Next Article in Journal
3D Nanocomposite with High Aspect Ratio Based on Polyaniline Decorated with Silver NPs: Synthesis and Application as Electrochemical Glucose Sensor
Previous Article in Journal
Conferring of Drought and Heat Stress Tolerance in Wheat (Triticum aestivum L.) Genotypes and Their Response to Selenium Nanoparticles Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nanotextured CeO2−SnO2 Composite: Efficient Photocatalytic, Antibacterial, and Energy Storage Fibers

1
Department of Chemistry, College of Science and Arts, Najran University, Najran 11001, Saudi Arabia
2
Promising Centre for Sensors and Electronic Devices (PCSED), Advanced Materials and Nano-Research Centre, Najran University, Najran 11001, Saudi Arabia
3
Department of Chemistry, College of Science, Albaha University, Albaha 65799, Saudi Arabia
4
Department of Biology, College of Science, Albaha University, Albaha 65799, Saudi Arabia
5
Nanostructured and Biomimetic Lab, Department of Nanotechnology, University of Kashmir Hazratbal, Srinagar 190006, India
6
Organic Materials and Fibers Engineering Department, Chonbuk National University, Jeonju 560011, Republic of Korea
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(6), 1001; https://doi.org/10.3390/nano13061001
Submission received: 9 February 2023 / Revised: 1 March 2023 / Accepted: 7 March 2023 / Published: 10 March 2023
(This article belongs to the Section Biology and Medicines)

Abstract

:
Bacterial infections remain a serious and pervasive threat to human health. Bacterial antibiotic resistance, in particular, lowers treatment efficacy and increases mortality. The development of nanomaterials has made it possible to address issues in the biomedical, energy storage, and environmental fields. This paper reports the successful synthesis of CeO2−SnO2 composite nanofibers via an electrospinning method using polyacrylonitrile polymer. Scanning and transmission electron microscopy assessments showed that the average diameter of CeO2−SnO2 nanofibers was 170 nm. The result of photocatalytic degradation for methylene blue dye displayed enhanced efficiency of the CeO2−SnO2 composite. The addition of SnO2 to CeO2 resulted in the enhancement of the light absorption property and enriched charge transmission of photoinduced electron–hole duos, which conspicuously contributed to momentous photoactivity augmentation. Composite nanofibers exhibited higher specific capacitance which may be accredited to the synergism between CeO2 and SnO2 particles in nanofibers. Furthermore, antibacterial activity was screened against Escherichia coli and CeO2−SnO2 composite nanofibers depicted excellent activity. The findings of this work point to new possibilities as an electrode material in energy storage systems and as a visible-light-active photocatalyst for the purification of chemical and biological contaminants, which would substantially benefit environmental remediation processes.

1. Introduction

Globally, environmental pollution is a serious problem, inflicting harm to life on the planet. Water pollution, among other types of pollution, has a significant negative impact on living species, including aquatic life. Water pollution is due to the discharge of harmful organic chemicals, such as dyes, acids, and antibiotics, into drinkable water bodies, such as rivers, lakes, and ponds, from textile, chemical, and pharmaceutical facilities. In nature, the majority of organic molecules are carcinogenic. Furthermore, water pollution sequentially causes soil pollution, which has a direct or indirect impact on daily living [1,2]. Synthetic color dyes, particularly those generated during textile washes, combine easily with water in comparison to chemicals and reagents; hence, the combination of effluents has hazardous potential. As a result, industrial effluent must be processed prior to being disposed of in the surroundings [3]. Dyes are colored aromatic organic complexes that capture light and provide color [4]. Because of these advantages, various dyes are utilized for a variety of applications in commerce, such as fabrics, foodstuffs, rubber, lithography, makeup, medication, plastic, concrete, and paper. Companies produce massive amounts of effluent, including carcinogenic and poisonous dyes, that pollute water and render it unsafe for human consumption. The textile sector is the largest dye-consuming business, relying on textile dyes, which are very complex molecules [5].
Methylene blue (MB), which is extensively used to color different fabrics, such as silk, wood, and cotton, as well as paper, is one of the most consumed ingredients in dye commerce [6]. Textile manufacturers typically release a considerable amount of MB dyes into natural water reservoirs, imposing health threats to beings and microbes [7]. Owing to its extreme toxicity, MB dye is dangerous to human health above a specific content level. MB is toxic, cancer-causing, and non-biodegradable, and it can put human well-being in danger and has a deleterious influence on the environment [7]. Human health dangers from MB comprise respiratory agony, gastrointestinal problems, vision, and digestive and mental illnesses [8]. Additionally, it produces indigestion, diarrhea, vomiting, cyanosis, dizziness, gastritis, jaundice, methemoglobinemia, tissue necrosis, and an escalation of heart rate, as well as untimely cell death in tissues and skin/eye irritability [7,9]. Contact with MB might cause skin redness and irritation [10]. For both aesthetic and toxicological reasons, MB liberation into surroundings poses a considerable concern. The management of wastewater containing MB dye before discharge into the environment is critical because of the negative effects on water quality and perception [11].
Likewise, major bacterial diseases are transmitted through wastewater. Therefore, microbiological management of drinking water should be the standard everywhere. Additionally, regular basic microbiological analysis of drinking water should be performed using culture methods to perceive the presence of pathogenic Escherichia coli. Hygienic water is essential for life, but an enormous number of persons are deprived of access to potable water, and many die as a result of waterborne bacterial infections [12]. To date, many water treatment technologies, including biodegradation, coagulation, adsorption, and photocatalysis, have been used to remove organic pollutants [13]. Multi-constituent photocatalysis of organic contaminants utilizing semiconducting NPs has gained popularity in recent decades because it is a low-cost, ecologically beneficial, and simple technology for treating dangerous contaminants in wastewater [4,14,15]. When compared with other ways, this technology is much more appealing due to the cheaper price of catalysts and the use of renewable energy [16].
One-dimensional (1D) nanostructured inorganic materials, especially nanofibers, are attracting a lot of attention. They were applied in various applications recently. One-dimensional metal oxides, as well as hybrids or composites [17], provide a wide range of material options, providing new options for future applications [18,19]. Electrospinning is one of the best flexible, multipurpose, and economical techniques that produce 1D nanofibers with a high aspect ratio, permeability, and adjustable penetrability. Remarkable progress was also made toward the utilization of such purposeful nanofibers in practical applications, such as fuel cells, lithium-ion arrays, solar booths, electronic sensors, photocatalysts, and supercapacitors [20]. Given the importance of renewable energy that is environmentally friendly, supercapacitors are in high demand; mostly because of their robust charge–discharge capacity, these supercapacitors are reliable [20].
Electrospinning has notable capability because of its proportionally low cost and relatively fast fabrication speed [21,22,23,24]. In recent years, cerium oxide (CeO2) has received significant attraction owing to its diverse properties, such as high-temperature stability and light absorption capability, as well as its broad range of catalytic applications [25,26,27]. It absorbs light in both the UV and visible regions; it was also applied as an effective oxidation catalyst, as it has high oxygen storage capability [28]. Moreover, it is also being used for the degradation of dyes and volatile organic compounds (VOCs) [29,30]. Currently, CeO2 is used as a photocatalyst and as an electrode material for supercapacitors [31,32]. Researchers are trying to enhance the properties of CeO2 for various applications by mixing or coupling with other semiconductors, such as CeO2/ZnFe2O4 [33], TiO2-WO3-CeO2 [34], CeO2/Ni-Al [35], CeO2/MnO2 [36], CeO2/SiO2 [37], CuO-CeO2 [38], CaO/CeO2 [39].
Tin oxide (SnO2) is an n-type semiconductor with a band gap of 3.6 eV. SnO2 has extensively been utilized as gas sensors, transparent conductive electrodes, solar cells, and photocatalysts [40,41,42]. It is expected that mixing SnO2 with CeO2 may enhance its activity for various applications. In the present study, CeO2−SnO2 composite nanofibers (NFs) were manufactured via electrospinning and screened for antibacterial, photocatalytic, and electrochemical applications. For decades, antibiotics have commonly been used to treat bacterial infections. However, the fast growth of antibiotic-resistant bacteria has caused several issues that place a significant burden on the medical community. As a result, the use of nanoparticles as an antibacterial substitute has been investigated. Regarding this situation, metal nanoparticles showed broad-spectrum antibacterial action. Furthermore, the use of nanomaterials in the biomedical area allows researchers to solve the issues of bacterial antibiotic resistance due to their distinct antibacterial mechanisms. Several metal and metal oxide-based nanomaterials were recently fully integrated into antibacterial applications and demonstrated remarkable performances [43]. Among them, Ce- and CeO2-based nanomaterials garnered a lot of attention across the world. Numerous investigations demonstrated that CeO2 nanoparticles have remarkable antibacterial activity [44,45]. Several investigations showed that CeO2 has an antibacterial impact on Staphylococcus aureus [46,47]. Furthermore, some research studies used agar diffusion and microdilution assays to investigate and confirm P. aeruginosa’s sensitivity to CeO2 [48]. In the same way, SnO2 also received attention as an antibacterial agent, where it was shown to inhibit the development of several bacterial strains, such as S. aureus and E. coli [49,50]. Furthermore, it was also found that SnO2 disinfects germs effectively once incapacitated with transition metal ions, for instance, Co@SnO2 and Ag@SnO2 nanoparticles exhibit powerful antimicrobial properties [51,52,53].
However, to the best of our knowledge, no research work has been published on the antibacterial behavior of CeO2−SnO2 composite NFs. This paper describes the facile manufacturing of CeO2−SnO2 composite NFs via electrospinning and their incredible photocatalytic, electrochemical applications, and antibacterial activity against E. coli. The composite was categorized using various physicochemical techniques, such as XRD, SEM, TEM, FT-IR, PL, and UV-vis. The photocatalytic activity was measured using MB disintegration under visible light radiation. The mechanism of improved photocatalytic activity of composite CeO2−SnO2 NFs was interpreted. The electrochemical efficiency of synthesized materials was scrutinized in terms of the cyclic voltammetry performance, which displayed that the CeO2−SnO2 composite possessed outstanding electrochemical efficacy as a supercapacitor material. Overall, the outcomes of this research highlight the innovative possibilities of utilizing these 1D high-aspect-ratio composite NFs as an electrode in energy storage systems and as a visible light active photocatalyst for the purification of chemical and biological contaminants, which would greatly assist in environmental remediation procedures. Henceforward, composites of these two materials (CeO2−SnO2) can serve as an excellent preliminary point for augmenting electrodes and photocatalytic efficiency. Additionally, this combination will be an imminent antimicrobial material with multiple functionalities. To recapitulate, the novelty of this work is that we used a facile electrospinning technique to fabricate composite NFs using cost-effective precursors. The CeO2−SnO2 composite NFs with a high aspect ratio are an auspicious future material and can be applied for antibacterial, photocatalyst, and electrode purposes, all in one.

