Next Article in Journal
The Potential of Glucosinolates and Their Hydrolysis Products as Inhibitors of Cytokine Storms
Next Article in Special Issue
A Soothing Lavender-Scented Electrospun Fibrous Eye Mask
Previous Article in Journal
Phthalate Esters in Different Types of Cosmetic Products: A Five-Year Quality Control Survey
Previous Article in Special Issue
Batch Preparation and Performance Study of Boehmite-Based Electrospun Nanofiber Separators for Lithium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress in Photocatalytic Applications of Electrospun Nanofibers: A Review

1
Department of Materials Science, Nanotechnology and Engineering Physics, Satbayev University, Almaty 050032, Kazakhstan
2
Institute of Nuclear Physics, Almaty 050032, Kazakhstan
3
Bes Saiman Group, Almaty 050057, Kazakhstan
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(20), 4824; https://doi.org/10.3390/molecules29204824
Submission received: 11 September 2024 / Revised: 7 October 2024 / Accepted: 9 October 2024 / Published: 11 October 2024

Abstract

:
Electrospun fiber-based photocatalysts demonstrate significant potential in addressing global environmental and energy challenges, primarily due to their high specific surface areas and unique properties. This review examines recent advances in the application of these materials in photocatalytic processes, with a particular focus on water splitting and hydrogen production. The principles of the electrospun method are described in detail, along with the operating parameters, material characteristics, and environmental conditions that affect the fiber formation. Additionally, the review discusses the challenges, advantages, and future prospects of photocatalysts incorporating carbon materials, metals, semiconductors, and hybrid structures with improved performance. These materials have the potential to significantly improve the efficiency of hydrogen energy production, water purification, and CO2 recovery, highlighting their importance in engineering sciences.

1. Introduction

Global environmental and energy issues have increasingly attracted attention from the scientific community due to the detrimental effects of fossil fuels on ecosystems [1]. The excessive use of oil, coal, and gas has led to a significant increase in greenhouse gas emissions, resulting in global warming and associated challenges, such as extreme weather events and rising sea levels [2,3]. For example, according to the World Health Organization (WHO), approximately 99 percent of the global population breathes air that fails to meet adequate quality standards, posing serious health risks. Each year, over 13 million people die worldwide due to air pollution and environmental causes [4].
One solution to these problems is to reduce reliance on fossil fuels and partially transition to clean renewable energy sources. A notable example of such a transition is the European Union, which, for the first time in 2021, produced 10.3% of the world’s electricity from wind and solar energy, doubling its share since 2015, when the Paris Agreement was signed [5]. In May 2022, the European Commission presented the REPowerEU plan to phase out fossil fuels, proposing an investment of EUR 210 billion in renewable energy and energy efficiency. According to this plan, the capacity of renewable energy in the EU is expected to reach 1236 GW by 2030 [6]. Similar initiatives are being adopted in other countries. For example, the United States passed legislation in 2022 aimed at reducing greenhouse gas emissions by 1 Gt, including tax credits for the production of clean hydrogen, the development of new clean technologies, air capture, and clean fuels [7].
Hydrogen, in particular, has garnered significant interest as an environmentally friendly energy carrier that is expected to play a key role in the decarbonization and electrification of major energy systems in the near future [8,9]. Its high energy density and unique properties make it suitable for a wide range of applications, including fuel cell vehicles, stationary power generation, and industrial processes. However, the existing difficulties in hydrogen production, storage, and transportation limit its widespread adoption [10,11]. Hydrogen production methods can be classified based on the hydrogen or energy source. These methods are divided into renewable [12] (e.g., wind energy [13], solar energy [14], geothermal energy [15,16]) and non-renewable [17] (e.g., fossil resources [13], coal gasification, catalytic reforming, partial oxidation) [18]. The use of solar energy to produce hydrogen has several advantages over traditional methods, including lower costs and scalability [19,20].
Photocatalytic water decomposition, discovered in 1972 [21], is a promising, efficient, and simple method for producing renewable hydrogen. However, despite numerous studies, the large-scale practical application of photocatalytic water decomposition to produce pure hydrogen remains challenging [22]. Nevertheless, ongoing research allows us to confidently discuss the prospects and competitiveness of this method for effective hydrogen production [23]. The primary issue limiting the practical application of current photocatalysts is the low quantum efficiency of solar energy conversion. This challenge can be addressed by creating hybrid composite structures based on photocatalysts, achieved through doping with various elements or combining them with the plasmonic effect of noble metals (Au, Ag, or Pt) or narrow-bandwidth semiconductors [24]. From this perspective, electrospinning is a versatile, simple, and cost-effective technique for producing composite structures with a high specific surface area-to-volume ratio and the ability to regulate their composition and morphology. Given that this method is widely used in various fields, such as water purification [25], tissue engineering [26,27], and the production and delivery of medicinal products [28,29], and its scalability, it has significant potential for developing effective photocatalytic systems. Unlike conventional photocatalysts, electrospun nanofiber photocatalysts have several advantages, such as significant specific surface area, high porosity, and the possibility of surface modification to enhance photoactivity [30].
Over the past decade, there has been a notable increase in research articles featuring the keywords “photocatalysis” and “electrospinning”. An analysis of publications indexed in the Scopus database from 2014 to 2024 reveals that 60,851 publications include the keyword “electrospinning” and 36,138 publications feature “photocatalysis”. Figure 1 suggest that research in this area will continue to increase. A quantitative analysis of the countries, institutions, and journals most engaged in research on photocatalysis and electrospinning shows that China is the most active, followed by India, the United States, South Korea, Japan, Germany, and Iran. In terms of citation impact, the United States leads, followed by Germany and Japan, indicating the wide use of the electrospinning method for developing composite photocatalytic systems. This review discusses the latest achievements in the formation of hybrid photocatalytic systems using electrospinning, specifically for hydrogen production. A detailed analysis of the effect of the method on the mechanism and efficiency of photocatalytic water decomposition is performed, along with the exploration of future prospects and challenges. Electrospinning is demonstrated as an efficient, low-cost approach to developing hybrid composite photocatalytic systems.

2. Principle of Electrospinning Method

Electrospinning is a method used to fabricate one-dimensional (1D) nanofibers from organic, inorganic, and hybrid materials [31,32]. The concept of electrospinning originated in 1600 with William Gilbert, who observed the formation of a drop of water in an electric field. In 1902, John Cooley and William Morton filed the first patents describing a prototype electrospinning rig. In 1934 and 1944, Anton Formhals filed several more patents, improving equipment to commercialize the electrospinning process [33]. The foundation of electrospinning research can be traced back to the pioneering work by Taylor from 1964 to 1969, which modelled the formation of spherical and conical shapes in polymer solutions or melt droplets under the influence of an electric field [34]. These studies initiated the development of the electrospinning method. In the early 1980s, Donaldson Co. Inc. in the United States began producing and marketing air filters fabricated using this method [33]. A new impetus for the development of electrospinning came with the introduction of electron microscopes in the early 1990s, which enabled the examination of nanoscale structures. Researchers, including Reneker and Rutledge, discovered that nanoscale fibers could be drawn from solutions of various organic polymers [35,36].
Electrospinning operates based on an electrohydrodynamic process in which a droplet of a liquid polymer is electrified to form a jet that is then stretched and elongated to produce a fiber. The typical electrospinning setup is relatively simple (Figure 2), making it accessible to most laboratories. The main elements of an electrospinning apparatus include a high-voltage power source, a needle through which the polymer solution is dispensed, and a conductive collector to collect the incoming polymer. These elements are combined into a single electrical circuit. During electrospinning, the liquid is squeezed out of the needle, forming pendant droplets due to surface tension. As the electrical voltage at the tip of the needle increases, the surface tension of the polymer solution is overcome, resulting in the formation of a Taylor cone from which a charged jet is ejected. Once the voltage is sufficient, the polymer jet rushes from the top of the cone towards the collector. The diameter of the resulting fibers depends on several factors. In air, part of the solvent evaporates, and the jet splits, depositing pure polymer on the collector as randomly or directionally aligned nanofibers with sizes ranging from nanometers or micrometers. The resulting material resembles a fibrous, porous soft fabric or a thin elastic coating. The fiber formation process can be divided into four stages [33].
Formation of a Taylor cone from which liquid is released.
Expansion of the charged jet along a straight line.
-
Stretching of the jet due to the increased electric field voltage, leading to electrical bending instability.
-
Solidification of the jet in the form of solid fiber(s) on a grounded collector.
The process of fiber formation by electrospinning is influenced by various operating, material, and environmental parameters.
First, the formation and deposition of fibers are affected by the electric field strength. Thinner fibers are usually formed at higher voltages, while thicker fibers or no fibers are formed at low voltages [38,39,40]. For example, increasing the voltage to 60 kV produced the thinnest nanofibers with a diameter of 190.21 ± 36.65 nm [41].
Second, the diameter and morphology of the resulting fibers are affected by the flow rate at which the polymer solution is fed into or ejected from the spinneret. To maintain a stable Taylor cone, the liquid flow rate must be adjusted continuously to match the voltage. A uniform Taylor cone results in the production of uniform nanofibers with narrow dispersion [42,43,44,45].
Third, the distance between the needle tip and the collector impacts the diameter and shape of the fibers. The minimum distance required ensures that the solvent has sufficient time to evaporate before the fiber reaches the collector. Increasing this distance results in thinner fibers. However, if the distance is too large or small, bead formation may occur [46,47].
Fourth, material properties, including the molecular weight of the polymer, viscosity of the solution, and concentration, play a crucial role in determining fiber size and structure. A higher polymer molecular weight or concentration increases fiber size. A study on the effects of PVA concentration and molecular weight (low, medium, and high) on the morphology of electrospun fibers showed that the formation of beads decreased with higher polymer concentration and molecular weight. However, at high molecular weights and concentrations, uneven and thick fibers were observed due to the increased solution viscosity [48]. High-viscosity colloidal polymer solutions can lead to an unstable jet state, resulting in fibers with a banded structure [49]. If the viscosity is too high, the solution may dry at the tip of the needle before electrospinning begins, complicating the process [47]. Notably, the choice of polymer significantly affects the formation of fibers; higher-viscosity polymer solutions typically produce fibers with larger diameters [50]. For example, using 14 wt% and 16 wt% polyvinylidene fluoride (PVDF) solutions, the 16 wt% solution yielded fewer beads and formed a membrane-like structure [51]. This is attributed to the formation of long-chain linkages in the polymer, which ensure the continuity of the jet [52].
Finally, environmental parameters such as temperature [53] and humidity significantly affect the electrospinning process and fiber morphology. Increasing the temperature to 45 °C decreased fiber diameter from 600 nm to 213 nm due to decreased solution viscosity and surface tension [54]. Similarly, increasing the relative humidity from 5.1% to 48.7% decreased fiber diameters from 253 nm to 144 nm. At humidity levels above 50%, beads formed on the fibers due to capillary instability [55].

3. Application of Electrospun One-Dimensional Photocatalysts

Electrospinning is a versatile, simple, and cost-effective method for producing high-quality 1D nanomaterials. However, there are certain problems associated with photocatalysts made from electrospun fibers. These problems include the recombination of charge carriers: the rapid recombination of electrons and holes reduces the efficiency of photocatalysis. Additionally, degradation of materials under ultraviolet light can lead to a loss of activity, and low quantum efficiency limits their application in visible light conditions [56].
To address these issues, the surfaces of the materials can be modified. Since the photocatalytic properties of fibers primarily depend on the catalytic components, this opens up opportunities for creating 1D fibrous materials with tunable chemical composition, morphology, high specific surface area, high porosity, and varying fiber diameters. Such 1D photocatalysts are particularly promising for various applications, as they provide better light harvesting and enhanced reaction efficiency. Additionally, doping and composite formation are possible: incorporating metallic or non-metallic dopants, as well as creating composites with other semiconductors, can reduce electron–hole recombination and extend the activity of photocatalysts (Figure 3) in the visible spectrum [57].
This section discusses the unique properties of 1D nanofibers and their applications in photocatalytic water decomposition, with a focus on various photocatalyst options incorporating carbon materials, metals, and semiconductors.

