Next Article in Journal
Modification of Graphite Sheet Anode with Iron (II, III) Oxide-Carbon Dots for Enhancing the Performance of Microbial Fuel Cell
Next Article in Special Issue
Synthesis and Application of Innovative and Environmentally Friendly Photocatalysts: A Review
Previous Article in Journal
Optimization of Plasmonic Copper Content at Copper-Modified Strontium Titanate (Cu-SrTiO3): Synthesis, Characterization, Photocatalytic Activity
Previous Article in Special Issue
Photocatalytic CO2 Conversion Using Anodic TiO2 Nanotube-CuxO Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advancements in Photocatalysis Coupling by External Physical Fields

1
School of Electrical and Energy Power Engineering, Yangzhou University, Yangzhou 225002, China
2
Research Center for Combustion and Environment Technology, Shanghai Jiao Tong University, Shanghai 200240, China
3
Institute of Solar Energy, Shanghai University of Electric Power, Shanghai 200090, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(9), 1042; https://doi.org/10.3390/catal12091042
Submission received: 19 August 2022 / Revised: 6 September 2022 / Accepted: 9 September 2022 / Published: 13 September 2022
(This article belongs to the Special Issue 10th Anniversary of Catalysts—Feature Papers in Photocatalysis)

Abstract

:
Photocatalysis is one of the most promising green technologies to utilize solar energy for clean energy achievement and environmental governance, such as artificial photosynthesis, water splitting, pollutants degradation, etc. Despite decades of research, the performance of photocatalysis still falls far short of the requirement of 5% solar energy conversion efficiency. Combining photocatalysis with the other physical fields has been proven to be an efficient way around this barrier which can improve the performance of photocatalysis remarkably. This review will focus on the recent advances in photocatalysis coupling by external physical fields, including Thermal-coupled photocatalysis (TCP), Mechanical-coupled photocatalysis (MCP), and Electromagnetism-coupled photocatalysis (ECP). In this paper, coupling mechanisms, materials, and applications of external physical fields are reviewed. Specifically, the promotive effect on photocatalytic activity by the external fields is highlighted. This review will provide a detailed and specific reference for photocatalysis coupling by external physical fields in a deep-going way.

1. Introduction

In order to achieve the carbon emission reduction target, photocatalysis has been widely considered to be a promising technology. After half a century of development, photocatalysis has tended to increasingly maturate, especially in many fields, such as pollutant degradation, water splitting, CO2 reduction, and desulfurization [1,2,3,4,5,6,7,8,9,10,11]. However, the main reason limiting photocatalytic technology from basic research to practical application is that the conversion efficiency of solar energy is still very low, which is far from meeting the requirements of commercial applications [12,13]. At present, there are two main ways to break the bottleneck of the efficiency of photocatalytic energy conversion: one is to enhance the absorption rate of the solar spectrum and the other is to efficiently promote the separation of photogenerated carriers.
Techniques for modification of photocatalytic materials, such as doping, compositing other materials, cocatalyst loading, morphology controlling, etc., have been confirmed to improve light absorption and photogenerated carrier separation efficiently [14,15,16,17,18]. However, the energy conversion efficiency is still very low although it has been improved to some extent by these modifications. Therefore, it is necessary to find a new way to achieve a breakthrough in photocatalytic performance. Studies have shown that photocatalytic activity can be greatly improved by adding other physical fields to the photocatalytic reaction process [19,20]. Generally, the assisted external field can provide additional energy to the photocatalytic system, which will power photogenerated charge separation and thereby improve the efficiency of photocatalysis. According to the energy form, as shown in Figure 1 the assisted physical field can be divided into the thermal field, mechanical field, and, electromagnetic field. In terms of energy generation types, it can also be divided into infrared thermal effect, microwave thermal effect, ultrasonic cavitation effect, friction electrostatic effect, external controllable electromagnetic field effect, and so on. One of the most common assisted physical fields is the thermal field. For example, in the photocatalytic reaction, most of the infrared light in the sunlight cannot be effectively used, which will dissipate into heat. Since infrared light and heat are inseparable, the infrared absorbing material can be composited into the photocatalytic material to collect the infrared band energy promoting the separation of photogenerated carriers. In addition, other non-contact external fields, such as the electric field, magnetic field, microwave field, and ultrasound field, can significantly improve the efficiency of the photocatalytic system, which have different mechanisms and respective advantages.
Therefore, we will provide a detailed and specific reference for photocatalysis coupling by external physical fields, including thermal-coupled photocatalysis (TCP), mechanical-coupled photocatalysis (MCP), and electromagnetism-coupled photocatalysis (ECP). This paper will mainly focus on coupling mechanisms, materials, and applications of external physical fields. Specifically, the promotive effect on photocatalytic activity by the external fields will be highlighted.