2. Materials and Methods

2.1. Synthesis of Pure CeO2 and CeO2−SnO2 Composite NFs

Pristine CeO2 NFs were fabricated via an electrospinning process devoid of tin(II)ethylhexanoate in sol-gel. A polyvinylpyrrolidone (PVP, 15 wt%) solution was made via the usual technique by melting PVP in dimethylformamide (DMF) under magnetic stimulation for 5 h at room temperature. Ce(NO3)3·6H2O (1.0 g) was then added to the PVP solution while vigorously shaking. The resulting sol-gel was electrospun. The samples were subsequently sintered in air for 2 h at 500 °C to eliminate the polymer.
In the general process, for the synthesis of CeO2 and CeO2-SnO2 composite NFs, PVP (15 wt%) solution was primed by liquefying PVP in DMF on a magnetic stirrer as abovementioned. Tin(II) ethylhexanoate (0.15 mL) was supplemented with 2 mL of ethanol, then transferred into 10 mL of PVP solution. Ce(NO3)3·6H2O (1.0 g) was added to produce the final solution. The acquired sol-gel was transported into a 10 mL needle with a stainless steel spike. A copper pin coupled to a high voltage generator was implanted in solution as a positive terminal, while a ground iron barrel roofed with a polyethylene leaf assisted as the counter electrode. The mixture was retained as a capillary by regulating the inclination angle. A voltage of 17.5 kV was maintained to produce the final resultant assortment. The distance between the tip of the needle and the collector was held at 15 cm. Primarily, as-spun composite NFs were desiccated at 80 °C for 24 h in a vacuum. Additionally, to eliminate the polymer, composites were calcined in air at 500 °C at a rate of 2 °C/min for 2 h.

2.2. Classification of Pure CeO2 and CeO2−SnO2 NFs

The XRD outlines of pure and composite NFs were measured using a Rigaku/Max-3A (Tokyo, Japan) with Cu Kα radiation (λ = 1.540 Å) over Bragg angles between 20° to 80°. To scan surface features, pristine and composite NFs were examined using field emission scanning electron microscopy (FE-SEM, JSM6700, JEOL, Tokyo, Japan) and high-resolution transmission electron microscopy (HRTEM-H-7650 Hitachi, Co., Tokyo, Japan). The elemental conformation of the samples was scrutinized using energy-dispersive X-ray spectroscopy (EDS) coupled to an SEM instrument. The EDS of the sample was taken on carbon tape with a Pt coating. The photoluminescence (PL) spectra of the NFs were measured at room temperature using the 325 nm line of a He-Cd laser at a power of 25 mW for excitation (Kimmon Koha, JP/IK 3302 R). The light absorbance of NFs was measured by means of a UV-vis diffused reflectance spectrum (UV-DRS, 525 Shimadzu).

2.3. Photocatalytic Degradation

Photocatalytic activity of CeO2−SnO2 composite has been performed by disintegration of MB dye in visible light treatment using 450 W Xenon lamp. In 250 mL of a solution, there was a dye concentration of 10 mgL−1 and 250 mg of photocatalyst; this was stirred for 1 h in the dark (at a neutral pH). Afterward, around 3 mL of aliquots was taken at regular time intervals from the solution and centrifuged. The absorbance of MB dye mixture was analyzed by a UV–Vis spectrophotometer (Shimadzu UV-3101, UV Probe) at a 664 nm absorbance peak.

2.4. Electrochemical Characterization

Cyclic voltammetry was undertaken in a three-electrode configuration by means of a potentiostat (Digi-Ivy, USA). A glassy carbon electrode was used, while Ag/AgCl and Pt were used as the working standard and the counter electrode, respectively. The specific capacitance (Cs) was measured from CV curves calculated graphically by integrating the area under the CV curve with help of the following equation [54]:
Cs   = Va Vc IVdV   1 w   ×   υ   × ( Va     Vc )
where w is the mass of the electrode and υ is the sweep rate (V/s).