3.1. Development of Various Electrospun Composite Nanofibers for Photocatalytic Applications

One of the most effective ways to improve the properties of 1D photocatalysts involves the incorporation of carbon materials, such as graphene, carbon nanotubes, and carbon nanoparticles. Graphene, a 2D carbon material with a honeycomb lattice structure, exhibits high electrical conductivity, sufficient mechanical strength, and good chemical stability, making it an ideal co-catalyst. It can be easily incorporated into conductive fibers by dispersing it in a polymer solution before electrospinning. The resulting fibers exhibit high electrical conductivity and a large surface area, facilitating efficient light absorption and charge separation on the catalyst surface [59,60,61]. In addition, graphene can be modified with various surfactants to enhance its adsorption efficiency. For example, reduced graphene oxide (rGO) and titanium dioxide (TiO2) composite fibers have shown significant improvements in the photocatalytic degradation of methyl orange compared to pure TiO2 [62]. Similar studies demonstrated that C/TiO2 nanofibers carbonized at 400 °C exhibit superior photocatalytic activity for MB degradation. This increased efficiency is attributed to the transfer of photogenerated electrons from the conduction band of TiO2 to the carbon during photocatalysis, leading to a more efficient separation of electrons and holes [63].
However, the primary limitation of graphene and graphene oxide as photocatalysts is their poor absorption of visible light. To address this, dye molecules that absorb visible light have been used as effective photosensitizers [64]. Positively charged dye molecules that absorb visible light are easily attracted to the negatively charged graphene oxide system due to the electrostatic forces of attraction. Studies [65] demonstrated that a photocatalyst comprising graphene oxide and positively charged dye molecules, without the inclusion of noble metals, exhibited activity two orders of magnitude higher than that of conventional TiO2-based catalysts.
There is also growing interest in carbon nitride, with approximately 10–12% of research on photocatalytic hydrogen production focusing on photocatalysts based on graphite-like carbon nitride (gC3N4) [66]. It is known for its stability, good light absorption (up to 460 nm) [67], large surface area and cost-effectiveness [68]. However, its photocatalytic efficiency is limited due to the high rate of electron–hole recombination, which prevents efficient reduction of H2 [69].
Semiconductor-based photocatalysts, particularly in the form of nanofibers, such as TiO2 [70], ZnO [71,72], and SrTiO3 [73,74,75], are known for their light-absorbing properties, high surface-area-to-volume ratio, and mechanical strength. TiO2 [76,77] exhibits high photocatalytic activity owing to its three main crystal structures with bandgap energies of 3.2, 3.0 and 3.1 eV. ZnO [78] has a higher light absorption capacity than TiO2; however, the issue of photocorrosion under UV radiation remains [79].
While these materials are extensively studied, semiconductors face challenges related to high bandgap energy and chemical stability. To improve their efficiency, many researchers have used metal or non-metal doping or combined them with other semiconductors that operate effectively under visible-light irradiation [80,81]. However, this approach limits the ability of semiconductors to absorb visible light, which constitutes the majority of the solar spectrum.
In this context, sulfide (CdS, MoS2) [82] and nitride (GaN, InN) [83] semiconductors have attracted considerable interest due to their small bandgaps, which enable efficient visible-light absorption. Sulfide semiconductors possess bandgaps ranging from 2.0 to 2.4 eV, enhancing their performance in the visible spectrum. However, a significant challenge with these materials is the high recombination rate of electron–hole pairs [84], which reduces their photocatalytic efficiency. Various approaches have been used to address this problem, including the addition of cocatalysts and the creation of heterojunctions [85], which aim to increase the lifetime of the charge carriers and improve photocatalytic efficiency.
Noble metals such as platinum, gold and palladium are widely used as cocatalysts for semiconductors [86,87,88,89]. These metals facilitate efficient electron transfer and reduce the probability of recombination. Additionally, they act as “traps” for electrons, extending their lifetime and increasing the number of photogenerated charge carriers.
Base metals such as iron, copper, nickel, and cobalt [90,91,92], are also used as cocatalysts. Although they are less efficient than noble metals, they offer a more cost-effective solution by creating additional active sites on the photocatalyst surface and enhancing charge-carrier dissociation.

3.2. Application of Electrospun Nanofibers for Hydrogen Production

Nanofibers obtained by electrospinning have recently been considered as promising candidates for applications in the photocatalytic production of solar hydrogen [73,93]. The process of photocatalytic water splitting to produce hydrogen involves three stages: light absorption leading to the formation of electron–hole pairs, charge separation and transport, followed by oxidation–reduction reactions. To efficiently split water, the photocatalyst must have a sufficient bandgap and suitable potentials for hydrogen and oxygen generation [94,95,96,97]. Therefore, there is a demand for the development of new, more efficient photocatalysts to harness solar energy effectively. A promising direction is the use of nanofiber photocatalysts, which have unique properties such as high surface area, improved charge transfer, and the ability to finely tune the morphology and structure of the material.
Electrospinning is an effective approach for synthesizing nanofiber photocatalysts [98]. The composites obtained using this method demonstrated improved characteristics compared to traditional photocatalysts, making them promising candidates for practical applications in hydrogen energy. In particular, sulfur-doped g-C3N4 nanofibers demonstrate 2.84 times higher activity in hydrogen evolution (632 μmol/h g) compared to bulk sulfur-doped g-C3N4, under similar conditions [93].
In addition, powder photocatalysts tend to settle, complicating recovery and reuse. In contrast, composite nanofiber membranes do not require mechanical stirring or ultrasound, which improves reaction stability. In particular, the ZIS/PAN membrane showed 3.7 times higher hydrogen production than ZIS powder while also simplifying the recovery process and demonstrating good stability, making it a promising solution for practical photocatalyst applications [99,100].
Carbon nanofibers obtained by electrospinning are highly effective at accumulating and transporting charge [74,101]. However, their photocatalytic efficiency may be hindered by issues such as limited active sites, rapid charge recombination, and high overpotential during hydrogen generation [93]. Possible solutions include the decoration of co-catalysts, which can increase the absorption of visible light, create more active sites, and influence charge carrier dynamics. For example, hydrogen evolution on electrospun porous TiO2 nanofibers with NiS and Pt cocatalysts, deposited via wet-chemical and self-assembly methods, increased 292 times compared to pure TiO2 nanofibers [102]. This significant increase in visible-light photocatalytic activity of the TiO2/NiS/Pt nanofibers upon deposition with cocatalysts can be attributed to enhanced absorption of visible light and more efficient separation of photogenerated electrons and holes.
In cocatalyst-electrospun nanofiber systems, selecting the optimal amount of cocatalyst is crucial for maximizing hydrogen generation efficiency. For example, one study [103] showed that as the content of the cocatalyst Cd0.5Co0.5S increased from 1.0 to 9.0 wt. %, the rate of hydrogen production using solar energy initially increased and then decreased. The highest efficiency of 4.55 mmol g−1 h−1 was achieved at 5.0 wt. % Cd0.5Co0.5S. The decrease in efficiency at higher concentrations was attributed to excessive cocatalyst content, which hinders charge separation, blocks active centers, and promotes the recombination of charge carriers, thus inhibiting photocatalytic water splitting [104]. As demonstrated in Table 1, the highest efficiency of photocatalytic hydrogen evolution was achieved with the nanofibers decorated with nanoparticle cocatalysts.
Photocatalytic hydrogen production based on electrospun nanofibers has great potential. However, several challenges remain for its practical implementation. These include (1) increasing the mechanical strength of nanofibers after heat treatment and (2) using safe and non-toxic solvents during electrospinning [37]. To achieve high-efficiency hydrogen generation, the development of photocatalytic material that absorbs light over a broad wavelength range and ensures efficient separation and migration of charged particles is essential. A particularly promising research direction is the fabrication of electrospun nanofibers with controlled properties and the formation of heterostructures with integrated nanoparticles, which could significantly increase their photocatalytic activity.