2. Thermal-Coupled Photocatalysis (TCP)

To break through the existing bottleneck of photocatalytic activity, thermal energy has been introduced into photocatalysis systems. Generally, traditional photothermal catalysis can be divided into three types [19,21,22,23,24,25,26,27]: (1) thermal-coupled photocatalysis (TCP); (2) photo-coupled thermocatalysis (PCT); and (3) photothermal co-catalysis (PTC). As for PCT, light is not the main driving force of the reaction, which is used only to provide heat and promote the thermocatalytic process. Therefore, we will focus on the TCP and PTC, where light is the main driving force for the reaction. Compared with conventional photocatalysis, TCP can combine the advantages of thermal and light energy, thereby greatly improving photocatalytic activity. The mechanism of efficiently photocatalytic activity for TCP can be summarized as follows: (1) thermal energy may help generate electron-hole pairs; (2) thermal energy can re-emit the trapped electrons and drive them rapidly for transfer to the surface of the photocatalyst; (3) thermal energy can increase the collision rate with more efficient collisions of molecules; and (4) the high surface temperature of photocatalyst will promote the dissociative adsorption of reactant molecules. Here, we will try to summarize the current research progress of TCP including the primary source of thermal energy, materials for TCP and, the main applications of TCP.

2.1. The Primary Source of Thermal Energy

2.1.1. External Direct Heating

The simplest method to realize TCP is to introduce an external heating process through traditional heat convection, heat conduction, and heat radiation. Increasing the external temperature can inhibit the recombination of photogenerated carriers and reduce the charge transfer resistance, thereby increasing the photocatalytic reaction rate. Huang et al. synthesized Co0.85Se/Mn0.3Cd0.7S nanorods for photocatalytic H2 production [28]. As shown in Figure 2a, the catalyst itself has no thermal catalytic activity. The H2 production was improved with the increase in temperature. Jing et al. reported a Cu2O-rGO/TiO2 photocatalyst whose H2 evolution rate (17,800 μmol g−1 h−1) under the photothermal condition (90 °C) is 4.7 times higher than that under only photocatalytic condition (25 °C), as shown in Figure 2b [29]. The photothermal effect facilitates the separation of photogenerated electron-hole pairs and simultaneously accelerates the reaction rate on the catalyst surface during H2 evolution. Li et al. synthesized a low-cost Cu2O/g-C3N4 heterojunction by a simple hydrothermal method and realized efficient photothermal conversion of CO2 into ethanol [30]. As shown in Figure 2c, under photothermal catalysis, the rate of ethanol is 1.89 times that of photocatalysis and 7.05 times that of thermocatalysis at 100 °C. The temperature enhanced the thermal motion of the radical, which promotes the formation of ethanol through the CH3 dimerization.