2.5. Antimicrobial Activity

The antimicrobial actions of CeO2 and CeO2−SnO2 composite NFs were screened against E. coli, which is ever-present throughout the atmosphere. The bacterial standard cultures were held on nutrient agar (NA) plates. Bacterial culture from the agar plates was seeded in 5 mL NS solution and adjusted to 1 × 106 CFU/mL. To evaluate the antibacterial effect, different concentrations of CeO2 NFs and CeO2−SnO2 composite NFs were prepared and screened. Preliminary screening was conducted using an agar diffusion technique, as described previously [55]. In short, 25 mL of agar holding 1 mL of microbial culture was placed in Petri dishes. Approximately 50 µL from each dosage was put into a 4 mm diameter well. Dishes were pre-incubated for 3 h at room temperature to enable the pre-diffusion of samples and then incubated for 24 h at 37 °C. As a negative control, DMSO was utilized, and the standard drug ciprofloxacin was used as the positive control. The inhibition potency was described as the absence of bacterial growth in the vicinity of holes and a caliper was used to determine the inhibition zone.

3. Results and Discussion

Electrospinning is a simple and versatile technology that uses electrostatic repulsion between surface charges to extract nanofibers from a viscoelastic fluid and can be used for a variety of materials [56]. In the present study, it was utilized successfully to make composite NFs with diameters as small as hundreds of nanometers. The XRD configuration of CeO2 and CeO2−SnO2 NFs after calcination at 500 °C are displayed in Figure 1. Specifically, all of the reflection crests in Figure 1a were assigned to the cubic fluorite structure of CeO2 (JCPDS no. 65-2975) at angles of 28.5°, 33.2°, 47.5°, 56.4°, and 59.1°, which relate to the (111), (200), (220), (311), and (222) crystal planes, respectively [26]. A small peak was observed at 52.5°, which was probably the peak of Ce(OH)3 [57]. In the XRD configuration of CeO2−SnO2 (Figure 1b), major diffracted crests identical to those of CeO2 NFs were visualized. Additionally, diffraction peaks from the tetragonal SnO2 (JCPDS card no. 41-1445) were also observed [58], which demonstrates the effective materialization of CeO2−SnO2 composite NFs. No other mixes of Ce and Sn were detected using XRD.
The morphologies of pure and composite NFs obtained after heating at 500 °C were demonstrated using SEM (Figure 2). From Figure 2a, it can be perceived that these arbitrarily oriented NFs had a uniform and continuous structure. Their lengths were measured as being several micrometers. The average diameter of the pure CeO2 NFs was 400 nm, whereas the average diameter of CeO2−SnO2 composite NFs was lower at 170 nm, which might have been due to the interaction between Ce and Sn (Figure 2b).
The arrows clearly show the characteristic sheath on the composite NFs, which was due to the presence of SnO2 (Figure 2c). The different salt content significantly increased the solution conductivity, resulting in thinner electrospun fibers. Due to increased spinnability in the presence of Ce and Sn precursors, uniform nanofibers were generated. Consequently, it was observed that a different salt addition had a considerable impact on the spinnability of the polymer solution. Additionally, because of the increased conductivity, the electric field accelerated the ions and the interactions between the polymer and the salt resulted in a less compact polymer structure. Our findings show that when salt (Ce and S) ions were propelled by an electric field, the polymer chain followed the ions’ motion, thereby decreasing the fiber diameter size. Analogous studies were reported for polymeric fibers by electrospinning in the presence of salts [59,60].
Figure 3a demonstrates the EDS results, which confirmed the presence of Ce, Sn, and O in the CeO2−SnO2 composite NFs. EDS analysis for the composite yielded an average atomic ratio of 1:0.15 for Ce/Sn. The detailed microstructure and morphology of composite NFs were further scanned using HR-TEM, the results of which are presented in Figure 3b. The surface consisted of two sets of lattices with spacings of 0.312 nm and 0.17 nm, corresponding to the interspace area between (111) and (211) of CeO2 and SnO2, respectively. It further verified the formation of a heterojunction between CeO2 and SnO2. The inset in Figure 3 demonstrates the low-resolution image of NFs. The TEM morphology of the CeO2−SnO2 NFs again showed that the average diameter of the composite NFs was 170 nm (inset of Figure 3). The corresponding selected area electron diffractions (SAED) arrangement of these NFs displayed a good ring pattern with no dislocations or imperfections, which can be attributed to the high crystalline phase of the composite sample, indicating the polycrystalline nature of the composite NFs (inset of Figure 3).
Figure 4 displays the FT-IR bands of the virgin CeO2 and composite CeO2−SnO2 NFs. The existence of a strong band prior to 600 cm−1 could be assigned to the Ce-O-Ce stretching vibration [61] (Figure 4a). Both spectra revealed bands around 3450 cm−1 and 1637 cm−1 conforming to O–H widening vibration of the remaining water molecules and hydroxyl clusters. In the nanocomposite NFs spectra (Figure 4b), the occurrence of a band at 610 cm−1 corresponded to the stretching mode of SnO2 [62]. The presence of both CeO2 and SnO2 peaks suggested the integration of these materials in the final composite NFs.
The optical characteristics of bare and composite NFs were assessed using UV-DRS (Figure 5). Figure 5A demonstrates the UV-DRS results of the CeO2 and CeO2−SnO2 NFs, which show the absorption edge within the visible region. The spectrum of the CeO2-SnO2 composite presented a slight red shift, which coordinated well with pure CeO2 and could be ascribed to exciton partial leakage into the CeO2 matrix. To estimate the band gaps of the prepared samples, the modified Kubelka–Munk function was plotted for (Ahν)2 versus the energy of the exciting light (hν) using the following formula:
  α = Constant   ( h ν     E g ) n h ν
From the inset in Figure 5A, the calculated band gaps of the CeO2 NFs and CeO2−SnO2 composite were found to be 2.87 eV and 2.67 eV, respectively. This means that the smaller band gap of CeO2−SnO2 composite NFs will harness more visible light than CeO2 NFs. These results are consistent with an earlier report [63]. The PL emission spectra of the CeO2 and CeO2−SnO2 composite samples were documented using an excitation wavelength (Figure 5B). The PL spectrum of CeO2 had prominent peaks at 364, 421, 483, and 530 nm [64,65,66]. Moreover, the CeO2−SnO2 composite showed weaker emission peaks than CeO2, indicating a lower recombination frequency of the photogenerated electron–hole pair.
The photoactivity of pure and composite nanofibers was assessed for the breakdown of MB dye under visible light (Figure 6). Figure 6A shows the MB concentration change calculated from its main absorption peak at 664 nm. From a comparative experiment, it can clearly be seen that CeO2−SnO2 composite photocatalyst displayed better photoactivity than the pure CeO2 NFs (Figure 6B). After 125 min of visible light irradiation, the photodegradation effectivenesses of the CeO2 and the CeO2−SnO2 composite photocatalysts for MB were 42% and 85%, respectively. The better photodegradation efficiency of the CeO2-SnO2 composite can be ascribed to the development of heterojunction between CeO2 and SnO2 and the parting of photogenerated electron–hole pairs.
Figure 7 shows the reusability capacity of CeO2−SnO2 composite NFs for the degradation of MB dye. It was demonstrated that even after five cycles, the CeO2−SnO2 composite still exhibited good photocatalytic efficiency. It can be perceived that as the number of cycles increased, the composite photocatalytic activity marginally declined. The CeO2−SnO2 catalyst degraded 85% of the MB dye solution on its initial application, but after repeated applications (5th cycle), it had achieved a decolorization efficiency of 81%.
This shows that the CeO2−SnO2 catalyst was a reliable and efficient composite for dye degradation. Erstwhile, many similar studies also reported on composite nanomaterials for MB dye degradation (Table 1).
A potential mechanism for photocatalytic MB degradation using a CeO2−SnO2 composite is hypothesized and illustrated in Figure 8. Under visible light irradiation, photons are absorbed by the CeO2 catalyst when energy is equivalent to or greater than the band gap. CeO2 gets excited and generates electrons (e) and holes (h+).
CeO2 + hν → (eCB)CeO2 + (eVB+)CeO2
The photoexcited electrons are transported to the SnO2, as the conduction band (CB) position of CeO2 is more negative than SnO2. Meanwhile, the transported electrons in the CB of SnO2 will interact with dissolved oxygen molecules to generate superoxide radical anions (O2). The superoxide radicals react with water and produce hydroperoxide radicals (HO2) and hydroxyl radicals (OH). Subsequently, the holes on the valence band (VB) of SnO2 migrate to the VB of the CeO2 since the valence band potential of CeO2 is more negative than that of SnO2. Simultaneously, the generated h+ can oxidize water molecules to produce OH. These radicals have a strong ability to degrade dye molecules (Figure 8). The MB dye is photo-oxidized into carbon dioxide and water molecules or inorganic ions as degradation products [63].
O2 + (eCB)SnO2 → O2
O2 + H2O → HO2 + OH
h+ + H2O → OH + H+
MB dye + Radical species (O2, OH) → Degradation products (CO2 + H2O)
Figure 9A demonstrates the cyclic voltammograms (CVs) of the pure CeO2 NFs and CeO2−SnO2 NFs. The CeO2−SnO2 composite NFs indicated higher current values and better voltammetric shapes compared with pure CeO2. The CeO2−SnO2 composite exhibited improved specific capacitance (413.9 Fg−1) compared with the CeO2 NFs (263.9 Fg−1) at a scan rate of 10 mV/s. No redox peaks were observed in the CV of the bare CeO2 NFs, which might have been due to the reason that there was little or no oxidation–reduction (corrosion or degradation) process going parallel to the double-layer charging. The presence of redox peaks in the CeO2−SnO2 composite NFs indicates that reversible and incessant faradaic oxidation–reduction responses were involved during the charging and discharging procedure [75,76]. It can be assumed that the electrochemical performances of composites were enhanced due to the synergistic effect between CeO2 and SnO2. Figure 9B shows the CV curves for different scan rates for the CeO2−SnO2 composite. The CeO2−SnO2 composite NFs confirmed elevated specific capacitance values of 413.9, 294.94, 147.6, and 16.1 F/g at scan rates of 5, 10, 50, and 100 mVs−1, correspondingly.
In the antibacterial bioassay, CeO2 NFs and CeO2−SnO2 composite NFs were tested against specified bacteria using an agar diffusion experiment at various doses. The CeO2 and CeO2−SnO2 composite NFs were found to be the most promising, with MICs of 50 µg/mL for E. coli (Figure 10). The augmented activity of the CeO2−SnO2 composite NFs might be attributed to their form and wide surface area [77], as well as the synergistic influence of CeO2 and SnO2 in the composite NFs. However, the antibacterial activities of the virgin CeO2 NFs and CeO2−SnO2 composite NFs were identical at a low concentration, the antibacterial impact varied with the increase in concentration and incubation time. The chemical components of the CeO2−SnO2 composite NFs primarily reacted with the bacterial outer cell wall before diffusing inside the inner wall, producing disarray and leaking through the disturbance in the inner cell content and distortion. A few previous investigations also showed that CeO2 nanoparticles have exceptional antibacterial activity [44,45]. In particular, some studies found that CeO2 has an antibacterial effect on S. aureus [46,47]. Furthermore, agar diffusion and microdilution experiments were utilized in certain studies to examine and validate P. aeruginosa’s sensitivity to CeO2 [48]. Similarly, SnO2 also attracted interest as an effective antibacterial, where it was shown to inhibit the development of several bacterial strains, such as S. aureus and E. coli [49,50]. Furthermore, it was observed that SnO2 supplemented with transition metal ions effectively disinfects microorganisms and possesses potent antibacterial activities [51,52,53].