3.3. Application of Electrospun Nanofibers for Water Treatment

Composite photocatalysts synthesized by electrospinning are promising materials for purifying water from organic pollutants [87]. As previously noted, nanofibers obtained using this method exhibit a high specific surface area and a developed porous structure, enhancing their adsorption properties and access to active photocatalytic centers. This facilitates the effective destruction of stable organic compounds, such as benzene rings and carbonyl groups [110].
The most widely used photocatalysts in nanofibers are metal oxides, such as TiO2 and ZnO, due to their high catalytic activity and environmental safety [111]. When exposed to ultraviolet or visible light, these materials generate electrons and holes, which react with oxygen to form active oxides and hydroxyl radicals. These highly reactive species break the stable chemical bonds in pollutant molecules, such as C=C and CN bonds, mineralizing them into carbon dioxide and water. However, the bandgaps of these materials, 3.2 eV (for TiO2) and 3.37 eV (for ZnO), limit their absorption to ultraviolet radiation, which constitutes only 4–5% of the solar spectrum, reducing their effectiveness in the visible range [112]. In addition, the rapid recombination of photogenerated electron–hole pairs reduces the number of charge carriers involved in redox reactions, thereby diminishing the photocatalytic efficiency [113].
To increase the photosensitivity and efficiency of TiO2, various modifications including doping with noble metals, transition metals, rare-earth elements, and nonmetals have been explored [30]. In particular, N,F-doping of TiO2-δ nanofibers, developed by researchers [114], increased the degradation rate of RhB, MB and Cr(VI) dyes by 11.8, 3.2 and 2.8 times, respectively, compared to commercial TiO2. Doping narrows the bandgap, enabling the photocatalyst to function under visible light. This effect is attributed to the hybridization of the 2p orbitals of nitrogen with the 2p orbitals of oxygen in the valence band of TiO2, enhancing the separation of electron–hole pairs and reducing recombination [115,116]. Fluorine doping also improves light absorption at long wavelengths due to the similarity in ionic radii between fluorine and oxygen. The researchers further suggested that multi-element doping could provide even more improvements in photocatalytic performance compared to single-component doping.
Another promising approach to improving the photocatalytic properties of materials involves the creation of binary and ternary composites based on various metal oxides. For example, combining TiO2 with ZnO or Bi2WO6 leads to the formation of heterostructures that promote more efficient charge separation and broaden the spectrum of photocatalytic activity [117,118]. Ternary composites, such as NT@BMO and NT@BWO, exhibit high photocatalytic efficiency due to the synergistic interaction between components, resulting in improved adsorption of pollutants and more effective degradation of organic molecules.
Similarly, the composite NT@BWMO, a ternary heterostructure with controlled morphology and composition, was obtained by depositing BWMO on the surface of N-TiO2 NF [114]. The [email protected], [email protected] and [email protected] samples synthesized at different W/Mo molar ratios (0.25, 0.75, 0.5, 0.5, and 0.75/0.25, respectively) exhibited > 99% tetracycline removal efficiency under visible-light irradiation. Among them, the [email protected] sample showed the highest tetracycline degradation rate (TC) of 0.0054 min−1, which is 9.0, 2.5 and 1.8 times higher than that of N-TiO2, NT@BMO and NT@BWO, respectively. The enhanced photocatalytic activity of [email protected] is attributed to the improved adsorption, optimal crystal size, a narrower bandgap, and enhanced visible-light absorption. Photoluminescence (PL) and photoelectrochemical performance (PEC) analyses confirmed that a lower tungsten ion content improved carrier mobility and increased the carrier separation rate. Radical scavenging experiments and electron paramagnetic resonance (EPR) results showed that ・O2 radicals and h+ holes played a crucial role in the photocatalytic degradation process, while the influence of hydroxyl radicals was minimal. The photocatalytic activity of [email protected] slightly decreased from 99.4% to 91%, indicating that the material can be reused multiple times without significant loss of efficiency.
Based on the data presented in Table 2, it can be concluded that the use of composites with multicomponent heterostructures achieves 99–100% water purification from various organic pollutants under the influence of visible light. However, the addition of g-C3N4 to nanofibers does not result in equally high photocatalytic activity. This is attributed to the limitations of the material, such as high charge recombination rates, low conductivity, and a tendency to aggregate, which reduces surface area. However, graphite-like carbon nitride (g-C3N4), consisting of tri-s-triazine structural units, is the most stable isomer, providing nanofibers with high thermal and chemical stability, thereby increasing their durability in photocatalytic applications [119].
Electrospinning allows the synthesis of materials with unique adsorption and photocatalytic properties, ensuring the complete removal of organic pollutants [120]. Nanofiber composites also exhibit high wear resistance, making them suitable for repeated use. However, scaling this technology for industrial application remains a challenge. To integrate this technology into mass production, new approaches that increase productivity without compromising material quality must be developed. Promising areas include increasing the speed of nanofiber formation, introducing multijet electrospinning, and developing cost-effective raw materials.
Table 2. Recent advances in electrospun fiber photocatalysts.
Table 2. Recent advances in electrospun fiber photocatalysts.
YearPhotocatalystLight SourcePollutantTimeEfficiencyRef.
2021 ZnFe2O4/Ag/AgBr UV light Rhodamine B 100 min86%[121]
2020 Bimetal-PANNM UV-visible Reactive blue 60 min99.99%[122]
2021 TiO2@Ag@Cu2O Visible light Methylene Blue90 min99%[123]
2020 ZnIn2S4/SnO2 Visible light Cr(VI) 80 min100%[124]
2020 Co-CdSe@ECNFs Visible light Methylene Blue90 min87%[125]
2021 ZnO UV light Methylene Blue85 min90%[87]
2021 Bi2O3/g-C3N4 Visible light Tetracycline 180 min~60%[126]
2020 Mn2+/ZnO Visible light Rhodamine B 260 min~80%[127]
2021 Ag3PO4-TiO2CNFs Visible light Methylene Blue10 min100%[128]
2021 Ag/BiVO4 Visible light Rhodamine B 20 min~100%[129]
2020 ZnO-TiO2CNFs Visible light Methylene Blue120 min~95%[130]
2023 g-C3N4 (TiO2/g-C3N4@LCNFs UV light Rhodamine B 90 min83.8%[131]
2022 Chitin-modified and graphene oxide (GO) bridged TiO2/carbon fibers (CGTC) Visible light Rhodamine B 60 min86.8%[132]
2022 PAN/Bi2MoO6/Ti3C2 (PAN/BT) UV-visible Tetracycline 180 min 90.3% [133]
2023 CuBi2O4@WO3 Visible light Tetracycline hydrochloride (TCH) 120 min70.42%[134]

3.4. Application of Electrospun Nanofibers for CO2 Reduction

One of the most promising applications of composite photocatalysts produced via electrospinning is the reduction of CO2 into high-value-added products [135]. This process not only mitigates carbon dioxide emissions but also generates clean energy sources, playing a pivotal role in sustainable development and the fight against climate change. In recent years, significant research efforts have been directed towards the development and optimization of materials for photocatalytic CO2 reduction [136].
A particularly promising approach for enhancing photocatalytic CO2 reduction efficiency is the synthesis of graphene-based nanostructured materials. Graphene, due to its superior charge carrier mobility, large surface area, structural flexibility, and chemical stability, has been widely investigated for improving the photocatalytic performance of semiconductors. Specifically, the development of homogeneous ternary nanocomposites composed of graphene, noble metals, and semiconductors—without agglomeration or over-packing of grapheme—is emerging as a reliable approach for CO2 photoreduction [137]. Utilizing negative electric potential in the coaxial electrospinning technique enables the production of core–shell nanofibers (NFs), where metal ions concentrate beneath the rGO layer that uniformly wraps the entire fiber. These rGO monolayers, concentrated on the surface of silver (Ag), efficiently transport and collect photogenerated electrons for CO2 reduction while enhancing light-capturing abilities, allowing the utilization of a broader light spectrum, from ultraviolet to visible. Under visible light, rGO/Ag/TiO2 NFs demonstrated 25 times higher CO2 photoreduction efficiency, producing 4301 μmol gNF−1 CH4 in 7 h, compared to conventional semiconductor nanofibers. Similarly, graphene’s efficacy is highlighted in a study combining graphene with PVDF (polyvinylidene fluoride) and TiO2, which achieved a yield of 363 μmol g−1 in 1 h, significantly higher than the result without graphene (28.3 μmol g−1 in 1 h) [138]. However, it is important to note that an excessive amount of graphene may complicate the electrospinning process and reduce photocatalytic efficiency due to aggregation, which blocks incident light and hinders photocatalysis.
In addition to graphene, doping semiconductors with single-atom metal catalysts has been recognized as an effective strategy for enhancing the efficiency of photocatalytic CO2 reduction [139]. Titanium dioxide (TiO2), known for its high photocatalytic activity, thermodynamic stability, non-toxicity, and low cost, has been widely utilized in CO2 reduction processes [140]. Electrospun TiO2 nanoparticles serve as ideal substrates for the growth of secondary nanostructures, facilitating the creation of heterojunction photocatalysts. These hybrid heterojunctions improve electron–hole separation, enhance light absorption, and promote reactant activation, resulting in superior photocatalytic performance. For instance, TiO2 nanofibers coated with graphitic carbon nitride (gC3N4) achieved CO and CH4 yields of 5.19 and 1.65 μmol/g, respectively, representing a 1.8-fold increase in CO2 conversion performance compared to gC3N4 alone [141]. However, TiO2 faces limitations such as low surface active site density, a high recombination rate of photogenerated charge carriers, and limited CO2 capture efficiency. To overcome these challenges, various modification strategies have been explored [142].
In parallel with TiO2, alternative semiconductor materials have been investigated for photocatalytic CO2 reduction, including metal oxides (e.g., ZnO), metal chalcogenides (e.g., ZnS, CdS), perovskite halides (e.g., CsPbBr3), MXenes (e.g., Ti3C2), layered double hydroxides, and metal–organic frameworks [143,144,145]. However, these materials often exhibit low practical efficiency due to rapid electron–hole recombination and limited sunlight utilization. Consequently, significant efforts have been made to develop more efficient photocatalysts by manipulating their morphology, adjusting the bandgap, and introducing metals to improve performance [146].
Recently, single-atom catalysts (SACs) have gained considerable attention in various catalytic reactions due to their unique physical and chemical properties [147]. Several SAC-based photocatalysts, including Cu/CN [148] and Pt@WS2 [149], have been synthesized. In particular, doping TiO2 with noble metals significantly enhances catalytic activity through three primary mechanisms: Fermi level alignment, efficient electron trapping, and the creation of thermal catalytic sites for adsorbed molecules and reaction intermediates [123]. For instance, co-deposition of Pt on TiO2 has demonstrated exceptional activity in CO2 conversion to CH4, outperforming pure TiO2 by a factor of 10 [150]. This enhanced performance is attributed to Pt nanoparticles acting as electron traps, facilitating charge separation on the TiO2 surface. On the other hand, Au/TiO2 nanofibers exhibit lower activity in CH4 production but demonstrate higher CO production, highlighting new possibilities for selective control of photocatalytic reaction products.
The key parameters and results of recent studies on the photocatalytic reduction of CO2 into value-added products are summarized in Table 3. Based on a comparative analysis, the Ni-MoP@NCPF photocatalyst exhibits the highest efficiency in reducing CO2 to CO [151] (953.33 μmol g−1h−1), while the Graphene@PVDF@TiO2 composite [138] shows superior efficiency in CH4 formation (363 μmol g−1 h−1) under visible-light irradiation. When comparing photocatalysts, it is essential to consider the use of sacrificial agents, carbon and hydrogen sources, and the type of radiation.

4. Conclusions

In conclusion, electrospun nanofiber photocatalysts demonstrate significant potential for practical application in hydrogen energy, water purification and CO2 recovery. The advantages and prospects of various photocatalyst modifications, including the addition of carbon materials, metals, and semiconductors, as well as the development of hybrid structures with improved characteristics, are discussed. Due to their unique physicochemical properties, such as large specific surface area and improved charge separation, these materials can significantly improve photocatalytic efficiency. However, for widespread implementation, several technical challenges must be addressed, including increasing the mechanical strength of nanofibers post-heat treatment and using safe, non-toxic solvents during synthesis. The prospects for further development of electrospun composite materials offer broad opportunities to enhance their photocatalytic properties. Research in this area can focus on optimizing the composition and structure of nanofibers, utilizing new co-catalysts, and modifying the surface to improve the efficiency of light absorption and solar energy conversion processes. A key challenge is addressing the issues related to electron–hole recombination, as well as developing cost-effective solutions to increase the lifespan and stability of photocatalysts. These directions could significantly expand the application of electrospun materials in environmentally friendly technologies and energy-saving processes, including photocatalytic hydrogen production and water purification.

Author Contributions

Writing—original draft preparation, A.S.; Conceptualization, methodology, investigation, N.I. and A.B.; Project administration, funding acquisition, A.D. and S.K.; Writing—review and editing, C.D. and Z.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by the Science Committee of the Ministry of Science and Higher Education and of the Republic of Kazakhstan (Grant No. BR18574073).