2.1.2. Near-Infrared Indirect Heating

Near-infrared indirect heating produces the photothermal effect by photothermal materials. Under near-infrared light irradiation, the absorbed photons could excite the electrons, which would migrate until they recombine with the holes. Therefore, there is an essential distinction between TCP via external direct heating and near-infrared indirect heating in terms of the way to enhance the photocatalytic activity. The photothermal effect can increase the temperature in a local area to induce a collective effect of multistage processes, which include charge excitation, charge carrier mobility, and reaction barriers. Photothermal material SnSe nanoparticles that serve as “hot islands” on ZnIn2S4 monolayer were proposed as an ideal model for investigating the role of the photothermal contribution in photocatalysis [31]. The energetic photoexcited electrons of nanoparticle SnSe interact with the lattice to transfer their energy to phonons for photothermal conversion. Therefore, SnSe acts as a “hot island” to release thermal energy, which increases the temperature around it. Then, the increased temperature in the local area that is caused by the photothermal effect can induce a collective effect to accelerate the photocatalytic reaction of ZnIn2S4 to enhance the photocatalytic performance, as shown in Figure 3. The core-shell structured WO3/CdS photocatalyst has been reported for full-spectrum solar energy conversion. The wrapped WO3 with abundant oxygen vacancy can store electrons activated by UV-visible light, which can then be released by NIR light. The solar-to-hydrogen conversion efficiency under simulated solar light irradiation reaches as high as 3.00% [32]. Non-precious metal co-catalysts S vacancies 1T-WS2, CuS, and plasmonic nanoparticles could also facilitate the driving of the photothermal effect [33,34].

2.1.3. Microwave Indirect Heating

The effect of microwave irradiation is a combination of thermal effects and non-thermal microwave effects, arising from the heating rate, super-heating or ‘‘hot spots’’, and the selective absorption of radiation by polar substances [35,36]. Therefore, it is of great practical significance to use the coupling of microwave and UV light to improve the efficiency of photocatalytic activity. The special thermal effect generated by microwave field is different from external direct heating and near-infrared indirect heating, which shows quickly increase temperature and more uniform temperature in solution. Ling et al. have successfully prepared uniformly Ag-loaded TiO2 for photocatalytic degradation of 4-CP and MO coupling by microwave irradiation [37]. The improved photocatalytic degradation activity of Ag/TiO2 was due to the thermal effect and non-thermal effect. Under microwave irradiation, silver nanoparticles can not only promote the charge separation process but also enhance the adsorption of organic pollutants on the surface of TiO2, as shown in Figure 4. Wang et al. reported a novel Z-scheme SrTiO3/MnFe2O4 nanocomposite for tetracycline (TC) degradation under MW-UV exposure [38]. It has been found that the combination of MW catalytic and UV photocatalytic oxidation technologies achieves superior SrTiO3/MnFe2O4 catalytic performance. The Z-scheme SrTiO3/MnFe2O4 system displays high catalytic activity under MW-UV. Particularly, the generated MW plasma accelerates electron transfer and separation of e-h+ pair, which promotes the enhancement of redox capacity and the oxidation of organics. Gayathri et al. found that the simultaneous irradiation by MW and UV is more efficient and synergistic for the mineralization of the dye compared to the respective individual energy sources [39]. The synergistic enhancement in the photocatalytic (UV/ZnO) degradation, as well as mineralization of rhodamine B (RhB) by MW, is attributed to the increased formation of reactive oxygen species (ROS) and the formation of surface defects in ZnO, which inhibits the recombination of photogenerated electrons and holes. MW can also enhance the photocatalytic degradation and mineralization of RhB even in systems with less dissolved oxygen because of the efficient transference of oxygen from the bulk of the catalyst to the surface.

2.2. Materials of Thermal-Coupled Photocatalysis

In thermal-coupled photocatalysis, there are three primary sources of thermal energy: external direct heating, near-infrared indirect heating, and microwave indirect heating. Generally, external direct heating and microwave indirect heating have nothing special requests for the materials. Here, we will focus on the near-infrared indirect heating materials which show a photothermal effect irradiated by near-infrared light. The surface temperature of materials will increase locally by converting solar energy into heat energy [22,23,24,40]. According to different photothermal conversion mechanisms (as shown in Figure 5), near-infrared light can be converted to heat by three categories of materials: (1) Metallic materials with localized surface plasmon effect; (2) Narrow-band semiconductors with non-radiative relaxation; and (3) Carbon-based materials with thermal vibration. Therefore, TCP can be realized by near-infrared indirect heating through the composite material of photocatalysts and near-infrared absorption materials.