4. Conclusions

In summary, CeO2−SnO2 composite NFs were prepared via electrospinning. As aforementioned, many strategies for removing MB and other textile dyes from industrial effluent were reported in the literature. However, in the present study, the photocatalysis approach was used. The addition of SnO2 to CeO2 significantly improved the photocatalytic MB degradation property and photoelectrochemical performance. After 125 min of irradiation, a photodegradation efficiency of 85% was achieved using the CeO2−SnO2 composite photocatalyst. Moreover, the CeO2−SnO2 composite NFs also showed higher specific capacitance. It is believed that these CeO2−SnO2 composite NFs with excellent photoactivity and electrochemical performance can be very promising for high-performance electrode materials in supercapacitors and as an efficient alternative photocatalyst. Rationally, due to their distinctive functional mechanism towards pathogens via reversible oxidation state transition between Ce(III) and Ce(IV), (IV) CeO2 nanoparticles with lower toxicity work as powerful antibacterial agents as well. The outcome of this study indicates the prospects of CeO2−SnO2 composite NFs to be exploited as an antimicrobial material, an electrode, and as a visible light active photocatalyst, which would greatly aid in the environmental cleanup process.

Author Contributions

Conceptualization, J.S.A., M.S.H. and T.A.; methodology, J.S.A. and M.S.H.; software, M.A.M.A.; validation, A.F.S. and M.F.; formal analysis, H.Y.K.; investigation, T.A.; resources, F.A.S.; data curation, A.F.S. and M.F.; writing—original draft preparation, M.S.H., T.A. and J.S.A.; writing—review and editing, T.A. and F.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are thankful to the Deanship of Scientific Research at Najran University for funding this work under the General Research Funding program “Functional Composite Materials for Water Remediation”, grant code (NU/DRP/SERC/12/45).

Data Availability Statement

Data is included in the main text.