Institutional Review Board Statement

The study did not require ethical approval.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Letcher, T.M. 9.01—Introduction to Environmental and Social Issues. In Comprehensive Renewable Energy, 2nd ed.; Letcher, T.M., Ed.; Elsevier: Oxford, UK, 2022; pp. 1–8. ISBN 978-0-12-819734-9. [Google Scholar]
  2. Sharma, S.; Basu, S.; Shetti, N.P.; Kamali, M.; Walvekar, P.; Aminabhavi, T.M. Waste-to-Energy Nexus: A Sustainable Development. Environ. Pollut. 2020, 267, 115501. [Google Scholar] [CrossRef] [PubMed]
  3. Yergaziyeva, G.; Kuspanov, Z.; Mambetova, M.; Khudaibergenov, N.; Makayeva, N.; Daulbayev, C. Advancements in Catalytic, Photocatalytic, and Electrocatalytic CO2 Conversion Processes: Current Trends and Future Outlook. J. CO2 Util. 2024, 80, 102682. [Google Scholar] [CrossRef]
  4. Nations, U. Renewable Energy—Powering a Safer Future. Available online: https://www.un.org/en/climatechange/raising-ambition/renewable-energy (accessed on 5 April 2023).
  5. Nations, U. The Paris Agreement. Available online: https://www.un.org/en/climatechange/paris-agreement (accessed on 5 April 2023).
  6. Directorate-General for Communication (European Commission). REPowerEU with Clean Energy; Publications Office of the European Union: Luxembourg, 2022; ISBN 978-92-76-52458-8. [Google Scholar]
  7. Environment, U.N. Emissions Gap Report. 2022. Available online: http://www.unep.org/resources/emissions-gap-report-2022 (accessed on 5 April 2023).
  8. Hermesmann, M.; Müller, T.E. Green, Turquoise, Blue, or Grey? Environmentally Friendly Hydrogen Production in Transforming Energy Systems. Prog. Energy Combust. Sci. 2022, 90, 100996. [Google Scholar] [CrossRef]
  9. Daulbayev, C.; Lesbayev, B.; Bakbolat, B.; Kaidar, B.; Sultanov, F.; Yeleuov, M.; Ustayeva, G.; Rakhymzhan, N. A Mini-Review on Recent Trends in Prospective Use of Porous 1D Nanomaterials for Hydrogen Storage. S. Afr. J. Chem. Eng. 2022, 39, 52–61. [Google Scholar] [CrossRef]
  10. Ratnakar, R.R.; Gupta, N.; Zhang, K.; van Doorne, C.; Fesmire, J.; Dindoruk, B.; Balakotaiah, V. Hydrogen Supply Chain and Challenges in Large-Scale LH2 Storage and Transportation. Int. J. Hydrogen Energy 2021, 46, 24149–24168. [Google Scholar] [CrossRef]
  11. Qureshi, F.; Yusuf, M.; Arham Khan, M.; Ibrahim, H.; Ekeoma, B.C.; Kamyab, H.; Rahman, M.M.; Nadda, A.K.; Chelliapan, S. A State-of-The-Art Review on the Latest Trends in Hydrogen Production, Storage, and Transportation Techniques. Fuel 2023, 340, 127574. [Google Scholar] [CrossRef]
  12. Insights into Renewable Hydrogen Energy: Recent Advances and Prospects. Mater. Sci. Energy Technol. 2020, 3, 319–327. [CrossRef]
  13. Razzhivin, I.A.; Andreev, M.V.; Kievec, A.V. Increasing the Stability of Hydrogen Production in the Wind Energy-Hydrogen System through the Use of Synthetic Inertia of the Wind Turbine. Int. J. Hydrogen Energy 2022, 47, 38495–38505. [Google Scholar] [CrossRef]
  14. Partho, A.T.; Tahir, M.; Tahir, B. Recent Advances in Covalent Organic Framework (COF) Nanotextures with Band Engineering for Stimulating Solar Hydrogen Production: A Comprehensive Review. Int. J. Hydrogen Energy 2022, 47, 34323–34375. [Google Scholar] [CrossRef]
  15. Steingrímsson, B.; Gunnlaugsson, E.; Saemundsson, K.; Axelsson, G. 7.02—Nature and Classification of Geothermal Resources. In Comprehensive Renewable Energy, 2nd ed.; Letcher, T.M., Ed.; Elsevier: Oxford, UK, 2022; pp. 3–17. ISBN 978-0-12-819734-9. [Google Scholar]
  16. Mahmoud, M.; Ramadan, M.; Naher, S.; Pullen, K.; Ali Abdelkareem, M.; Olabi, A.-G. A Review of Geothermal Energy-Driven Hydrogen Production Systems. Therm. Sci. Eng. Prog. 2021, 22, 100854. [Google Scholar] [CrossRef]
  17. Muritala, I.K.; Guban, D.; Roeb, M.; Sattler, C. High Temperature Production of Hydrogen: Assessment of Non-Renewable Resources Technologies and Emerging Trends. Int. J. Hydrogen Energy 2020, 45, 26022–26035. [Google Scholar] [CrossRef]
  18. Monga, D.; Shetti, N.P.; Basu, S.; Kakarla, R.R. Recent Advances in Various Processes for Clean and Sustainable Hydrogen Production. Nano-Struct. Nano-Objects 2023, 33, 100948. [Google Scholar] [CrossRef]
  19. Mandade, P. Chapter 5—Introduction, Basic Principles, Mechanism, and Challenges of Photocatalysis. In Handbook of Nanomaterials for Wastewater Treatment; Bhanvase, B., Sonawane, S., Pawade, V., Pandit, A., Eds.; Micro and Nano Technologies; Elsevier: Amsterdam, The Netherlands, 2021; pp. 137–154. ISBN 978-0-12-821496-1. [Google Scholar]
  20. Kuspanov, Z.; Bakbolat, B.; Baimenov, A.; Issadykov, A.; Yeleuov, M.; Daulbayev, C. Photocatalysts for a Sustainable Future: Innovations in Large-Scale Environmental and Energy Applications. Sci. Total Environ. 2023, 885, 163914. [Google Scholar] [CrossRef]
  21. Ahmed, S.N.; Haider, W. Heterogeneous Photocatalysis and Its Potential Applications in Water and Wastewater Treatment: A Review. Nanotechnology 2018, 29, 342001. [Google Scholar] [CrossRef]
  22. Nishiyama, H.; Yamada, T.; Nakabayashi, M.; Maehara, Y.; Yamaguchi, M.; Kuromiya, Y.; Nagatsuma, Y.; Tokudome, H.; Akiyama, S.; Watanabe, T.; et al. Photocatalytic Solar Hydrogen Production from Water on a 100-M2 Scale. Nature 2021, 598, 304–307. [Google Scholar] [CrossRef]
  23. Goto, Y.; Hisatomi, T.; Wang, Q.; Higashi, T.; Ishikiriyama, K.; Maeda, T.; Sakata, Y.; Okunaka, S.; Tokudome, H.; Katayama, M.; et al. A Particulate Photocatalyst Water-Splitting Panel for Large-Scale Solar Hydrogen Generation. Joule 2018, 2, 509–520. [Google Scholar] [CrossRef]
  24. Tai, L.N.; Long, P.D.; Hong Le, N.T.; Van Hong, L.; Linh, P.T.; Khuyen, B.X.; Tung, B.S.; Lam, V.D. Photocatalytic and Water-Splitting Properties of TiO2 and Ag–TiO2 Films in the Visible Light Region. AIP Adv. 2021, 11, 075118. [Google Scholar] [CrossRef]
  25. Peng, R.; Zhang, S.; Yao, Y.; Wang, J.; Zhu, X.; Jiang, R.; Zhang, J.; Zhang, W.; Wang, C. MOFs Meet Electrospinning: New Opportunities for Water Treatment. Chem. Eng. J. 2023, 453, 139669. [Google Scholar] [CrossRef]
  26. Liverani, L.; Guarino, V.; La Carrubba, V.; Boccaccini, A.R. Porous Biomaterials and Scaffolds for Tissue Engineering. In Encyclopedia of Biomedical Engineering; Narayan, R., Ed.; Elsevier: Oxford, UK, 2019; pp. 188–202. ISBN 978-0-12-805144-3. [Google Scholar]
  27. Mohammadalizadeh, Z.; Bahremandi-Toloue, E.; Karbasi, S. Recent Advances in Modification Strategies of Pre- and Post-Electrospinning of Nanofiber Scaffolds in Tissue Engineering. React. Funct. Polym. 2022, 172, 105202. [Google Scholar] [CrossRef]
  28. Ghosal, K.; Augustine, R.; Zaszczynska, A.; Barman, M.; Jain, A.; Hasan, A.; Kalarikkal, N.; Sajkiewicz, P.; Thomas, S. Novel Drug Delivery Systems Based on Triaxial Electrospinning Based Nanofibers. React. Funct. Polym. 2021, 163, 104895. [Google Scholar] [CrossRef]
  29. Tan, S.M.; Teoh, X.Y.; Le Hwang, J.; Khong, Z.P.; Sejare, R.; Almashhadani, A.Q.; Assi, R.A.; Chan, S.Y. Electrospinning and Its Potential in Fabricating Pharmaceutical Dosage Form. J. Drug Deliv. Sci. Technol. 2022, 76, 103761. [Google Scholar] [CrossRef]
  30. Asgari, S.; Mohammadi Ziarani, G.; Badiei, A.; Ajalloueian, F.; Vasseghian, Y. Electrospun Composite Nanofibers as Novel High-Performance and Visible-Light Photocatalysts for Removal of Environmental Pollutants: A Review. Environ. Res. 2022, 215, 114296. [Google Scholar] [CrossRef] [PubMed]
  31. Tebyetekerwa, M.; Ramakrishna, S. What Is Next for Electrospinning? Matter 2020, 2, 279–283. [Google Scholar] [CrossRef]
  32. SalehHudin, H.S.; Mohamad, E.N.; Mahadi, W.N.L.; Muhammad Afifi, A. Multiple-Jet Electrospinning Methods for Nanofiber Processing: A Review. Mater. Manuf. Process. 2018, 33, 479–498. [Google Scholar] [CrossRef]
  33. Xue, J.; Wu, T.; Dai, Y.; Xia, Y. Electrospinning and Electrospun Nanofibers: Methods, Materials, and Applications. Chem. Rev. 2019, 119, 5298–5415. [Google Scholar] [CrossRef]
  34. Ghosal, K.; Agatemor, C.; Tucker, N.; Kny, E.; Thomas, S. Electrical Spinning to Electrospinning: A Brief History; The Royal Society of Chemistry: Cambridge, UK, 2018. [Google Scholar] [CrossRef]
  35. Haider, A.; Haider, S.; Kang, I.-K. A Comprehensive Review Summarizing the Effect of Electrospinning Parameters and Potential Applications of Nanofibers in Biomedical and Biotechnology. Arab. J. Chem. 2018, 11, 1165–1188. [Google Scholar] [CrossRef]
  36. Shepa, I.; Mudra, E.; Dusza, J. Electrospinning through the Prism of Time. Mater. Today Chem. 2021, 21, 100543. [Google Scholar] [CrossRef]
  37. Panda, P.K.; Sahoo, B.; Ramakrishna, S. Electrospun Nanofibers for Photocatalytic Water Treatment and Hydrogen Generation Application: A Review. Int. J. Hydrogen Energy 2023, 48, 37193–37208. [Google Scholar] [CrossRef]
  38. Joshi, B.; Samuel, E.; Kim, Y.; Yarin, A.L.; Swihart, M.T.; Yoon, S.S. Review of Recent Progress in Electrospinning-Derived Freestanding and Binder-Free Electrodes for Supercapacitors. Coord. Chem. Rev. 2022, 460, 214466. [Google Scholar] [CrossRef]
  39. Can-Herrera, L.A.; Oliva, A.I.; Dzul-Cervantes, M.a.A.; Pacheco-Salazar, O.F.; Cervantes-Uc, J.M. Morphological and Mechanical Properties of Electrospun Polycaprolactone Scaffolds: Effect of Applied Voltage. Polymers 2021, 13, 662. [Google Scholar] [CrossRef]
  40. Tlili, I.; Alkanhal, T.A. Nanotechnology for Water Purification: Electrospun Nanofibrous Membrane in Water and Wastewater Treatment. J. Water Reuse Desalination 2019, 9, 232–248. [Google Scholar] [CrossRef]
  41. Wei, L.; Sun, R.; Liu, C.; Xiong, J.; Qin, X. Mass Production of Nanofibers from Needleless Electrospinning by a Novel Annular Spinneret. Mater. Des. 2019, 179, 107885. [Google Scholar] [CrossRef]
  42. Zhao, K.; Wang, W.; Yang, Y.; Wang, K.; Yu, D.-G. From Taylor Cone to Solid Nanofiber in Tri-Axial Electrospinning: Size Relationships. Results Phys. 2019, 15, 102770. [Google Scholar] [CrossRef]
  43. Ngadiman, N.H.A.; Noordin, M.Y.; Idris, A.; Kurniawan, D. Effect of Electrospinning Parameters Setting towards Fiber Diameter. Adv. Mater. Res. 2014, 845, 985–988. [Google Scholar] [CrossRef]
  44. Sharma, G.K.; James, N.R.; Sharma, G.K.; James, N.R. Electrospinning: The Technique and Applications. In Recent Developments in Nanofibers Research; IntechOpen: London, UK, 2022; ISBN 978-1-80356-387-9. [Google Scholar]
  45. Abdul Hameed, M.M.; Mohamed Khan, S.A.P.; Thamer, B.M.; Rajkumar, N.; El-Hamshary, H.; El-Newehy, M. Electrospun Nanofibers for Drug Delivery Applications: Methods and Mechanism. Polym. Adv. Technol. 2023, 34, 6–23. [Google Scholar] [CrossRef]