2.2.1. Metallic Materials with Localized Surface Plasmon Effect

Under illumination, when the frequency of the light matches the resonant frequency of conduction-band electrons, electromagnetic waves can induce collective oscillations of electrons at the interface between a conducting material and a dielectric medium, such as air or water. Because it is very localized, this phenomenon is called localized surface plasmon resonance (LSPR) [41,42,43]. LSPR excitation can lead to a significant enhancement of the local electric field and a high concentration of energetic (hot) electrons on the surface of the plasmonic structure, a process that leads to a rapid increase in the local surface temperature of the materials. The excited hot carriers then distribute the thermal charge through electron–electron or electron–phonon scattering. Therefore, LSPR involves three distinct processes: enhancement of the local electric field, generation of hot electrons, and photothermal conversion.
The LSPR effect has been observed in many metal nanoparticles (NPs) and their alloys, such as Au, Ag, Cu, AuCu, and AuAl, etc. [44,45,46,47,48,49,50,51,52]. Schatz et al. investigated the localized surface plasmon resonance (LSPR) of Cu NPs [44]. It was found that Cu shows strong and narrow LSPR peaks comparable to Ag and Au. Yan et al. reported using plasmonic Au−Pd nanostructures to directly capture visible to near-infrared light for chemical reactions [45]. The Au NPs are tightly bound to the Pd NPs for efficient light harvesting. Under plasmonic excitation, catalytic reactions are induced and accelerated through plasmonic photocatalysis and photothermal conversion. Under 809 nm laser irradiation at 1.68 W, the yield of the Suzuki coupling reaction was two times higher than when thermally heated to the same temperature. Liu et al. developed embedded coupled metal nanostructures, which proved to be an effective strategy to improve the utilization of LSPR effects (as shown in Figure 6) [46]. Compared with traditionally isolated Au nanoparticles, the contribution of Au nano chains to the enhanced light conversion rate is increased by a factor of 3.5. Ma et al. developed a series of Au-Al intermetallic compounds with a high LSPR effect [47]. The results show that the hot carrier generation rate of AuAl2 and AuAl is higher than that of Au, and the energy rate of the resulting hot carriers is higher than that of Au. At the same time, the transport performance of hot carriers in AuAl2 and AuAl is also better than that of Au.
Besides metal nanoparticles, some semiconductor materials with metallic phases are also candidates for LSPR, such as CuS, Ti3C2, TiN, WO3-x, etc. [42,53,54,55,56,57,58,59,60]. Dai et al. demonstrated the reversible displacement of surface-dependent localized surface plasmon resonances (LSPR) in CuS nanodisks [53]. As shown in Figure 7, LSPR curves (curves 1–4) shift with increasing surface-treatment cycles. Tian et al. prepared unique Cu7S4-Cu9S8 heterostructured hollow cubes with remarkable photothermal effects [55]. It is shown that the strong near-infrared plasmonic absorption of Cu7S4-Cu9S8 promotes the hot carrier transport between Cu7S4 and Cu9S8, thereby promoting the catalytic reaction. Shi et al. successfully synthesized biocompatible Ti3C2 nanosheets (MXenes) for cancer therapy [56]. Due to the LSPR effect of the semi-metallic Ti3C2 nanosheets, the ultrathin Ti3C2 nanosheets have strong absorption and conversion efficiency for NIR laser irradiation (808 nm). Naik et al. demonstrated that TiN is a plasmonic material in visible and near-infrared regions [57]. The research results show that TiN has comparable performance to Au in plasmonic applications, and is significantly better than gold and silver in conversion optics and other aspects. Manthiram et al. demonstrated that the metallic phase of WO3-x nanoparticles possesses strong and tunable localized surface plasmon resonance [42]. Xue et al. prepared Pd-loaded plasmonic WO3-x nanowires and demonstrated that the plasmon excitation of WO3-x nanowires by long wavelength irradiation can greatly promote Suzuki coupling reactions catalyzed by the attached Pd nanoparticles [58].