Acknowledgments

The authors are thankful to the Deanship of Scientific Research at Najran University for funding this work under the General Research Funding program, grant code (NU/DRP/SERC/12/45).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guo, W.; Luo, H.; Jiang, Z.; Fang, D.; Chi, J.; Shangguan, W.; Wang, Z.; Wang, L.; Lee, A.F. Ge-doped cobalt oxide for electrocatalytic and photocatalytic water splitting. ACS Catal. 2022, 12, 12000–12013. [Google Scholar] [CrossRef]
  2. Hunge, Y.M.; Yadav, A.A.; Kang, S.-W.; Lim, S.J.; Kim, H. Visible light activated MoS2/ZnO composites for photocatalytic degradation of ciprofloxacin antibiotic and hydrogen production. J. Photochem. Photobiol. A Chem. 2023, 434, 114250. [Google Scholar] [CrossRef]
  3. Das, T.K.; Das, N.C. Advances on catalytic reduction of 4-nitrophenol by nanostructured materials as benchmark reaction. Int. Nano Lett. 2022, 12, 223–242. [Google Scholar] [CrossRef]
  4. Yadav, A.A.; Hunge, Y.M.; Kang, S.-W.; Fujishima, A.; Terashima, C. Enhanced Photocatalytic Degradation Activity Using the V2O5/RGO Composite. Nanomaterials 2023, 13, 338. [Google Scholar] [CrossRef]
  5. Fong, W.M.; Affam, A.C.; Chung, W.C. Synthesis of Ag/Fe/CAC for colour and COD removal from methylene blue dye wastewater. Int. J. Environ. Sci. Technol. 2020, 17, 3485–3494. [Google Scholar] [CrossRef]
  6. Han, T.H.; Khan, M.M.; Kalathil, S.; Lee, J.; Cho, M.H. Simultaneous enhancement of methylene blue degradation and power generation in a microbial fuel cell by gold nanoparticles. Ind. Eng. Chem. Res. 2013, 52, 8174–8181. [Google Scholar] [CrossRef]
  7. Khan, I.; Saeed, K.; Zekker, I.; Zhang, B.; Hendi, A.H.; Ahmad, A.; Ahmad, S.; Zada, N.; Ahmad, H.; Shah, L.A. Review on methylene blue: Its properties, uses, toxicity and photodegradation. Water 2022, 14, 242. [Google Scholar] [CrossRef]
  8. Abdelrahman, E.A.; Hegazey, R.; El-Azabawy, R.E. Efficient removal of methylene blue dye from aqueous media using Fe/Si, Cr/Si, Ni/Si, and Zn/Si amorphous novel adsorbents. J. Mater. Res. Technol. 2019, 8, 5301–5313. [Google Scholar] [CrossRef]
  9. Jawad, A.H.; Abdulhameed, A.S.; Mastuli, M.S. Acid-factionalized biomass material for methylene blue dye removal: A comprehensive adsorption and mechanism study. J. Taibah Univ. Sci. 2020, 14, 305–313. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, Y.; Peng, Q.; Akhtar, N.; Chen, X.; Huang, Y. Microporous carbon material from fish waste for removal of methylene blue from wastewater. Water Sci. Technol. 2020, 81, 1180–1190. [Google Scholar] [CrossRef]
  11. Zamel, D.; Hassanin, A.H.; Ellethy, R.; Singer, G.; Abdelmoneim, A. Novel bacteria-immobilized cellulose acetate/poly (ethylene oxide) nanofibrous membrane for wastewater treatment. Sci. Rep. 2019, 9, 18994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Cabral, J.P. Water microbiology. Bacterial pathogens and water. Int. J. Environ. Res. Public Health 2010, 7, 3657–3703. [Google Scholar] [CrossRef] [PubMed]
  13. Hunge, Y.; Yadav, A.; Kang, S.-W.; Kim, H. Facile synthesis of multitasking composite of Silver nanoparticle with Zinc oxide for 4-nitrophenol reduction, photocatalytic hydrogen production, and 4-chlorophenol degradation. J. Alloys Compd. 2022, 928, 167133. [Google Scholar] [CrossRef]
  14. Chandra, R.; Nath, M. Controlled synthesis of AgNPs@ ZIF-8 composite: Efficient heterogeneous photocatalyst for degradation of methylene blue and congo red. J. Water Process Eng. 2020, 36, 101266. [Google Scholar] [CrossRef]
  15. Hunge, Y.M.; Yadav, A.A.; Kang, S.-W.; Mohite, B.M. Role of Nanotechnology in Photocatalysis Application. Recent Pat. Nanotechnol. 2023, 17, 5–7. [Google Scholar] [CrossRef] [PubMed]
  16. Muhd Julkapli, N.; Bagheri, S.; Bee Abd Hamid, S. Recent advances in heterogeneous photocatalytic decolorization of synthetic dyes. Sci. World J. 2014, 2014, 692307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Alorabi, A.Q.; Hassan, M.S.; Algethami, J.S.; Baghdadi, N.E. Synthesis and characterization of Ag-AgVO3/Cu2O heterostructure with improved visible-light photocatalytic performance. Sci. Prog. 2021, 104, 00368504211050300. [Google Scholar] [CrossRef]
  18. Zhang, H.; Zhang, F.; Wei, Y.; Miao, Q.; Li, A.; Zhao, Y.; Yuan, Y.; Jin, N.; Li, G. Controllable Design and Preparation of Hollow Carbon-Based Nanotubes for Asymmetric Supercapacitors and Capacitive Deionization. ACS Appl. Mater. Interfaces 2021, 13, 21217–21230. [Google Scholar] [CrossRef]
  19. Han, Y.; Sun, C.; Gao, K.; Ding, S.; Miao, Z.; Zhao, J.; Yang, Z.; Wu, P.; Huang, J.; Li, Z. Heterovalent oxynitride GaZnON nanowire as novel flexible anode for lithium-ion storage. Electrochim. Acta 2022, 408, 139931. [Google Scholar] [CrossRef]
  20. Hassan, M.S.; Amna, T.; Alqarni, L.S.; Alqahtani, H.S.; Alnaam, Y.A.; Almusabi, S.; Alzharani, A.A. High aspect ratio TiO2–Mn3O4 heterostructure: Proficient nanorods for pathogen inhibition and supercapacitor application. Mater. Sci. Technol. 2023, 1–10. [Google Scholar] [CrossRef]
  21. Gao, S.; Jiang, J.; Li, X.; Ye, F.; Fu, Y.; Zhao, L. Electrospun polymer-free nanofibers incorporating hydroxypropyl-β-cyclodextrin/Difenoconazole via supramolecular assembly for antifungal activity. J. Agric. Food Chem. 2021, 69, 5871–5881. [Google Scholar] [CrossRef] [PubMed]
  22. Ding, D.; Li, Z.; Yu, S.; Yang, B.; Yin, Y.; Zan, L.; Myung, N.V. Piezo-photocatalytic flexible PAN/TiO2 composite nanofibers for environmental remediation. Sci. Total Environ. 2022, 824, 153790. [Google Scholar] [CrossRef]
  23. Han, W.-H.; Li, X.; Yu, G.-F.; Wang, B.-C.; Huang, L.-P.; Wang, J.; Long, Y.-Z. Recent Advances in the Food Application of Electrospun Nanofibers. J. Ind. Eng. Chem. 2022, 110, 15–26. [Google Scholar] [CrossRef]
  24. Xue, C.; Liu, Y.; Zhao, J.; Li, X.; Zhang, J.; Zhang, J. Nitrogen-doped carbon nanofibers derived from phenolic-resin-based analogues for high-performance lithium-ion batteries. Solid State Ion. 2022, 376, 115854. [Google Scholar] [CrossRef]
  25. Payan, A.; Aghdam, N.C.; Soltan, J. Catalytic oxidation of acetone in air over Ag modified CeO2 catalysts under VUV irradiation: Comparison of different treatment process, performance, and mechanism studies. J. Environ. Chem. Eng. 2022, 10, 107253. [Google Scholar] [CrossRef]