  46. Shi, X.; Zhou, W.; Ma, D.; Ma, Q.; Bridges, D.; Ma, Y.; Hu, A. Electrospinning of Nanofibers and Their Applications for Energy Devices. J. Nanomater. 2015, 2015, e140716. [Google Scholar] [CrossRef]
  47. Li, Y.; Li, Q.; Tan, Z. A Review of Electrospun Nanofiber-Based Separators for Rechargeable Lithium-Ion Batteries. J. Power Sources 2019, 443, 227262. [Google Scholar] [CrossRef]
  48. Mwiiri, F.K.; Daniels, R. Influence of PVA Molecular Weight and Concentration on Electrospinnability of Birch Bark Extract-Loaded Nanofibrous Scaffolds Intended for Enhanced Wound Healing. Molecules 2020, 25, 4799. [Google Scholar] [CrossRef]
  49. Lin, H.-N.; Peng, T.-Y.; Kung, Y.-R.; Chiou, Y.-J.; Chang, W.-M.; Wu, S.-H.; Mine, Y.; Chen, C.-Y.; Lin, C.-K. Effects of the Methyl Methacrylate Addition, Polymerization Temperature and Time on the MBG@PMMA Core-Shell Structure and Its Application as Addition in Electrospun Composite Fiber Bioscaffold. Ceram. Int. 2023, 49, 7630–7639. [Google Scholar] [CrossRef]
  50. Ki, C.S.; Baek, D.H.; Gang, K.D.; Lee, K.H.; Um, I.C.; Park, Y.H. Characterization of Gelatin Nanofiber Prepared from Gelatin–Formic Acid Solution. Polymer 2005, 46, 5094–5102. [Google Scholar] [CrossRef]
  51. Suresh, S.; Becker, A.; Glasmacher, B. Impact of Apparatus Orientation and Gravity in Electrospinning-A Review of Empirical Evidence. Polymers 2020, 12, 2448. [Google Scholar] [CrossRef] [PubMed]
  52. Ramakrishna, S. An Introduction to Electrospinning and Nanofibers; World Scientific: Hackensack, NJ, USA, 2005; ISBN 978-981-256-761-1. [Google Scholar]
  53. Yang, G.-Z.; Li, H.-P.; Yang, J.-H.; Wan, J.; Yu, D.-G. Influence of Working Temperature on The Formation of Electrospun Polymer Nanofibers. Nanoscale Res. Lett. 2017, 12, 55. [Google Scholar] [CrossRef] [PubMed]
  54. Ramazani, S.; Karimi, M. Investigating the Influence of Temperature on Electrospinning of Polycaprolactone Solutions. e-Polymers 2014, 14, 323–333. [Google Scholar] [CrossRef]
  55. Wang, J.; Wang, Z.; Ni, J.; Li, L. Electrospinning for Flexible Sodium-Ion Batteries. Energy Storage Mater. 2022, 45, 704–719. [Google Scholar] [CrossRef]
  56. Joly, D.; Jung, J.-W.; Kim, I.-D.; Demadrille, R. Electrospun Materials for Solar Energy Conversion: Innovations and Trends. J. Mater. Chem. C 2016, 4, 10173–10197. [Google Scholar] [CrossRef]
  57. Lu, X.; Wang, C.; Wei, Y. One-Dimensional Composite Nanomaterials: Synthesis by Electrospinning and Their Applications. Small 2009, 5, 2349–2370. [Google Scholar] [CrossRef]
  58. Hariganesh, S.; Vadivel, S.; Maruthamani, D.; Rangabhashiyam, S. Chapter 12—Disinfection by-Products in Drinking Water: Detection and Treatment Methods. In Disinfection By-Products in Drinking Water; Prasad, M.N.V., Ed.; Butterworth-Heinemann: Oxford, UK, 2020; pp. 279–304. ISBN 978-0-08-102977-0. [Google Scholar]
  59. Mayorov, A.S.; Gorbachev, R.V.; Morozov, S.V.; Britnell, L.; Jalil, R.; Ponomarenko, L.A.; Blake, P.; Novoselov, K.S.; Watanabe, K.; Taniguchi, T.; et al. Micrometer-Scale Ballistic Transport in Encapsulated Graphene at Room Temperature. Nano Lett. 2011, 11, 2396–2399. [Google Scholar] [CrossRef]
  60. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef]
  61. Stoller, M.D.; Park, S.; Zhu, Y.; An, J.; Ruoff, R.S. Graphene-Based Ultracapacitors. Nano Lett. 2008, 8, 3498–3502. [Google Scholar] [CrossRef]
  62. Nasr, M.; Balme, S.; Eid, C.; Habchi, R.; Miele, P.; Bechelany, M. Enhanced Visible-Light Photocatalytic Performance of Electrospun rGO/TiO2 Composite Nanofibers. J. Phys. Chem. C 2017, 121, 261–269. [Google Scholar] [CrossRef]
  63. Song, M.; Cao, H.; Zhu, Y.; Wang, Y.; Zhao, S.; Huang, C.; Zhang, C.; He, X. Electrochemical and Photocatalytic Properties of Electrospun C/TiO2 Nanofibers. Chem. Phys. Lett. 2020, 747, 137355. [Google Scholar] [CrossRef]
  64. Yao, Y. Visible-Light Photocatalysis of Carbon-Based Materials; IntechOpen Limited: London, UK, 2018; ISBN 978-1-78923-029-1. [Google Scholar]
  65. Latorre-Sánchez, M.; Lavorato, C.; Puche, M.; Fornés, V.; Molinari, R.; Garcia, H. Visible-Light Photocatalytic Hydrogen Generation by Using Dye-Sensitized Graphene Oxide as a Photocatalyst. Chem.—A Eur. J. 2012, 18, 16774–16783. [Google Scholar] [CrossRef] [PubMed]
  66. Zhurenok, A.V.; Vasilchenko, D.B.; Kozlova, E.A. Comprehensive Review on G-C3N4-Based Photocatalysts for the Photocatalytic Hydrogen Production under Visible Light. Int. J. Mol. Sci. 2023, 24, 346. [Google Scholar] [CrossRef] [PubMed]
  67. Singla, S.; Sharma, S.; Basu, S.; Shetti, N.P.; Reddy, K.R. Graphene/Graphitic Carbon Nitride-Based Ternary Nanohybrids: Synthesis Methods, Properties, and Applications for Photocatalytic Hydrogen Production. FlatChem 2020, 24, 100200. [Google Scholar] [CrossRef]
  68. Mishra, A.; Mehta, A.; Kainth, S.; Basu, S. A Comparative Study on the Effect of Different Precursors for Synthesis and Efficient Photocatalytic Activity of G-C3N4/TiO2/Bentonite Nanocomposites. J. Mater. Sci. 2018, 53, 13126–13142. [Google Scholar] [CrossRef]
  69. Kim, Y.K.; Kim, M.; Hwang, S.-H.; Lim, S.K.; Park, H.; Kim, S. CdS-Loaded Flexible Carbon Nanofiber Mats as a Platform for Solar Hydrogen Production. Int. J. Hydrogen Energy 2015, 40, 136–145. [Google Scholar] [CrossRef]
  70. Dai, Z.; Ren, P.; Cao, Q.; Gao, X.; He, W.; Xiao, Y.; Jin, Y.; Ren, F. Synthesis of TiO2@lignin Based Carbon Nanofibers Composite Materials with Highly Efficient Photocatalytic to Methylene Blue Dye. J. Polym. Res. 2020, 27, 108. [Google Scholar] [CrossRef]
  71. Abbasi, S.; Hasanpour, M.; Ekrami-Kakhki, M.-S. Removal Efficiency Optimization of Organic Pollutant (Methylene Blue) with Modified Multi-Walled Carbon Nanotubes Using Design of Experiments (DOE). J. Mater. Sci. Mater. Electron. 2017, 28, 9900–9910. [Google Scholar] [CrossRef]
  72. Roozban, N.; Abbasi, S.; Ghazizadeh, M. The Experimental and Statistical Investigation of the Photo Degradation of Methyl Orange Using Modified MWCNTs with Different Amount of ZnO Nanoparticles. J. Mater. Sci. Mater. Electron. 2017, 28, 7343–7352. [Google Scholar] [CrossRef]
  73. Sultanov, F.; Daulbayev, C.; Bakbolat, B.; Daulbayev, O.; Bigaj, M.; Mansurov, Z.; Kuterbekov, K.; Bekmyrza, K. Aligned Composite SrTiO3/PAN Fibers as 1D Photocatalyst Obtained by Electrospinning Method. Chem. Phys. Lett. 2019, 737, 136821. [Google Scholar] [CrossRef]
  74. Daulbayev, C.; Sultanov, F.; Korobeinyk, A.V.; Yeleuov, M.; Azat, S.; Bakbolat, B.; Umirzakov, A.; Mansurov, Z. Bio-Waste-Derived Few-Layered Graphene/SrTiO3/PAN as Efficient Photocatalytic System for Water Splitting. Appl. Surf. Sci. 2021, 549, 149176. [Google Scholar] [CrossRef]
  75. Serik, A.; Kuspanov, Z.; Bissenova, M.; Idrissov, N.; Yeleuov, M.; Umirzakov, A.; Daulbayev, C. Effective Photocatalytic Degradation of Sulfamethoxazole Using PAN/SrTiO3 Nanofibers. J. Water Process Eng. 2024, 66, 106052. [Google Scholar] [CrossRef]
  76. Pereira, M.F.G.; Nascimento, M.M.; Cardoso, P.H.N.; Oliveira, C.Y.B.; Tavares, G.F.; Araújo, E.S. Preparation, Microstructural Characterization and Photocatalysis Tests of V5+-Doped TiO2/WO3 Nanocomposites Supported on Electrospun Membranes. Inorganics 2022, 10, 143. [Google Scholar] [CrossRef]
  77. Guerrero-Pérez, M.O. Research Progress on the Applications of Electrospun Nanofibers in Catalysis. Catalysts 2022, 12, 9. [Google Scholar] [CrossRef]
  78. Hilal Elhousseini, M.; Isık, T.; Kap, Ö.; Verpoort, F.; Horzum, N. Dual Remediation of Waste Waters from Methylene Blue and Chromium (VI) Using Thermally Induced ZnO Nanofibers. Appl. Surf. Sci. 2020, 514, 145939. [Google Scholar] [CrossRef]
  79. Navidpour, A.H.; Salehi, M.; Salimijazi, H.R.; Kalantari, Y.; Azarpour Siahkali, M. Photocatalytic Activity of Flame-Sprayed Coating of Zinc Ferrite Powder. J. Therm. Spray Technol. 2017, 26, 2030–2039. [Google Scholar] [CrossRef]
  80. Karim, A.V.; Shriwastav, A. Degradation of Amoxicillin with Sono, Photo, and Sonophotocatalytic Oxidation under Low-Frequency Ultrasound and Visible Light. Environ. Res. 2021, 200, 111515. [Google Scholar] [CrossRef]
  81. Bharathi, P.; Harish, S.; Archana, J.; Navaneethan, M.; Ponnusamy, S.; Muthamizhchelvan, C.; Shimomura, M.; Hayakawa, Y. Enhanced Charge Transfer and Separation of Hierarchical CuO/ZnO Composites: The Synergistic Effect of Photocatalysis for the Mineralization of Organic Pollutant in Water. Appl. Surf. Sci. 2019, 484, 884–891. [Google Scholar] [CrossRef]
  82. Ullah, H.; Haneef, Z.; Ahmad, A.; Butler, I.S.; Nasir Dara, R.; Rehman, Z. MoS2 and CdS Photocatalysts for Water Decontamination: A Review. Inorg. Chem. Commun. 2023, 153, 110775. [Google Scholar] [CrossRef]
  83. Paraskevopoulou, A.; Pandis, P.; Argirusis, C.; Sourkouni, G. Sonochemical Synthesis of Indium Nitride Nanoparticles and Photocatalytic Composites with Titania. Ceramics 2024, 7, 478–490. [Google Scholar] [CrossRef]
  84. Alahmadi, N. Recent Progress in Photocatalytic Removal of Environmental Pollution Hazards in Water Using Nanostructured Materials. Separations 2022, 9, 264. [Google Scholar] [CrossRef]
  85. Samadi, M.; Moshfegh, A.Z. Recent Developments of Electrospinning-Based Photocatalysts in Degradation of Organic Pollutants: Principles and Strategies. ACS Omega 2022, 7, 45867–45881. [Google Scholar] [CrossRef] [PubMed]
  86. Duan, Z.; Huang, Y.; Zhang, D.; Chen, S. Electrospinning Fabricating Au/TiO2 Network-like Nanofibers as Visible Light Activated Photocatalyst. Sci. Rep. 2019, 9, 8008. [Google Scholar] [CrossRef] [PubMed]
  87. Pantò, F.; Dahrouch, Z.; Saha, A.; Patanè, S.; Santangelo, S.; Triolo, C. Photocatalytic Degradation of Methylene Blue Dye by Porous Zinc Oxide Nanofibers Prepared via Electrospinning: When Defects Become Merits. Appl. Surf. Sci. 2021, 557, 149830. [Google Scholar] [CrossRef]
  88. Ruiz-Ramírez, L.R.; Torres-Pérez, J.; Medellín-Castillo, N.; Reyes-López, S.Y. Photocatalytic Degradation of Oxytetracycline by SiO2–TiO2–Ag Electrospun Fibers. Solid State Sci. 2023, 140, 107188. [Google Scholar] [CrossRef]
  89. Zhang, Y.; Liu, Y.; Xie, S.; Huang, H.; Guo, G.; Dai, H.; Deng, J. Supported Ceria-Modified Silver Catalysts with High Activity and Stability for Toluene Removal. Environ. Int. 2019, 128, 335–342. [Google Scholar] [CrossRef]
  90. Lontio Fomekong, R.; Saruhan, B. Synthesis of Co3+ Doped TiO2 by Co-Precipitation Route and Its Gas Sensing Properties. Front. Mater. 2019, 6, 252. [Google Scholar] [CrossRef]
  91. Na, K.-H.; Kim, B.-S.; Yoon, H.-S.; Song, T.-H.; Kim, S.-W.; Cho, C.-H.; Choi, W.-Y. Fabrication and Photocatalytic Properties of Electrospun Fe-Doped TiO2 Nanofibers Using Polyvinyl Pyrrolidone Precursors. Polymers 2021, 13, 2634. [Google Scholar] [CrossRef]
