R.B.; Kamoun, N.T. Structural, optical, photoluminescence and electrical properties of p-CuO/n-ZnO: Sn and p-CuO/n-α-Fe2O3 efficient hetero-junctions for optoelectronic applications. J. Lumin. 2022, 241, 118457. [Google Scholar] [CrossRef]
  50. Marpally, J. Photocatalytic Degradation of Malachite Green and Rhodamine B Dye over SnO2-CuO Binary Metal Oxide Nanocomposite under UV Light. Ph.D. Thesis, Department of Chemistry, National Institute of Technology, Rourkela, Odisha, India, 2015. [Google Scholar]
  51. Zhang, C.; Lin, J. Defect-related luminescent materials: Synthesis, emission properties and applications. Chem. Soc. Rev. 2012, 41, 7938–7961. [Google Scholar] [CrossRef]
  52. Ghadikolaei, S.S.C. Solar photovoltaic cells performance improvement by cooling technology: An overall review. Int. J. Hydrogen Energy 2021, 46, 10939–10972. [Google Scholar] [CrossRef]
  53. Machkih, K.; Oubaki, R.; Makha, M. A Review of CIGS Thin Film Semiconductor Deposition via Sputtering and Thermal Evaporation for Solar Cell Applications. Coatings 2024, 14, 1088. [Google Scholar] [CrossRef]
  54. Barbato, M.; Artegiani, E.; Bertoncello, M.; Meneghini, M.; Trivellin, N.; Mantoan, E.; Romeo, A.; Mura, G.; Ortolani, L.; Zanoni, E. CdTe solar cells: Technology, operation and reliability. J. Phys. D: Appl. Phys. 2021, 54, 333002. [Google Scholar] [CrossRef]
  55. Paul, R.; Shukla, S.; Lenka, T.R.; Talukdar, F.A.; Goyal, V.; Boukortt, N.E.I.; Menon, P.S. Recent progress in CZTS (CuZnSn sulfide) thin-film solar cells: A review. J. Mater. Sci. Mater. Electron. 2024, 35, 226. [Google Scholar] [CrossRef]
  56. Kim, D.S.; Min, B.K. Strategies to Enhance the Performance of Cu (In, Ga)(S, Se) 2 Thin-Film Solar Cells by Doping Approaches. Korean J. Chem. Eng. 2024, 41, 3771–3781. [Google Scholar] [CrossRef]
  57. Yang, A.; Hou, Q.; Yin, X.; Sha, S. First-principle study of the effects of biaxial strain on the photocatalytic and magnetic mechanisms of ZnO with Sm doping and point defects (VZn, Hi). Vacuum 2021, 189, 110225. [Google Scholar] [CrossRef]
  58. Shah, U.A.; Wang, A.; Irfan Ullah, M.; Ishaq, M.; Shah, I.A.; Zeng, Y.; Abbasi, M.S.; Umair, M.A.; Farooq, U.; Liang, G.-X.; et al. A deep dive into Cu2ZnSnS4 (CZTS) solar cells: A review of exploring roadblocks, breakthroughs, and shaping the future. Small 2024, 20, 2310584. [Google Scholar] [CrossRef] [PubMed]
  59. Chen, D.; Yang, G.; Shao, Z.; Ciucci, F. NanoscaledSm-doped CeO2 buffer layers for intermediate-temperature solid oxide fuel cells. Electrochem. Commun. 2013, 35, 131–134. [Google Scholar] [CrossRef]
  60. Al-Musawi, T.J.; Rajiv, P.; Mengelizadeh, N.; Arghavan, F.S.; Balarak, D. Photocatalytic efficiency of CuNiFe2O4 nanoparticles loaded on multi-walled carbon nanotubes as a novel photocatalyst for ampicillin degradation. J. Mol. Liq. 2021, 337, 116470. [Google Scholar] [CrossRef]
  61. de Carvalho Costa, L.R.; de Oliveira, J.T.; Jurado-Davila, V.; Féris, L.A. Degradation of ampicillin by combined process: Adsorption and Fenton reaction. Environ. Technol. Innov. 2022, 26, 102365. [Google Scholar] [CrossRef]
  62. Yde Carvalho Costa, L.R.; de Oliveira, J.T.; Jurado-Davila, V.; Féris, L.A. Caffeine and ampicillin degradation by ozonation: Addressing pathways, performance and eco-toxicity. Chem. Eng. Sci. 2024, 288, 119817. [Google Scholar] [CrossRef]
  63. Nguyen, V.N.; Wang, S.L.; Nguyen, T.H.; Nguyen, V.B.; Doan, M.D.; Nguyen, A.D. Preparation and Characterization of Chitosan/Starch Nanocomposites Loaded with Ampicillin to Enhance Antibacterial Activity against Escherichia coli. Polymers 2024, 16, 2647. [Google Scholar] [CrossRef]
  64. Batool, I.; Albalawi, K.; Khan, A.U.; Tahir, K.; Khan, Z.U.H.; Zaki, M.E.; Saleh, E.A.M.; Alabbad, E.A.; Althagafi, T.M.; Abdulaziz, F. The construction of novel CuO/SnO2@g-C3N4 photocatalyst for efficient degradation of ciprofloxacin, methylene blue and photoinhibition of bacteria through efficient production of reactive oxygen species. Environ. Res. 2023, 231, 116086. [Google Scholar] [CrossRef]
  65. Singh, J.; Singh, G.P.; Kumar, S.; Jain, R.K.; Gasso, S.; Singh, B.; Singh, K.; Singh, A.; Singh, R.C. Probing structural, optical and magnetic properties of Sm-doped ZnO nanomaterials via experimental and DFT approach: Enhanced photocatalytic degradation and antibacterial performance. Colloids Surfaces A Physicochem. Eng. Asp. 2023, 668, 131470. [Google Scholar] [CrossRef]
  66. Li, G.; Wang, R.; Wang, B.; Zhang, J. Sm-doped mesoporous g-C3N4 as efficient catalyst for degradation of tylosin: Influencing factors and toxicity assessment. Appl. Surf. Sci. 2020, 517, 146212. [Google Scholar] [CrossRef]
  67. Nosrati, R.; Olad, A.; Maramifar, R. Degradation of ampicillin antibiotic in aqueous solution by ZnO/polyaniline nanocomposite as photocatalyst under sunlight irradiation. Environ. Sci. Pollut. Res. 2012, 19, 2291–2299. [Google Scholar] [CrossRef] [PubMed]
  68. Raizada, P.; Kumari, J.; Shandilya, P.; Singh, P. Kinetics of photocatalytic mineralization of oxytetracycline and ampicillin using activated carbon supported ZnO/ZnWO4 nanocomposite in simulated wastewater. Desalination Water Treat. 2017, 79, 204–213. [Google Scholar] [CrossRef]