  26. Hassan, M.S.; Khan, R.; Amna, T.; Yang, J.; Lee, I.-H.; Sun, M.-Y.; EL-Newehy, M.H.; Al-Deyab, S.S.; Khil, M.-S. The influence of synthesis method on size and toxicity of CeO2 quantum dots: Potential in the environmental remediation. Ceram. Int. 2016, 42, 576–582. [Google Scholar] [CrossRef]
  27. Das, T.K.; Remanan, S.; Ghosh, S.; Ghosh, S.K.; Das, N.C. Efficient synthesis of catalytic active silver nanoparticles illuminated cerium oxide nanotube: A mussel inspired approach. Environ. Nanotechnol. Monit. Manag. 2021, 15, 100411. [Google Scholar] [CrossRef]
  28. Dong, F.; Tanabe, T.; Takahashi, N.; Shinjoh, H. Investigation of the effective oxygen storage and release performances on the Pt/CeO2-ZrO2 catalysts by breakthrough method. Catal. Today 2019, 332, 259–266. [Google Scholar] [CrossRef]
  29. Chu, Z.; Li, J.; Lan, Y.-P.; Chen, C.; Yang, J.; Ning, D.; Xia, X.; Mao, X. KCl–LiCl molten salt synthesis of LaOCl/CeO2-g-C3N4 with excellent photocatalytic-adsorbed removal performance for organic dye pollutant. Ceram. Int. 2022, 48, 15439–15450. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Wu, M.; Wang, Y.; Zhao, X.; Leung, D.Y. Low-cost and efficient Mn/CeO2 catalyst for photocatalytic VOCs degradation via scalable colloidal solution combustion synthesis method. J. Mater. Sci. Technol. 2022, 116, 169–179. [Google Scholar] [CrossRef]
  31. Liu, Z.; Zheng, J.; Duan, L.; Zhu, Z. Biomass-assisted synthesis of CeO2 nanorods for CO2 photoreduction under visible light. ACS Appl. Nano Mater. 2021, 4, 4226–4237. [Google Scholar] [CrossRef]
  32. Xue, Z.; Lv, L.; Tian, Y.; Tan, S.; Ma, Q.; Tao, K.; Han, L. Co3S4 nanoplate arrays decorated with oxygen-deficient CeO2 nanoparticles for supercapacitor applications. ACS Appl. Nano Mater. 2021, 4, 3033–3043. [Google Scholar] [CrossRef]
  33. AL-Shwaiman, H.A.; Akshhayya, C.; Syed, A.; Bahkali, A.H.; Elgorban, A.M.; Das, A.; Varma, R.S.; Khan, S.S. Fabrication of intimately coupled CeO2/ZnFe2O4 nano-heterojunction for visible-light photocatalysis and bactericidal application. Mater. Chem. Phys. 2022, 279, 125759. [Google Scholar] [CrossRef]
  34. Munawar, T.; Mukhtar, F.; Nadeem, M.S.; Manzoor, S.; Ashiq, M.N.; Mahmood, K.; Batool, S.; Hasan, M.; Iqbal, F. Fabrication of dual Z-scheme TiO2-WO3-CeO2 heterostructured nanocomposite with enhanced photocatalysis, antibacterial, and electrochemical performance. J. Alloys Compd. 2021, 898, 162779. [Google Scholar] [CrossRef]
  35. Zheng, J.; Wang, J.; Qu, J.; Wu, M.; Xu, Z. CeO2/Ni-Al layered double hydroxide composite electrode for the enhancement of specific capacitance and capacitance retention performance. Appl. Clay Sci. 2022, 216, 106370. [Google Scholar] [CrossRef]
  36. Xie, A.; Wang, H.; Zhu, Z.; Zhang, W.; Li, X.; Wang, Q.; Luo, S. Mesoporous CeO2-α-MnO2-reduced graphene oxide composite with ultra-high stability as a novel electrode material for supercapacitor. Surf. Interfaces 2021, 25, 101177. [Google Scholar] [CrossRef]
  37. Yin, L.; Xu, J.; Zhang, B.; Wang, L.; Tao, W.; Teng, X.; Ning, W.; Zhang, Z. A facile fabrication of highly dispersed CeO2/SiO2 aerogel composites with high adsorption desulfurization performance. Chem. Eng. J. 2022, 428, 132581. [Google Scholar] [CrossRef]
  38. Lu, Y.; Duan, L.; Sun, Z.; Chen, J. Flame spray pyrolysis synthesized CuO–CeO2 composite for catalytic combustion of C3H6. Proc. Combust. Inst. 2021, 38, 6513–6520. [Google Scholar] [CrossRef]
  39. Thirunavukkarasu, A.; Nithya, R. Adsorption of acid orange 7 using green synthesized CaO/CeO2 composite: An insight into kinetics, equilibrium, thermodynamics, mass transfer and statistical models. J. Taiwan Inst. Chem. Eng. 2020, 111, 44–62. [Google Scholar] [CrossRef]
  40. Van, K.N.; Huu, H.T.; Thi, V.N.N.; Le Thi, T.L.; Truong, D.H.; Truong, T.T.; Dao, N.N.; Vo, V.; Vasseghian, Y. Facile construction of S-scheme SnO2/g-C3N4 photocatalyst for improved photoactivity. Chemosphere 2022, 289, 133120. [Google Scholar] [CrossRef]
  41. Panday, M.; Upadhyay, G.K.; Purohit, L. Sb incorporated SnO2 nanostructured thin films for CO2 gas sensing and humidity sensing applications. J. Alloys Compd. 2022, 904, 164053. [Google Scholar] [CrossRef]
  42. Qureshi, A.A.; Javed, S.; Javed, H.M.A.; Akram, A.; Mustafa, M.S.; Ali, U.; Nisar, M.Z. Facile formation of SnO2–TiO2 based photoanode and Fe3O4@ rGO based counter electrode for efficient dye-sensitized solar cells. Mater. Sci. Semicond. Process. 2021, 123, 105545. [Google Scholar] [CrossRef]
  43. Algethami, J.S.; Hassan, M.S.; Alorabi, A.Q.; Alhemiary, N.A.; Fallatah, A.M.; Alnaam, Y.; Almusabi, S.; Amna, T. Manganese Ferrite–Hydroxyapatite Nanocomposite Synthesis: Biogenic Waste Remodeling for Water Decontamination. Nanomaterials 2022, 12, 1631. [Google Scholar] [CrossRef] [PubMed]
  44. Farias, I.A.P.; Santos, C.C.L.d.; Sampaio, F.C. Antimicrobial activity of cerium oxide nanoparticles on opportunistic microorganisms: A systematic review. BioMed. Res. Int. 2018, 2018, 1923606. [Google Scholar] [CrossRef] [Green Version]
  45. Pop, O.L.; Mesaros, A.; Vodnar, D.C.; Suharoschi, R.; Tăbăran, F.; Magerușan, L.; Tódor, I.S.; Diaconeasa, Z.; Balint, A.; Ciontea, L. Cerium oxide nanoparticles and their efficient antibacterial application in vitro against gram-positive and gram-negative pathogens. Nanomaterials 2020, 10, 1614. [Google Scholar] [CrossRef] [PubMed]
  46. Reddy Yadav, L.; Manjunath, K.; Archana, B.; Madhu, C.; Raja Naika, H.; Nagabhushana, H.; Kavitha, C.; Nagaraju, G. Fruit juice extract mediated synthesis of CeO2 nanoparticles for antibacterial and photocatalytic activities. Eur. Phys. J. Plus 2016, 131, 1–10. [Google Scholar] [CrossRef]
  47. Gopinath, K.; Karthika, V.; Sundaravadivelan, C.; Gowri, S.; Arumugam, A. Mycogenesis of cerium oxide nanoparticles using Aspergillus niger culture filtrate and their applications for antibacterial and larvicidal activities. J. Nanostructure Chem. 2015, 5, 295–303. [Google Scholar] [CrossRef] [Green Version]
  48. Ravishankar, T.N.; Ramakrishnappa, T.; Nagaraju, G.; Rajanaika, H. Synthesis and characterization of CeO2 nanoparticles via solution combustion method for photocatalytic and antibacterial activity studies. ChemistryOpen 2015, 4, 146–154. [Google Scholar] [CrossRef]
  49. Phukan, A.; Bhattacharjee, R.P.; Dutta, D.K. Stabilization of SnO2 nanoparticles into the nanopores of modified Montmorillonite and their antibacterial activity. Adv. Powder Technol. 2017, 28, 139–145. [Google Scholar] [CrossRef]