2.2.2. Narrow-Band Semiconductors with Non-Radiative Relaxation

Unlike plasmonic materials, some narrow-band semiconductors generate heat through the nonradiative relaxation of electron-hole pairs. Under illumination, when the semiconductor receives energy greater than or equal to its forbidden bandwidth, electron-hole pairs are generated inside the semiconductor. Afterwards, the excited electrons and holes can relax to lower energy levels and release energy in the form of photons. Phonon relaxation leads to localized heating of the lattice and constitutes a temperature distribution that depends on its optical absorption [22,40,61].
TiO2 is a typical representative of this photothermal conversion mechanism. Huang et al. report a large-scale production method for the synthesis of black TiO2 by aluminum reduction, and the obtained samples have a unique crystalline core-amorphous shell structure (TiO2@TiO2-x) [62]. Under the illumination of the AM1.5G Xe lamp solar simulator, the temperature of TiO2@TiO2-x increases, which leads to more electronic excitation and relaxation and enhanced thermal emission. Black TiO2 absorbs approximately 65% of the total solar energy through improved visible and infrared absorption. Ozin et al. synthesized black TiOx (x < 2) by reducing white P25 TiO2 nanocrystals with magnesium (Mg) [63], as shown in Figure 8a. The research results show that under the simulated illumination of 1 solar intensity, black TiOx can effectively generate water vapor, and its photothermal conversion efficiency is as high as 50%. Yu et al. showed a slow photon effect in TiO2 photonic crystals (TiO2 PCs) with extraordinary photothermal effects that outperform commercial TiO2 and TiO2 nanotube arrays for photocatalytic methane production, as shown in Figure 8b [64]. Liu et al. proposed a novel diatom-like photothermal material using TiO2 thin-film bilayer melamine foam (TiO2-PANi@MF) for solar-driven clean water production [65]. The material can achieve full-spectrum light absorption and photothermal conversion. The evaporation rate of 2.12 kg m−2 h−1 was achieved under 1 sun irradiation, and the solar steam conversion efficiency was as high as 88.9%.
In addition to TiO2, some semiconductor materials with defects also have near-infrared light-to-heat conversion properties. Chou et al. reported the NIR-photothermal properties of chemically exfoliated MoS2 (ceMoS2) [66]. The absorbance of ceMoS2 in NIR is approximately 7.8 times higher than that of graphene oxide (GO). Its extinction coefficient at 800 nm is 29.2 L g−1 cm−1, which is higher than that of gold nanorods. Jun et al. have demonstrated for the first time that chemically exfoliated MoS2 can be an efficient photothermal material for the creation of fresh water in solar evaporators [67]. Due to the efficient photothermal conversion of ce-MoS2 nanosheets, the solar evaporation efficiency can reach 76% under the light intensity of 0.76 kW/m2. Guo et al. proposed oxygen-deficient molybdenum oxide quantum dots (MoO3-x QDs), which have an absorption spectrum matching that of sunlight in both the visible and near-infrared regions, resulting in a good interfacial heating effect on water evaporation [68]. As shown in Figure 9a, theoretical models suggest that the unique photothermal properties of MoO3-x QDs are caused by the oxygen defect level rather than the LSPR effect. Qi et al. proposed a new route to activate inert In(OH)3 into 2D black In2O3-x nanosheets via photoinduced defect engineering, as shown in Figure 9b [69]. Due to a large number of oxygen vacancies on the layers, the black In2O3-x nanosheets have a strong photothermal conversion ability. In the photothermal catalytic water gas shift experiment (RWGS), the CO production rate is as high as 103.21 mmol g−1 h−1.