  69. Hajji, M.; Ajili, M.; Jebbari, N.; Loreiro, A.G.; Kamoun, N.T. Photocatalytic performance and solar cell applications of coupled semiconductor CuO–ZnO sprayed thin films: Coupling effect between oxides. Opt. Mater. 2023, 140, 113798. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of the undoped and Sm-doped CuO–SnO2:F thin films.
Figure 1. X-ray diffraction patterns of the undoped and Sm-doped CuO–SnO2:F thin films.
Technologies 13 00197 g001
Figure 2. Raman spectra of the undoped and 6% Sm-doped CuO–SnO2:F thin films.
Figure 2. Raman spectra of the undoped and 6% Sm-doped CuO–SnO2:F thin films.
Technologies 13 00197 g002
Figure 3. SEM images, TEM spectra, elemental mapping, and EDS of the undoped (a,c,e,g) and 6% Sm-doped CuO–SnO2:F thin films (b,d,f,h).
Figure 3. SEM images, TEM spectra, elemental mapping, and EDS of the undoped (a,c,e,g) and 6% Sm-doped CuO–SnO2:F thin films (b,d,f,h).
Technologies 13 00197 g003aTechnologies 13 00197 g003bTechnologies 13 00197 g003c
Figure 4. BET nitrogen adsorption isotherm plot of the 6% Sm-doped CuO–SnO2:F thin films.
Figure 4. BET nitrogen adsorption isotherm plot of the 6% Sm-doped CuO–SnO2:F thin films.
Technologies 13 00197 g004
Figure 5. XPS spectra of the undoped and 6% Sm-doped CuO–SnO2:F thin films. (a) XPS survey spectrum of the sample, (bf) High-resolution XPS spectra of specific elements.
Figure 5. XPS spectra of the undoped and 6% Sm-doped CuO–SnO2:F thin films. (a) XPS survey spectrum of the sample, (bf) High-resolution XPS spectra of specific elements.
Technologies 13 00197 g005
Figure 6. (a) Transmission (T), (b) reflection (R), (c) absorbance, (d) Tauc plot, (e) refractive index n, (f) extinction coefficient k, (g) real (εr) and (h) imaginary (εi) parts of the dielectric constant (ε), and (i) volume energy loss function (VELF). Spectra of the undoped and Sm–doped CuO–SnO2:F thin films.
Figure 6. (a) Transmission (T), (b) reflection (R), (c) absorbance, (d) Tauc plot, (e) refractive index n, (f) extinction coefficient k, (g) real (εr) and (h) imaginary (εi) parts of the dielectric constant (ε), and (i) volume energy loss function (VELF). Spectra of the undoped and Sm–doped CuO–SnO2:F thin films.
Technologies 13 00197 g006aTechnologies 13 00197 g006b
Figure 7. Pl spectra of the undoped and Sm-doped CuO–SnO2 coupled oxide thin films at different Sm doping levels.
Figure 7. Pl spectra of the undoped and Sm-doped CuO–SnO2 coupled oxide thin films at different Sm doping levels.
Technologies 13 00197 g007
Figure 8. The structure and the mesh (a) and the J–V curve (b) of our solar cell structure using Silvaco.
Figure 8. The structure and the mesh (a) and the J–V curve (b) of our solar cell structure using Silvaco.
Technologies 13 00197 g008aTechnologies 13 00197 g008b
Figure 9. Absorption spectra (a) and photodegradation rates (b) of ampicillin using various catalysts (undoped and samarium-doped CuO–SnO2:F thin films at different samarium concentration), degradation mechanism (c), and scavenger tests (d).
Figure 9. Absorption spectra (a) and photodegradation rates (b) of ampicillin using various catalysts (undoped and samarium-doped CuO–SnO2:F thin films at different samarium concentration), degradation mechanism (c), and scavenger tests (d).
Technologies 13 00197 g009
Table 1. Structural (D and δ are the crystallite size and the dislocation density) parameters of undoped and Sm-doped CuO–SnO2 coupled oxide thin films at different Sm doping levels.
Table 1. Structural (D and δ are the crystallite size and the dislocation density) parameters of undoped and Sm-doped CuO–SnO2 coupled oxide thin films at different Sm doping levels.
Thin LayerD (nm)δ (×10−4 nm−2)Lattice Parameters (A°)
CuO–SnO2:F2812.7a = 4.683
b = 3.421
c = 5.146
2% Sm-doped CuO–SnO2:F3110.4a = 4.685
b = 3.426
c = 5.147
4% Sm-doped CuO–SnO2:F1830.8a = 4.686
b = 3.428
c = 5.149
6% Sm-doped CuO–SnO2:F1544.4a = 4.689
b = 3.429
c = 5.151
Table 2. Optical bandgaps of the undoped and Sm-doped CuO–SnO2 coupled oxide thin films at different Sm doping levels.
Table 2. Optical bandgaps of the undoped and Sm-doped CuO–SnO2 coupled oxide thin films at different Sm doping levels.
Thin LayersGap Energy (eV)
CuO–SnO2:F1.91
2% Sm-doped CuO–SnO2:F2.34
4% Sm-doped CuO–SnO2:F2.52
6% Sm-doped CuO–SnO2:F2.04
Table 3. Simulation parameters of Cu (In,Ga)Se2 (CIGS)-, CdTe-, and Cu2ZnSnS4 (CZTS)-based solar cell absorber layers.
Table 3. Simulation parameters of Cu (In,Ga)Se2 (CIGS)-, CdTe-, and Cu2ZnSnS4 (CZTS)-based solar cell absorber layers.
Absorber LayerVoc (V)Jsc (mA)FF (%)Efficiency (%)
CdTe0.2812.66709.36
CZTS1.3912.727213.22
CIGS0.72308615.98
Table 4. Ampicillin degradation efficiency of various metaloxides.
Table 4. Ampicillin degradation efficiency of various metaloxides.
Metal Oxide Deposition MethodDegradation Efficiency in 2 Hunder Sunlight Irradiation Reference
ZnOHydrothermal synthesis41%[67]
ZnO/ZnWO4 nanocompositeActivated carbon-supported method83%[68]
8% bismuth-dopedCuO–ZnOSpray pyrolysis77%[26]
6% Sm-doped CuO–SnO2Spray pyrolysis86%Our work
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