  50. Kumari, M.M.; Philip, D. Synthesis of biogenic SnO2 nanoparticles and evaluation of thermal, rheological, antibacterial and antioxidant activities. Powder Technol. 2015, 270, 312–319. [Google Scholar] [CrossRef]
  51. Chandran, D.; Nair, L.S.; Balachandran, S.; Rajendra Babu, K.; Deepa, M. Structural, optical, photocatalytic, and antimicrobial activities of cobalt-doped tin oxide nanoparticles. J. Sol-Gel Sci. Technol. 2015, 76, 582–591. [Google Scholar] [CrossRef]
  52. Nasir, Z.; Shakir, M.; Wahab, R.; Shoeb, M.; Alam, P.; Khan, R.H.; Mobin, M. Co-precipitation synthesis and characterization of Co doped SnO2 NPs, HSA interaction via various spectroscopic techniques and their antimicrobial and photocatalytic activities. Int. J. Biol. Macromol. 2017, 94, 554–565. [Google Scholar] [CrossRef] [PubMed]
  53. Qamar, M.A.; Shahid, S.; Khan, S.A.; Zaman, S.; Sarwar, M.N. Synthesis characterization, optical and antibacterial studies of Co-doped SnO2 nanoparticles. Dig. J. Nanomater. Biostruct. 2017, 12, 1127–1135. [Google Scholar]
  54. Hassan, M.S.; Amna, T.; Yang, O.-B.; El-Newehy, M.H.; Al-Deyab, S.S.; Khil, M.-S. Smart copper oxide nanocrystals: Synthesis, characterization, electrochemical and potent antibacterial activity. Colloids Surf. B Biointerfaces 2012, 97, 201–206. [Google Scholar] [CrossRef] [PubMed]
  55. Jorgensen, J.H.; Hindler, J.F.; Reller, L.B.; Weinstein, M.P. New consensus guidelines from the Clinical and Laboratory Standards Institute for antimicrobial susceptibility testing of infrequently isolated or fastidious bacteria. Clin. Infect. Dis. 2007, 44, 280–286. [Google Scholar] [CrossRef] [Green Version]
  56. Xue, J.; Xie, J.; Liu, W.; Xia, Y. Electrospun nanofibers: New concepts, materials, and applications. Acc. Chem. Res. 2017, 50, 1976–1987. [Google Scholar] [CrossRef]
  57. Nilchi, A.; Yaftian, M.; Aboulhasanlo, G.; Rasouli Garmarodi, S. Adsorption of selected ions on hydrous cerium oxide. J. Radioanal. Nucl. Chem. 2009, 279, 65–74. [Google Scholar] [CrossRef]
  58. Bao, H.; Zhang, Z.; Hua, Q.; Huang, W. Compositions, structures, and catalytic activities of CeO2@ Cu2O nanocomposites prepared by the template-assisted method. Langmuir 2014, 30, 6427–6436. [Google Scholar] [CrossRef]
  59. Wang, B.-B.; Wang, X.-D.; Wang, T.-H. Microscopic mechanism for the effect of adding salt on electrospinning by molecular dynamics simulations. Appl. Phys. Lett. 2014, 105, 121906. [Google Scholar] [CrossRef]
  60. Topuz, F.; Abdulhamid, M.A.; Holtzl, T.; Szekely, G. Nanofiber engineering of microporous polyimides through electrospinning: Influence of electrospinning parameters and salt addition. Mater. Des. 2021, 198, 109280. [Google Scholar] [CrossRef]
  61. Ho, C.; Yu, J.C.; Kwong, T.; Mak, A.C.; Lai, S. Morphology-controllable synthesis of mesoporous CeO2 nano- and microstructures. Chem. Mater. 2005, 17, 4514–4522. [Google Scholar] [CrossRef]
  62. Wan, W.; Li, Y.; Ren, X.; Zhao, Y.; Gao, F.; Zhao, H. 2D SnO2 nanosheets: Synthesis, characterization, structures, and excellent sensing performance to ethylene glycol. Nanomaterials 2018, 8, 112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Manibalan, G.; Murugadoss, G.; Thangamuthu, R.; Kumar, R.M.; Jayavel, R.; Kumar, M.R. Enhanced photocatalytic performance of heterostructure CeO2–SnO2 nanocomposite via hydrothermal route. Mater. Res. Express 2019, 6, 075032. [Google Scholar] [CrossRef]
  64. Subhan, A.; Ahmed, T.; Awal, M.R.; Kim, B.M. Structure and photoluminescence studies of CeO2 CuAlO2 mixed metal oxide fabricated by co-precipitation method. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 135, 466–471. [Google Scholar] [CrossRef] [PubMed]
  65. Malleshappa, J.; Nagabhushana, H.; Prasad, B.D.; Sharma, S.; Vidya, Y.; Anantharaju, K. Structural, photoluminescence and thermoluminescence properties of CeO2 nanoparticles. Optik 2016, 127, 855–861. [Google Scholar] [CrossRef]
  66. Kumar, S.; Ojha, A.K.; Patrice, D.; Yadav, B.S.; Materny, A. One-step in situ synthesis of CeO2 nanoparticles grown on reduced graphene oxide as an excellent fluorescent and photocatalyst material under sunlight irradiation. Phys. Chem. Chem. Phys. 2016, 18, 11157–11167. [Google Scholar] [CrossRef]
  67. Chae, B.W.; Amna, T.; Hassan, M.S.; Al-Deyab, S.S.; Khil, M.-S. CeO2-Cu2O composite nanofibers: Synthesis, characterization photocatalytic and electrochemical application. Adv. Powder Technol. 2017, 28, 230–235. [Google Scholar] [CrossRef]
  68. Qiu, J.; Liu, F.; Yue, C.; Ling, C.; Li, A. A recyclable nanosheet of Mo/N-doped TiO2 nanorods decorated on carbon nanofibers for organic pollutants degradation under simulated sunlight irradiation. Chemosphere 2019, 215, 280–293. [Google Scholar] [CrossRef]
  69. Jeong, W.-H.; Amna, T.; Ha, Y.-M.; Hassan, M.S.; Kim, H.-C.; Khil, M.-S. Novel PANI nanotube@ TiO2 composite as efficient chemical and biological disinfectant. Chem. Eng. J. 2014, 246, 204–210. [Google Scholar] [CrossRef]
  70. Gadisa, B.T.; Appiah-Ntiamoah, R.; Kim, H. In-situ derived hierarchical ZnO/Zn-C nanofiber with high photocatalytic activity and recyclability under solar light. Appl. Surf. Sci. 2019, 491, 350–359. [Google Scholar] [CrossRef]
  71. Yasin, A.S.; Obaid, M.; El-Newehy, M.H.; Al-Deyab, S.S.; Barakat, N.A. Influence of TixZr(1−x)O2 nanofibers composition on the photocatalytic activity toward organic pollutants degradation and water splitting. Ceram. Int. 2015, 41, 11876–11885. [Google Scholar] [CrossRef]
  72. Wang, H.; Huang, X.; Li, W.; Gao, J.; Xue, H.; Li, R.K.; Mai, Y.-W. TiO2 nanoparticle decorated carbon nanofibers for removal of organic dyes. Colloids Surf. A Physicochem. Eng. Asp. 2018, 549, 205–211. [Google Scholar] [CrossRef]
  73. Zhao, J.; Lu, Q.; Wang, Q.; Ma, Q. α-Fe2O3 nanoparticles on Bi2MoO6 nanofibers: One-dimensional heterostructures synergistic system with enhanced photocatalytic activity. Superlattices Microstruct. 2016, 91, 148–157. [Google Scholar] [CrossRef]
  74. Yousef, A.; Barakat, N.A.; Amna, T.; Unnithan, A.R.; Al-Deyab, S.S.; Kim, H.Y. Influence of CdO-doping on the photoluminescence properties of ZnO nanofibers: Effective visible light photocatalyst for waste water treatment. J. Lumin. 2012, 132, 1668–1677. [Google Scholar] [CrossRef]