2.2.3. Carbon-Based Materials with Thermal Vibration

Carbon-based materials have always been a research hotspot in the field of catalysis due to their excellent chemical stability and high light absorption [70,71]. Carbon-based materials possess abundant conjugated π-bonds and thus can convert solar energy into thermal energy through thermal vibration mechanisms. Under sunlight illumination, a large number of conjugated π bonds require only a small amount of energy to excite loosely held electrons, followed by electron relaxation via electron–electron and electron–phonon scattering. This causes the molecules to generate heat, which in turn causes the macroscopic heating of the material [72,73,74].
Carbon nanotubes (CNTs) are a classic carbon-based photothermal material. Using carbon nanotubes (CNTs) to modify flexible wood films (F-Wood/CNTs), Hu et al. created an efficient solar steam generation device [75]. Benefiting from the excellent light absorption of the black CNT coating, as shown in Figure 10a, its solar steam power generation device exhibits a high efficiency of 81% at 10 kW cm−2. An all-carbon nanotube (CNT) hybrid film was prepared by Jia et al. The hybrid film has ultra-high solar absorbance, low thermal conductivity, and excellent light-to-heat conversion capability [76]. The photothermal conversion efficiency of a single hybrid film is 87.4% under 1 sun illumination. Zhang et al. used SiO2 nanoparticles to graft carbon nanotubes (CNTs) to prepare CNTs-SiO2 superhydrophobic coatings [77]. Due to the superhydrophobicity of the coating surface and the remarkable photothermal conversion ability of CNTs, the CNTs-SiO2 coating can significantly delay the water freezing time and effectively melt the ice layer within seconds under laser irradiation.
Graphene, including graphene oxide (GO) and reduced graphene oxide (rGO), has long been known for its powerful photothermal conversion capabilities. Dai et al. developed nanosheets of reduced graphene oxide (nano-rGO) with high near-infrared (NIR) absorbance, as shown in Figure 10b [78]. The single-layered nano-rGO sheets have an average lateral dimension of ∼20 nm, and their NIR absorption is six times higher than that of non-reduced covalent PEGylated nano-GO. Jiang et al. synthesized two-dimensional rGO and one-dimensional multi-walled CNTs (MWCNTs) as rGO-MWCNT-based films to enhance photothermal performance [79]. The surface of the rGO-MWCNT-based film is rough, thereby reducing the reflection of light and thereby realizing almost all light trapping. Under 1 sun illumination, the surface temperature is as high as 78 °C, which is 10 °C higher than the result of pure rGO film. The solar thermal conversion efficiency can reach 80.4%. Qiu et al. prepared a reduced graphene oxide-based (rGO-based) covalent organic framework hydrogel (KTG) with a three-dimensional porous structure as a platform for enhanced uranium capture via photothermal desalination [80]. KTG can obtain a special uranium capture capacity (521.6 mg g−1) under one sunlight exposure, which is 42.4% higher than that under dark conditions.
Polymers, such as polypyrrole (PPy), and polydopamine (PDA), are also very powerful in photothermal conversion. Dai et al. constructed uniform PPy nanoparticles [81]. As shown in Figure 10c, the PPy nanoparticles have good colloidal stability and strong NIR absorption, and their photothermal conversion efficiency is better than that of the well-known Au nanorods. Jun et al. demonstrated a polydopamine (PDA)-coated polyvinylidene fluoride (PVDF) photothermal membrane for solar-driven membrane distillation [82]. The film has remarkable light absorption and excellent light-to-heat conversion performance. Under the irradiation of 0.75 kW/m2, its distillation performance efficiency is 45%. Xu et al. improved light-harvesting efficiency over a wide spectral and angular range by exploiting self-organized surface structures on large-area multilayer polypyrrole (PPy) nanosheets [83]. The solar heat conversion efficiency is as high as 95.33%. In addition, when used for solar steam power generation, the measurement efficiency can reach 92% under 1 sun illumination. Chen et al. developed a porous framework photothermal microgroove-structured aerogel (PDA/PEI/PPy@PI-MS MGA, pppMGA) [84]. Combining the porous micro-grooved structure of the pppMGA evaporation interface with the light absorption ability of PPy, the light absorption rate (98%) was effectively improved. Under 1 sun illumination, the pppMGA evaporator can achieve a high evaporation rate (∼1.38 kg m−2 h−1) and a high photothermal conversion efficiency (∼93.04%).

2.3. Applications of Thermal-Coupled Photocatalysis

Thermal-coupled photocatalysis has applications in many fields, including artificial photosynthesis, water splitting, pollutant degradation, etc. The thermal-coupled photocatalysis achievements in recent years are summarized and listed in Table 1.