  92. Ghoreishian, S.M.; Ranjith, K.S.; Lee, H.; Park, B.; Norouzi, M.; Nikoo, S.Z.; Kim, W.-S.; Han, Y.-K.; Huh, Y.S. Tuning the Phase Composition of 1D TiO2 by Fe/Sn Co-Doping Strategy for Enhanced Visible-Light-Driven Photocatalytic and Photoelectrochemical Performances. J. Alloys Compd. 2021, 851, 156826. [Google Scholar] [CrossRef]
  93. Wang, L.; Sun, M.; Zhou, B.; Rao, Y.; Yan, T.; Du, B.; Shao, Y. Electrospinning Mediated Solvothermal Preparation of Hierarchical MGa2O4/ZnIn2S4 (M = Ni, Co) Hollow Nanofibers with p-n Heterojunction for Boosting Photocatalytic H2 Evolution. Fuel 2024, 374, 132448. [Google Scholar] [CrossRef]
  94. Kuspanov, Z.; Umirzakov, A.; Serik, A.; Baimenov, A.; Yeleuov, M.; Daulbayev, C. Multifunctional Strontium Titanate Perovskite-Based Composite Photocatalysts for Energy Conversion and Other Applications. Int. J. Hydrogen Energy 2023, 48, 38634–38654. [Google Scholar] [CrossRef]
  95. Kudaibergen, A.D.; Kuspanov, Z.B.; Issadykov, A.N.; Beisenov, R.E.; Mansurov, Z.A.; Yeleuov, M.A.; Daulbayev, C.B. Synthesis, Structure, and Energetic Characteristics of Perovskite Photocatalyst SrTiO3: An Experimental and DFT Study. Eurasian Chem.-Technol. J. 2023, 25, 139–146. [Google Scholar] [CrossRef]
  96. Bissenova, M.; Umirzakov, A.; Mit, K.; Mereke, A.; Yerubayev, Y.; Serik, A.; Kuspanov, Z. Synthesis and Study of SrTiO3/TiO2 Hybrid Perovskite Nanotubes by Electrochemical Anodization. Molecules 2024, 29, 1101. [Google Scholar] [CrossRef] [PubMed]
  97. Matus, E.V.; Ismagilov, I.Z.; Mikhaylova, E.S.; Ismagilov, Z.R. Hydrogen Production from Coal Industry Methane. Eurasian Chem.-Technol. J. 2022, 24, 69–91. [Google Scholar] [CrossRef]
  98. Sun, F.; Qi, H.; Xie, Y.; Ma, Q.; He, W.; Xu, D.; Wang, G.; Yu, W.; Wang, T.; Dong, X. Flexible Self-Supporting Bifunctional [TiO2/C]//[Bi2WO6/C] Carbon-Based Janus Nanofiber Heterojunction Photocatalysts for Efficient Hydrogen Evolution and Degradation of Organic Pollutant. J. Alloys Compd. 2020, 830, 154673. [Google Scholar] [CrossRef]
  99. Zhang, Y.; Niu, L.; Li, Z.; Yang, T.; Liu, Y.; Kang, Z. A Recyclable ZnIn2S4/PAN Photocatalytic Nanofiber Membrane for Boosting Visible Light Hydrogen Evolution in Seawater without Cocatalyst. Appl. Catal. B Environ. Energy 2024, 357, 124300. [Google Scholar] [CrossRef]
  100. Ismagilov, I.Z.; Matus, E.V.; Kuznetsov, V.V.; Yashnik, S.A.; Kerzhentsev, M.A.; Gerritsen, G.; Abbenhuis, H.C.; Ismagilov, Z.R. Application of POSS Nanotechnology for Preparation of Efficient Ni Catalysts for Hydrogen Production. Eurasian Chem.-Technol. J. 2017, 19, 3–16. [Google Scholar] [CrossRef]
  101. Sultanov, F.; Daulbayev, C.; Azat, S.; Kuterbekov, K.; Bekmyrza, K.; Bakbolat, B.; Bigaj, M.; Mansurov, Z. Influence of Metal Oxide Particles on Bandgap of 1D Photocatalysts Based on SrTiO3/PAN Fibers. Nanomaterials 2020, 10, 1734. [Google Scholar] [CrossRef]
  102. Chang, Y.-C.; Zeng, C.-J.; Chen, C.-Y.; Tsay, C.-Y.; Lee, G.-J.; Wu, J.J. NiS/Pt Loaded on Electrospun TiO2 Nanofiber with Enhanced Visible-Light-Driven Photocatalytic Hydrogen Production. Mater. Res. Bull. 2023, 157, 112041. [Google Scholar] [CrossRef]
  103. Peng, H.; Yong, J.; Wang, H.; Gou, Y.; Wang, F.; Zheng, X. Dual CdS–CoS/S,N-Doped TiO2 Nanofibers for Efficient Visible-Light Induced H2 Evolution. Int. J. Hydrogen Energy 2022, 47, 31269–31278. [Google Scholar] [CrossRef]
  104. Li, R.; Luan, J.; Zhang, Y.; Jiang, L.; Yan, H.; Chi, Q.; Yan, Z. A Review of Efficient Photocatalytic Water Splitting for Hydrogen Production. Renew. Sustain. Energy Rev. 2024, 206, 114863. [Google Scholar] [CrossRef]
  105. Abu-Sari, S.M.; Ang, B.C.; Wan Daud, W.M.A.; Patah, M.F.A. Visible-Light-Driven Photocatalytic Hydrogen Production on Defective, Sulfur Self-Doped g-C3N4 Nanofiber Fabricate via Electrospinning Method. J. Environ. Chem. Eng. 2023, 11, 109318. [Google Scholar] [CrossRef]
  106. Shang, Q.; Wang, J.; Yang, J.; Wang, Z.; Wang, G.; Wang, K.; Wu, X.; Li, J. Photocatalytic Hydrogen Evolution Coupled with Tetracycline Photodegradation over S-Scheme BaTiO3/Ag2S Dual-Function Nanofibers: Performance and Mechanism. Appl. Surf. Sci. 2023, 635, 157760. [Google Scholar] [CrossRef]
  107. Chang, Y.-C.; Syu, S.-Y.; Hsu, P.-C. Construction of In2S3–In(OH)3–ZnS Nanofibers for Boosting Photocatalytic Hydrogen Evolution. Int. J. Hydrogen Energy 2024, 86, 24–35. [Google Scholar] [CrossRef]
  108. Peng, H.; Luo, M.; Zheng, X. Boosting Solar-Light Photocatalytic Activity of Ni2P-Decorated ZnCr2O4 Nanofibers for H2 Evolution and Tetracycline Elimination. Ceram. Int. 2024, 50, 26474–26481. [Google Scholar] [CrossRef]
  109. Al-Enizi, A.M.; Karim, A.; Yousef, A. A Novel Method for Fabrication of Electrospun Cadmium Sulfide Nanoparticles- Decorated Zinc Oxide Nanofibers as Effective Photocatalyst for Water Photosplitting. Alex. Eng. J. 2023, 65, 825–835. [Google Scholar] [CrossRef]
  110. Aravind, M.; Amalanathan, M.; Aslam, S.; Noor, A.E.; Jini, D.; Majeed, S.; Velusamy, P.; Alothman, A.A.; Alshgari, R.A.; Saleh Mushab, M.S.; et al. Hydrothermally Synthesized Ag-TiO2 Nanofibers (NFs) for Photocatalytic Dye Degradation and Antibacterial Activity. Chemosphere 2023, 321, 138077. [Google Scholar] [CrossRef]
  111. Mousa, H.M.; Alenezi, J.F.; Mohamed, I.M.A.; Yasin, A.S.; Hashem, A.-F.M.; Abdal-hay, A. Synthesis of TiO2@ZnO Heterojunction for Dye Photodegradation and Wastewater Treatment. J. Alloys Compd. 2021, 886, 161169. [Google Scholar] [CrossRef]
  112. Villa, K.; Galán-Mascarós, J.R.; López, N.; Palomares, E. Photocatalytic Water Splitting: Advantages and Challenges. Sustain. Energy Fuels 2021, 5, 4560–4569. [Google Scholar] [CrossRef]
  113. Ahmed, M.A.; Mohamed, A.A. Recent Progress in Semiconductor/Graphene Photocatalysts: Synthesis, Photocatalytic Applications, and Challenges. RSC Adv. 2022, 13, 421–439. [Google Scholar] [CrossRef]
  114. Chen, Y.; Li, A.; Fu, X.; Peng, Z. Electrospinning-Based (N,F)-Co-Doped TiO2-δ Nanofibers: An Excellent Photocatalyst for Degrading Organic Dyes and Heavy Metal Ions under Visible Light. Mater. Chem. Phys. 2022, 291, 126672. [Google Scholar] [CrossRef]
  115. Yasin, A.S.; Mohamed, I.M.A.; Mousa, H.M.; Park, C.H.; Kim, C.S. Facile Synthesis of TiO2/ZrO2 Nanofibers/Nitrogen Co-Doped Activated Carbon to Enhance the Desalination and Bacterial Inactivation via Capacitive Deionization. Sci. Rep. 2018, 8, 541. [Google Scholar] [CrossRef] [PubMed]
  116. Le, P.H.; Hieu, L.T.; Lam, T.-N.; Hang, N.T.N.; Truong, N.V.; Tuyen, L.T.C.; Phong, P.T.; Leu, J. Enhanced Photocatalytic Performance of Nitrogen-Doped TiO2 Nanotube Arrays Using a Simple Annealing Process. Micromachines 2018, 9, 618. [Google Scholar] [CrossRef] [PubMed]
  117. Enculescu, M.; Costas, A.; Evanghelidis, A.; Enculescu, I. Fabrication of ZnO and TiO2 Nanotubes via Flexible Electro-Spun Nanofibers for Photocatalytic Applications. Nanomaterials 2021, 11, 1305. [Google Scholar] [CrossRef] [PubMed]
  118. Chen, G.; Wang, Y.; Shen, Q.; Xiong, X.; Ren, S.; Dai, G.; Wu, C. Fabrication of TiO2 Nanofibers Assembled by Bi2WO6 Nanosheets with Enhanced Visible Light Photocatalytic Activity. Ceram. Int. 2020, 46, 21304–21310. [Google Scholar] [CrossRef]
  119. Lou, C.-W.; Xie, M.-M.; Yang, Y.-D.; Wang, H.-Y.; Wang, Z.-K.; Zhang, L.; Hsieh, C.-T.; Liu, L.-Y.; Lin, M.-C.; Li, T.-T. Carbon Nanofiber Membranes Loaded with MXene@g-C3N4: Preparation and Photocatalytic Property. Nanomaterials 2024, 14, 896. [Google Scholar] [CrossRef]
  120. Liu, Y.; Hao, M.; Zhou, C.; Yang, B.; Jiang, S.; Huang, J.; Chen, Z.; Liu, Y.; Ramakrishna, S. Chapter 9—Scale-up Strategies for Electrospun Nanofiber Production. In Electrospun and Nanofibrous Membranes; Kargari, A., Matsuura, T., Shirazi, M.M.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2023; pp. 205–266. ISBN 978-0-12-823032-9. [Google Scholar]
  121. Sabzehmeidani, M.M.; Karimi, H.; Ghaedi, M.; Avargani, V.M. Construction of Efficient and Stable Ternary ZnFe2O4/Ag/AgBr Z-Scheme Photocatalyst Based on ZnFe2O4 Nanofibers under LED Visible Light. Mater. Res. Bull. 2021, 143, 111449. [Google Scholar] [CrossRef]
  122. Yi, S.; Sun, S.; Zhang, Y.; Zou, Y.; Dai, F.; Si, Y. Scalable Fabrication of Bimetal Modified Polyacrylonitrile (PAN) Nanofibrous Membranes for Photocatalytic Degradation of Dyes. J. Colloid Interface Sci. 2020, 559, 134–142. [Google Scholar] [CrossRef]
  123. Li, X.; Raza, S.; Liu, C. Directly Electrospinning Synthesized Z-Scheme Heterojunction TiO2@Ag@Cu2O Nanofibers with Enhanced Photocatalytic Degradation Activity under Solar Light Irradiation. J. Environ. Chem. Eng. 2021, 9, 106133. [Google Scholar] [CrossRef]
  124. Li, X.; Chen, D.; Li, N.; Xu, Q.; Li, H.; He, J.; Lu, J. Hollow SnO2 Nanotubes Decorated with ZnIn2S4 Nanosheets for Enhanced Visible-Light Photocatalytic Activity. J. Alloys Compd. 2020, 843, 155772. [Google Scholar] [CrossRef]
  125. Panthi, G.; Kwon, O.H.; Kuk, Y.-S.; Gyawali, K.R.; Park, Y.W.; Park, M. Ternary Composite of Co-Doped CdSe@electrospun Carbon Nanofibers: A Novel Reusable Visible Light-Driven Photocatalyst with Enhanced Performance. Catalysts 2020, 10, 348. [Google Scholar] [CrossRef]
  126. Zhang, L.; Shen, Q.; Zheng, S.; Yu, L.; Huang, F.; Zhang, C.; Sheng, J.; Yang, H. Direct Electrospinning Preparation of Z-Scheme Mixed-Crystal Bi2O3/g-C3N4 Composite Photocatalysts with Enhanced Visible-Light Photocatalytic Activity. New J. Chem. 2021, 45, 14522–14531. [Google Scholar] [CrossRef]
  127. Wang, Y.; Hao, X.; Wang, Z.; Dong, M.; Cui, L. Facile Fabrication of Mn2+-Doped ZnO Photocatalysts by Electrospinning. R. Soc. Open Sci. 2020, 7, 191050. [Google Scholar] [CrossRef] [PubMed]
  128. Pant, B.; Prasad Ojha, G.; Acharya, J.; Park, M. Ag3PO4-TiO2-Carbon Nanofiber Composite: An Efficient Visible-Light Photocatalyst Obtained from Electrospinning and Hydrothermal Methods. Sep. Purif. Technol. 2021, 276, 119400. [Google Scholar] [CrossRef]
  129. Xing, Y.; Cheng, J.; Li, H.; Lin, D.; Wang, Y.; Wu, H.; Pan, W. Electrospun Ceramic Nanofibers for Photocatalysis. Nanomaterials 2021, 11, 3221. [Google Scholar] [CrossRef]
  130. Pant, B.; Ojha, G.P.; Kuk, Y.-S.; Kwon, O.H.; Park, Y.W.; Park, M. Synthesis and Characterization of ZnO-TiO2/Carbon Fiber Composite with Enhanced Photocatalytic Properties. Nanomaterials 2020, 10, 1960. [Google Scholar] [CrossRef]
  131. Zhai, G.; Zhou, J.; Xie, M.; Jia, C.; Hu, Z.; Xiang, H.; Zhu, M. Improved Photocatalytic Property of Lignin-Derived Carbon Nanofibers through Catalyst Synergy. Int. J. Biol. Macromol. 2023, 233, 123588. [Google Scholar] [CrossRef]