Charrada, G.; Yahmadi, B.; Alhalaili, B.; Hajji, M.; Derouich, S.G.; Vidu, R.; Kamoun, N.T. Synthesis of Sm-Doped CuO–SnO2:FSprayed Thin Film: An Eco-Friendly Dual-Function Solution for the Buffer Layer and an Effective Photocatalyst for Ampicillin Degradation. Technologies 2025, 13, 197. https://doi.org/10.3390/technologies13050197

AMA Style

Charrada G, Yahmadi B, Alhalaili B, Hajji M, Derouich SG, Vidu R, Kamoun NT. Synthesis of Sm-Doped CuO–SnO2:FSprayed Thin Film: An Eco-Friendly Dual-Function Solution for the Buffer Layer and an Effective Photocatalyst for Ampicillin Degradation. Technologies. 2025; 13(5):197. https://doi.org/10.3390/technologies13050197

Chicago/Turabian Style

Charrada, Ghofrane, Bechir Yahmadi, Badriyah Alhalaili, Moez Hajji, Sarra Gam Derouich, Ruxandra Vidu, and Najoua Turki Kamoun. 2025. "Synthesis of Sm-Doped CuO–SnO2:FSprayed Thin Film: An Eco-Friendly Dual-Function Solution for the Buffer Layer and an Effective Photocatalyst for Ampicillin Degradation" Technologies 13, no. 5: 197. https://doi.org/10.3390/technologies13050197

APA Style

Charrada, G., Yahmadi, B., Alhalaili, B., Hajji, M., Derouich, S. G., Vidu, R., & Kamoun, N. T. (2025). Synthesis of Sm-Doped CuO–SnO2:FSprayed Thin Film: An Eco-Friendly Dual-Function Solution for the Buffer Layer and an Effective Photocatalyst for Ampicillin Degradation. Technologies, 13(5), 197. https://doi.org/10.3390/technologies13050197

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