2.3.1. Artificial Photosynthesis

Photosynthesis is one of the most important reactions on Earth. Artificial photosynthesis attempts to replicate natural photosynthesis, thereby achieving the goal of reducing CO2. Artificial photosynthesis can simply be defined as the process of converting CO2 and water into carbon-based fuels under sunlight [141,142,143]. Products from CO2 reduction include CH4, CO, methanol, methane, etc. However, the single photocatalytic method suffers from disadvantages, such as low charge separation efficiency, poor light-harvesting ability, and less reduction selectivity. In contrast, thermal-coupled photocatalysis couples solar and thermal energy, providing a more reliable solution.
Combining the photocatalytic properties of TiO2 with the thermocatalytic properties of CeO2, Fiorenza et al. synthesized TiO2-CeO2-based catalysts for artificial photosynthesis [89]. Furthermore, the addition of Co-Cu oxide increases the number of oxygen vacancies, which facilitates the separation of carriers. As shown in Figure 11, under the combined action of thermal energy and solar energy, the CO and CH4 yields can reach 12 μmol g−1 h−1 and 1.84 μmol g−1 h−1, respectively. Xuan et al. designed SrTiO3/Cu@Ni/TiN catalysts for CO2 reduction to ethanol [85]. Non-noble metallic materials, such as Cu, Ni, and TiN were used to enhance the plasmonic effect for efficiently capturing solar full-spectrum energy. The optimal ethanol evolution rate is 21.3 μmol g−1 h−1, and the ethanol selectivity is 79%. Lou et al. constructed a plasmonic semiconductor heterostructure photocatalyst with pyroelectric black phosphorus (BP) and plasmonic tungsten oxide (WO) for artificial photosynthesis [86]. Under the irradiation of visible light and NIR light, the plasmonic thermal effect of WO can increase the local temperature to 86 °C, triggering the pyroelectric effect of BP, generating pyroelectric carriers, and enhancing the electron transfer from BP to WO. Thus, the photothermal catalytic performance is enhanced. Its CO generation reaches 26.1 μmol g−1 h−1, which is 7 times and 17 times that of plasma WO and pyroelectric BP, respectively. Xu et al. developed a novel lead-free perovskite Cs3Sb2I9 photocatalyst to efficiently reduce CO2 to CO and methane via photothermal co-catalysis [92]. Its optimal production rate is 95.7 μmol g−1 h−1, which is 87 times and 5.2 times higher than that of pure photocatalysis and pure thermal catalysis, respectively. The study found that the defective active sites generated by thermal excitation of Cs3Sb2I9 photocatalysts can adsorb and activate CO2, while the large number of electrons excited by light can provide CO2 reduction ability, and their synergistic effect enhances the catalytic activity. Cui et al. reported ultrathin porous g-C3N4 nanosheets decorated with AuCu alloy NPs for photothermal catalytic CO2 reduction to ethanol [49]. The strong interaction between metal and g-C3N4 accelerates the transfer of photogenerated charges. In addition, the increase in temperature leads to enhanced thermal motion of molecules, leading to a synergistic effect of photocatalysis and thermocatalysis. At 120 °C, the yield and selectivity of ethanol is 0.89 mmol g−1 h−1, which is 4.2 times higher than that of photocatalysis and 7.6 times higher than that of thermal catalysis.