  132. Yu, W.; Zheng, B.; Mao, K.; Jiang, J.; Luo, B.; Wu, X.; Tao, T.; Min, X.; Mi, R.; Huang, Z.; et al. Interfacial Structure and Photocatalytic Degradation Performance of Graphene Oxide Bridged Chitin-Modified TiO2/Carbon Fiber Composites. J. Clean. Prod. 2022, 361, 132261. [Google Scholar] [CrossRef]
  133. Zhang, J.-J.; Kai, C.-M.; Zhang, F.-J.; Wang, Y.-R. Novel PAN/Bi2MoO6/Ti3C2 Ternary Composite Membrane via Electrospinning with Enhanced Photocatalytic Degradation of Tetracycline. Colloids Surf. A Physicochem. Eng. Asp. 2022, 648, 129255. [Google Scholar] [CrossRef]
  134. Zhang, L.; Shen, Q.; Huang, F.; Jiang, L.; Liu, J.; Sheng, J.; Li, Y.; Yang, H. Electrospinning Directly Synthesis of 0D/1D CuBi2O4@WO3 Nanofiber Photocatalyst with S-Scheme Heterojunction. Appl. Surf. Sci. 2023, 608, 155064. [Google Scholar] [CrossRef]
  135. Fu, J.; Jiang, K.; Qiu, X.; Yu, J.; Liu, M. Product Selectivity of Photocatalytic CO2 Reduction Reactions. Mater. Today 2020, 32, 222–243. [Google Scholar] [CrossRef]
  136. Adekoya, D.; Tahir, M.; Amin, N.A.S. Recent Trends in Photocatalytic Materials for Reduction of Carbon Dioxide to Methanol. Renew. Sustain. Energy Rev. 2019, 116, 109389. [Google Scholar] [CrossRef]
  137. Kang, S.; Hwang, J. rGO-Wrapped Ag-Doped TiO2 Nanofibers for Photocatalytic CO2 Reduction under Visible Light. J. Clean. Prod. 2022, 374, 134022. [Google Scholar] [CrossRef]
  138. Wang, X.; Zhang, Z.; Huang, Z.; Dong, P.; Nie, X.; Jin, Z. Electrospun PVDF Nanofibers Decorated with Graphene and Titania for Improved Visible Light Photocatalytic Methanation of CO2. Plasmonics 2020, 15, 717–725. [Google Scholar] [CrossRef]
  139. Lu, N.; Zhang, M.; Jing, X.; Zhang, P.; Zhu, Y.; Zhang, Z. Electrospun Semiconductor-Based Nano-Heterostructures for Photocatalytic Energy Conversion and Environmental Remediation: Opportunities and Challenges. Energy Environ. Mater. 2023, 6, e12338. [Google Scholar] [CrossRef]
  140. Xu, F.; Tan, H.; Fan, J.; Cheng, B.; Yu, J.; Xu, J. Electrospun TiO2-Based Photocatalysts. Sol. RRL 2021, 5, 2000571. [Google Scholar] [CrossRef]
  141. Kang, S.; Im, T.; Koh, M.; Lee, C.S. Facile Fabrication of Electrospun Black Titania Nanofibers Decorated with Graphitic Carbon Nitride for the Application of Photocatalytic CO2 Reduction. J. CO2 Util. 2020, 41, 101230. [Google Scholar] [CrossRef]
  142. Song, J.; Guan, R.; Xie, M.; Dong, P.; Yang, X.; Zhang, J. Advances in Electrospun TiO2 Nanofibers: Design, Construction, and Applications. Chem. Eng. J. 2022, 431, 134343. [Google Scholar] [CrossRef]
  143. Ma, H.; Wu, X.; Li, X.; Liu, J.; Dong, H.; Liu, Y.; Niu, L.; Zhang, F.; Wang, W.; Shao, C.; et al. Photocatalytic CO2 Reduction to Ethanol by ZnCO2O4/ZnO Janus Hollow Nanofibers. Inorg. Chem. 2024, 63, 15735–15751. [Google Scholar] [CrossRef]
  144. Wei, Z.; Ji, T.; Zhou, X.; Guo, J.; Yu, X.; Liu, H.; Wang, J. Synergistic Enhancement of Photocatalytic CO2 Reduction by Built-in Electric Field/Piezoelectric Effect and Surface Plasmon Resonance via PVDF/CdS/Ag Heterostructure. Small 2023, 19, 2304202. [Google Scholar] [CrossRef]
  145. Wu, Y.; Wu, J.; Zhu, C.; Zhang, L.; Yan, J. MXene-Intercalation Induced Ordered Brick-Mortar Structures of Allomorph Junctions for Enhanced Flexibility in TiO2 Nanofibers and Photocatalytic Efficiency. Chem. Eng. J. 2023, 465, 142798. [Google Scholar] [CrossRef]
  146. Sun, Y.; Zhang, X.; Zhang, M.; Ge, M.; Wang, J.; Tang, Y.; Zhang, Y.; Mi, J.; Cai, W.; Lai, Y.; et al. Rational Design of Electrospun Nanofibers for Gas Purification: Principles, Opportunities, and Challenges. Chem. Eng. J. 2022, 446, 137099. [Google Scholar] [CrossRef]
  147. Wang, S.; Wang, L.; Wang, D.; Li, Y. Recent Advances of Single-Atom Catalysts in CO2 Conversion. Energy Environ. Sci. 2023, 16, 2759–2803. [Google Scholar] [CrossRef]
  148. Li, L.; ul Hasan, I.M.; Farwa1; He, R.; Peng, L.; Xu, N.; Niazi, N.K.; Zhang, J.-N.; Qiao, J. Copper as a Single Metal Atom Based Photo-, Electro-, and Photoelectrochemical Catalyst Decorated on Carbon Nitride Surface for Efficient CO2 Reduction: A Review. Nano Res. Energy 2022, 1, e9120015. [Google Scholar] [CrossRef]
  149. Zhu, L.; Qin, C.; Wang, Y.; Cao, J. Single-Atom Pt Supported on Non-Metal Doped WS2 for Photocatalytic CO2 Reduction: A First-Principles Study. Appl. Surf. Sci. 2023, 626, 157252. [Google Scholar] [CrossRef]
  150. Zhang, Z.; Wang, Z.; Cao, S.-W.; Xue, C. Au/Pt Nanoparticle-Decorated TiO2 Nanofibers with Plasmon-Enhanced Photocatalytic Activities for Solar-to-Fuel Conversion. J. Phys. Chem. C 2013, 117, 25939–25947. [Google Scholar] [CrossRef]
  151. Gong, S.; Hou, M.; Niu, Y.; Teng, X.; Liu, X.; Xu, M.; Xu, C.; Ka-Man Au, V.; Chen, Z. Molybdenum Phosphide Coupled with Highly Dispersed Nickel Confined in Porous Carbon Nanofibers for Enhanced Photocatalytic CO2 Reduction. Chem. Eng. J. 2022, 427, 131717. [Google Scholar] [CrossRef]
  152. Photocatalytic CO2 Reduction of C/ZnO Nanofibers Enhanced by an Ni-NiS Cocatalyst—Nanoscale (RSC Publishing). Available online: https://pubs.rsc.org/en/content/articlelanding/2020/nr/c9nr10451h (accessed on 11 September 2024).
  153. Shao, X.; Li, K.; Li, J.; Cheng, Q.; Wang, G.; Wang, K. Investigating S-Scheme Charge Transfer Pathways in NiS@Ta2O5 Hybrid Nanofibers for Photocatalytic CO2 Conversion. Chin. J. Catal. 2023, 51, 193–203. [Google Scholar] [CrossRef]
  154. Kang, S.; Khan, H.; Lee, C.; Kwon, K.; Sunyong Lee, C. Investigation of Hydrophobic MoSe2 Grown at Edge Sites on TiO2 Nanofibers for Photocatalytic CO2 Reduction. Chem. Eng. J. 2021, 420, 130496. [Google Scholar] [CrossRef]
  155. Prado, A.C.F.; Malafatti, J.O.D.; Oliveira, J.A.; Ribeiro, C.; Joya, M.R.; Luz, A.P.; Paris, E.C. Preparation and Application of Nb2O5 Nanofibers in CO2 Photoconversion. Nanomaterials 2021, 11, 3268. [Google Scholar] [CrossRef]
  156. Li, Y.; Ren, Z.; Gu, M.; Duan, Y.; Zhang, W.; Lv, K. Synergistic Effect of Interstitial C Doping and Oxygen Vacancies on the Photoreactivity of TiO2 Nanofibers towards CO2 Reduction. Appl. Catal. B Environ. 2022, 317, 121773. [Google Scholar] [CrossRef]
  157. Zhang, Z.; Chen, R.; Wang, L.; Chen, X.; Ding, J.; Zhang, J.; Wan, H.; Guan, G. Synergistic Effect of Cu2+ and Cu+ in SrTiO3 Nanofibers Promotes the Photocatalytic Reduction of CO2 to Methanol. Appl. Surf. Sci. 2023, 609, 155297. [Google Scholar] [CrossRef]
  158. Kang, S.; Khan, H.; Lee, C.S. CO2 Selectivity of Flower-like MoS2 Grown on TiO2 Nanofibers Coated with Acetic Acid-Treated Graphitic Carbon Nitride. Sol. Energy Mater. Sol. Cells 2021, 221, 110890. [Google Scholar] [CrossRef]
Figure 1. (a) Statistics from the Scopus database showing publication trends and keyword matches with the search string TITLE (“photocatalysis”) AND TITLE-ABS-KEY (“electrospinning”) from 2014 to 4 September 2024. (b) Distribution of publications by country.
Figure 1. (a) Statistics from the Scopus database showing publication trends and keyword matches with the search string TITLE (“photocatalysis”) AND TITLE-ABS-KEY (“electrospinning”) from 2014 to 4 September 2024. (b) Distribution of publications by country.
Molecules 29 04824 g001
Figure 2. Schematic diagram of a typical electrospinning set up for preparation of nanofibers (reproduced from [37] with permission of Elsevier, 2019).
Figure 2. Schematic diagram of a typical electrospinning set up for preparation of nanofibers (reproduced from [37] with permission of Elsevier, 2019).
Molecules 29 04824 g002
Figure 3. General mechanism of a photocatalytic reaction (reproduced from [58] with permission of Elsevier, 2020).
Figure 3. General mechanism of a photocatalytic reaction (reproduced from [58] with permission of Elsevier, 2020).
Molecules 29 04824 g003
Table 1. Recent research results (2022–2024) on the use of electrospun nanofiber-based photocatalysts in H2 production.
Table 1. Recent research results (2022–2024) on the use of electrospun nanofiber-based photocatalysts in H2 production.
YearPhotocatalystLight SourceSacrificial AgentH2 Evolution Rate (mmol h−1g−1) and AQYRef.
2023Sg-C3N4 nanofiberMetal halide 400 W, full spectrum20 vol.% methanol0.632[105]
2023TiO2/NiS/Pt nanofiber5 W blue LED light, λmax = 420 nm50 vol.% methanol4.411[102]
2024NiGa2O4/ZnIn2S4 nanofiber300 W Xe-lamp, AM 1.5 filter, 41.7 mW/cm210 vol.% TEOA9.292[93]
2024CoGa2O4/ZnIn2S4 nanofiber300 W Xe-lamp, AM 1.5 filter, 41.7 mW/cm210 vol.% TEOA6.283[93]
2022Cd0.5Co0.5S/SN-TiO2 nanofiber300 W Xe-lamp2.4 g Na2S, and 1.26 g Na2SO3 into 100 mL deionized H2O4.55 and AQY of 19.01% at 410 nm[103]
2024ZnIn2S4/PAN nanofiber membraneVisible light (420 nm ≤ λ ≤ 700 nm)10 vol.% TEOA1.836 and AQY of 1.77% at 365 nm[99]
2023S-scheme BaTiO3/Ag2S nanofiber300 W Xe-lampNa2S (0.35 mol/L) and Na2SO3 (0.25 mol/L)0.597[106]
2024In2S3–In(OH)3–ZnS nanofibers5 W blue LED light (λmax = 420 nm, 41.7 mW cm−2)0.1 M Na2S solution0.2236[107]
2024C–Ni2P/ZnCr2O4 nanofibersXe lamp intensity of 350 mW cm−20.2 g Na2S, and 0.2 g Na2SO3 into 100 mL deionized H2O0.5759 and AQY of 15.25% at 420 nm[108]
2023CdS NPs-decorated ZnO nanofibers500 W Xe lamp with 425 nm band pass filter0.35 M Na2S and 0.25 M Na2SO30.820[109]
Table 3. Recent research results (2020–2024) on the use of electrospun nanofiber-based photocatalysts for CO2 reduction into value-added products.
Table 3. Recent research results (2020–2024) on the use of electrospun nanofiber-based photocatalysts for CO2 reduction into value-added products.
Photocatalytic NanofibersLight SourceReagentProductsReaction RateRef.
g-C3N4/black titania300 W Xe-arc lampCO2 + H2O + TEOACO and CH45.19 and 1.65 μmol/g[141]
(rGO)-wrapped Ag/TiO2500 W Xe lamp with a 400-nm long pass filterCO2 + H2O vaporCH44.301 μmol g−1[137]
Ni-NiS/C/ZnO350 W simulated solar Xe arc lamp, 10,117 μW cm−2CO2 + H2O + NaHCO3CO and CH45.86 and 1.14 μmol g−1 h−1[152]
NiS@Ta2O5Xe lamp, 920 mW cm−2CO2 + H2OCO and CH443.27 and 6.56 μmol g−1 h−1[153]
TiO2/MoSe2300 W Xe-arc lamp, 12 mW/cm2CO2 + H2O + TEOACH4 and CO174.02 and 478.46 μmol/g[154]
Nb2O518 W mercury lamp, 254 nmCO2 + H2O vaporCO and CH48.5 and 0.55 μmol g−1[155]
Ni-MoP@NCPF300 W Xe lamp with a UVCUT 420-nm filterCO2 + acetonitrile/H2O + TEOACO953.33 μmol g−1h−1[151]
C doped TiO2300 W Xe lamp, AM 1.5 filterCO2 + H2O + NaHCO3 + H2SO4CH455.17 μmol g−1 h−1[156]
SrTi1-xCuxO3-H2300 W Xe lamp, (400 nm < λ < 780 nm)CO2 + H2OCH3OH5.38 μmol g−1 h−1[157]
Graphene@PVDF@TiO2Two 300 W visible light sources (UV < 5%)CO2 + H2OCH4363 μmol g−1 h−1[138]
TiO2/MoS2/g-C3N300 W Xe-arc lamp, 12 mW/cm2CO2 + H2O + TEOACH421.78 μmol g−1[158]
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