  75. Zhou, F.; Liu, Q.; Kang, D.; Gu, J.; Zhang, W.; Zhang, D. A 3D hierarchical hybrid nanostructure of carbon nanotubes and activated carbon for high-performance supercapacitors. J. Mater. Chem. A 2014, 2, 3505–3512. [Google Scholar] [CrossRef]
  76. Priyadharsan, A.; Vasanthakumar, V.; Karthikeyan, S.; Raj, V.; Shanavas, S.; Anbarasan, P.M. Multi-functional properties of ternary CeO2/SnO2/rGO nanocomposites: Visible light driven photocatalyst and heavy metal removal. J. Photochem. Photobiol. A Chem. 2017, 346, 32–45. [Google Scholar] [CrossRef]
  77. Ann, L.C.; Mahmud, S.; Bakhori, S.K.M.; Sirelkhatim, A.; Mohamad, D.; Hasan, H.; Seeni, A.; Rahman, R.A. Antibacterial responses of zinc oxide structures against Staphylococcus aureus, Pseudomonas aeruginosa and Streptococcus pyogenes. Ceram. Int. 2014, 40, 2993–3001. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) CeO2 and (b) CeO2−SnO2 composite NFs calcined at 500 °C.
Figure 1. XRD patterns of (a) CeO2 and (b) CeO2−SnO2 composite NFs calcined at 500 °C.
Nanomaterials 13 01001 g001
Figure 2. SEM profiles of (a) CeO2 and (b,c) CeO2−SnO2 nanofibers at low and high magnifications after calcination at 500 °C. The arrows clearly show the characteristic sheath on composite NFs, which was absent for the bare CeO2 NFs.
Figure 2. SEM profiles of (a) CeO2 and (b,c) CeO2−SnO2 nanofibers at low and high magnifications after calcination at 500 °C. The arrows clearly show the characteristic sheath on composite NFs, which was absent for the bare CeO2 NFs.
Nanomaterials 13 01001 g002
Figure 3. (a) EDX spectra and (b) HR-TEM image of the prepared CeO2−SnO2 composite NFs (the inset shows a low-resolution image and SAED pattern).
Figure 3. (a) EDX spectra and (b) HR-TEM image of the prepared CeO2−SnO2 composite NFs (the inset shows a low-resolution image and SAED pattern).
Nanomaterials 13 01001 g003
Figure 4. FT-IR configurations of the (a) CeO2 and (b) CeO2−SnO2 NFs.
Figure 4. FT-IR configurations of the (a) CeO2 and (b) CeO2−SnO2 NFs.
Nanomaterials 13 01001 g004
Figure 5. (A) UV-DRS pattern (inset: plot of the transformed Kubelka–Munk function versus the energy band gap) and (B) PL spectra of (a) CeO2 and (b) CeO2−SnO2 composite NFs.
Figure 5. (A) UV-DRS pattern (inset: plot of the transformed Kubelka–Munk function versus the energy band gap) and (B) PL spectra of (a) CeO2 and (b) CeO2−SnO2 composite NFs.
Nanomaterials 13 01001 g005
Figure 6. (A) Absorption spectra of MB dye using CeO2-SnO2 composite NFs for different periods and (B) relative photocatalytic action of (a) CeO2 and (b) a CeO2-SnO2 nanocomposite for MB dye.
Figure 6. (A) Absorption spectra of MB dye using CeO2-SnO2 composite NFs for different periods and (B) relative photocatalytic action of (a) CeO2 and (b) a CeO2-SnO2 nanocomposite for MB dye.
Nanomaterials 13 01001 g006
Figure 7. Recyclability studies of CeO2−SnO2 composite NFs for five consecutive cycles. Different color indicates different cycles
Figure 7. Recyclability studies of CeO2−SnO2 composite NFs for five consecutive cycles. Different color indicates different cycles
Nanomaterials 13 01001 g007
Figure 8. Schematic representation of the energy band configuration and electron−hole pair separation in CeO2−SnO2 NFs.
Figure 8. Schematic representation of the energy band configuration and electron−hole pair separation in CeO2−SnO2 NFs.
Nanomaterials 13 01001 g008
Figure 9. (a) Cyclic voltammograms of the CeO2 and CeO2−SnO2 NFs electrodes at a scanning rate of 5 mV/s and (b) CVs of CeO2−SnO2 NFs at different scanning rates.
Figure 9. (a) Cyclic voltammograms of the CeO2 and CeO2−SnO2 NFs electrodes at a scanning rate of 5 mV/s and (b) CVs of CeO2−SnO2 NFs at different scanning rates.
Nanomaterials 13 01001 g009
Figure 10. (a) E. coli susceptibility to CeO2 NFs and CeO2−SnO2 composite NFs (50–400 µg/mL). Ciprofloxacin was used as the standard antibiotic. The data are given as mean values and standard deviations of three replicates. * p < 0.005 and ** p < 0.01 vs. control. (b) Pictorial illustration of bacterial death and disintegration.
Figure 10. (a) E. coli susceptibility to CeO2 NFs and CeO2−SnO2 composite NFs (50–400 µg/mL). Ciprofloxacin was used as the standard antibiotic. The data are given as mean values and standard deviations of three replicates. * p < 0.005 and ** p < 0.01 vs. control. (b) Pictorial illustration of bacterial death and disintegration.
Nanomaterials 13 01001 g010
Table 1. Comparison of photocatalytic work results of different composite nanomaterials for MB dye degradation.
Table 1. Comparison of photocatalytic work results of different composite nanomaterials for MB dye degradation.
No.PhotocatalystDyeLight SourceTime (min)Degradation Efficiency (%)Reference
1.V2O5/RGO composite MBUV/visible10098.85[4]
2.HAP−MnFe2O4 nanocompositesMBVisible15088[43]
3.CeO2−Cu2O composite nanofibersMBUV/visible18092[67]
4.Mo/N−doped TiO2 nanorods@CNFsMBVisible18079.8[68]
5.PANI nanotube@TiO2 compositeMBVisible30085[69]
6.C−doped ZnO nanofiberMBSolar30>95[70]
7.TiO2/ZrO2 composite nanofibersMBVisible18082.7[71]
8.TiO2−decorated carbon nanofibersMBUV18097.4[72]
9.α−Fe2O3/Bi2MoO6 composite nanofibersMBSunlight24094.8[73]
10.ZnO/CdO alloy nanofibersMBVisible270>90[74]
11.Nanotextured CeO2−SnO2 composite fibersMBVisible12585This study
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

Algethami, J.S.; Hassan, M.S.; Amna, T.; Sheikh, F.A.; Alhamami, M.A.M.; Seliem, A.F.; Faisal, M.; Kim, H.Y. Nanotextured CeO2−SnO2 Composite: Efficient Photocatalytic, Antibacterial, and Energy Storage Fibers. Nanomaterials 2023, 13, 1001. https://doi.org/10.3390/nano13061001

AMA Style

Algethami JS, Hassan MS, Amna T, Sheikh FA, Alhamami MAM, Seliem AF, Faisal M, Kim HY. Nanotextured CeO2−SnO2 Composite: Efficient Photocatalytic, Antibacterial, and Energy Storage Fibers. Nanomaterials. 2023; 13(6):1001. https://doi.org/10.3390/nano13061001

Chicago/Turabian Style

Algethami, Jari S., M. Shamshi Hassan, Touseef Amna, Faheem A. Sheikh, Mohsen A. M. Alhamami, Amal F. Seliem, M. Faisal, and H. Y. Kim. 2023. "Nanotextured CeO2−SnO2 Composite: Efficient Photocatalytic, Antibacterial, and Energy Storage Fibers" Nanomaterials 13, no. 6: 1001. https://doi.org/10.3390/nano13061001

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