Serik, A.; Idrissov, N.; Baratov, A.; Dikov, A.; Kislitsin, S.; Daulbayev, C.; Kuspanov, Z. Recent Progress in Photocatalytic Applications of Electrospun Nanofibers: A Review. Molecules 2024, 29, 4824. https://doi.org/10.3390/molecules29204824

AMA Style

Serik A, Idrissov N, Baratov A, Dikov A, Kislitsin S, Daulbayev C, Kuspanov Z. Recent Progress in Photocatalytic Applications of Electrospun Nanofibers: A Review. Molecules. 2024; 29(20):4824. https://doi.org/10.3390/molecules29204824

Chicago/Turabian Style

Serik, Aigerim, Nurlan Idrissov, Aibol Baratov, Alexey Dikov, Sergey Kislitsin, Chingis Daulbayev, and Zhengisbek Kuspanov. 2024. "Recent Progress in Photocatalytic Applications of Electrospun Nanofibers: A Review" Molecules 29, no. 20: 4824. https://doi.org/10.3390/molecules29204824

APA Style

Serik, A., Idrissov, N., Baratov, A., Dikov, A., Kislitsin, S., Daulbayev, C., & Kuspanov, Z. (2024). Recent Progress in Photocatalytic Applications of Electrospun Nanofibers: A Review. Molecules, 29(20), 4824. https://doi.org/10.3390/molecules29204824

Article Metrics

Back to TopTop