2.3.2. Water Splitting

To effectively solve the consumption of fossil fuels and the serious environmental problems that accompany combustion, modern society has been looking for a clean, renewable, and cheap energy production method. Solar-driven photocatalytic water splitting for H2 production is a potentially effective solution to the energy crisis, especially with few environmental concerns [144,145,146]. However, there are widespread problems in photocatalytic water splitting for H2 production, including low solar energy utilization efficiency, poor response to visible and infrared light, and rapid recombination of photogenerated electrons and holes. To improve these problems, introducing thermal energy into a photocatalytic H2 production system has become a research hotspot.
Li et al. prepared Ag-polydopamine nanoparticles co-decorated defective mesoporous carbon nitride nanosheet assemblies (Ag-PDA/DCN) [96]. The photocatalytic H2 evolution rate of this material is as high as 3840 μmol g−1 h−1, driven by visible and near-infrared light, which is seven times higher than that of PDA/DCN. This is mainly due to the rapid electron transfer achieved by Ag-PDA nanoparticles, which further expands the photoresponse of DCN, resulting in an obvious photothermal effect, which further enhances the photocatalytic activity. Yu et al. used an S-vacancy 1T-WS2 (Vs-1T-WS2) as a photothermal co-catalyst and complexed it with CdS to promote photocatalytic H2 evolution [33]. The catalyst can achieve H2 evolution performance of 70.9 mmol g−1 h−1 at 500 nm with an apparent quantum yield (AQY) of 39.1%. The introduction of Vs-1T-WS2 not only expanded the light absorption range but also accelerated the directional migration of electrons from CdS to S vacancies through the induced photothermal effect. Guo et al. proposed an efficient two-phase photocatalytic H2 production system [103]. The system uses a charred wood substrate to convert liquid water into water vapor and split the water vapor into H2 in the presence of light without additional energy. The particle photocatalytic system based on the wood/CoO system achieved H2 production rate as high as 220.74 μmol h−1 cm2, indicating that the photothermal-photocatalytic biphasic system is cost-effective and beneficial for practical applications. Ho et al. have successfully synthesized Ag@SiO2@TiO2/Au catalysts with dual effects of enhanced charge transfer and photothermal heating for photo-splitting water for H2 production (as shown in Figure 12) [104]. The catalyst has an extremely high H2 generation rate of 30.2 mmol g−1 h−1. Ag@SiO2 absorbs visible light for efficient photothermal conversion, thereby promoting the increase in the number of active sites for photocatalytic reactions. The Au Nps reduced the charge recombination in TiO2 and further enhanced the photocatalytic performance. Jiang et al. reported the incorporation of single-atom silver into g-C3N4 (SAAg-g-CN) as a low-cost and efficient photocatalyst [101]. The unique coordination of SAAg-g-CN can shorten the electron transfer pathway and facilitate H2 desorption, thereby advancing the photocatalytic H2 production process. Furthermore, the strong N-Ag covalent bond masks the weakness of the metal NP cocatalyst in terms of growth and agglomeration under high-temperature reactions, ensuring the remarkable performance of SAAg-g-CN in thermal-assisted photocatalytic H2 production. The average H2 evolution rate (75 μmol g−1 h−1) of this catalyst is nine times that of g-C3N4.

2.3.3. Pollutants Degradation

Organic dyes are the most notorious pollutants in the aquatic environment. The low-cost degradation of organic fuels is of great significance in industrial production. Thermal-coupled photocatalysis degradation of organic pollutants is low-cost and can be performed under mild conditions, so it is considered a promising technology. At present, researchers have made more research results on the degradation of common organic dyes, such as methylene blue (MB) [109,110,111,112], rhodamine B (RhB) [113,114,115,116], and methyl orange (MO) [117,118], etc. Zhou et al. prepared a polydopamine/defective ultrathin mesoporous graphitic carbon nitride (PDA/DCN) Z-scheme organic assembly [109]. Under the light, the removal rate of MB was as high as 98% within 70 min. PDA extends the photoresponse to the near-infrared region and produces a pronounced photothermal effect, thereby enhancing the catalytic performance. Bao et al. fabricated CoFe2O4-modified BiOCl layered microspheres (CFO-BiOCl) [115]. The CFO-BiOCl sample has obvious photothermal catalytic activity, decomposing 99% of RhB within 5 min under full-spectrum light irradiation. Its excellent photocatalytic performance can be attributed to the suitable band arrangement of CoFe2O4 and BiOCl as well as the photothermal effect, which elevates the organic pollutant molecules to a more active state (as shown in Figure 13). Tian et al. integrated porous Au/TiO2 with butterfly P. paris’s wings (PW) as an Au/TiO2_PW photocatalyst [117]. The Au NPs enhanced the photothermal effect of the catalyst, and the PW could adjust the temperature, thus avoiding overheating and reducing the catalytic activity. The material exhibits excellent MO degradation performance due to the synergistic effect between photothermal and photocatalysis. Due to the synergistic effect of photothermal and photocatalysis, the degradation rate of MO was 74.2% within 30 min, which was 3.4 times that of TiO2.
Volatile organic compounds (VOCs) are the main air pollutants on the earth, which have great harm to human health and the environment. At present, the thermal-coupled photocatalysis degradation of VOCs has made great progress. Common VOCs, including toluene [120,121,122,123,124], benzene [136], phenol [132,133], and 2,4-dichlorophenol [125,126,127,128], be efficiently degraded under thermal-coupled photocatalysis. Jiang et al. investigated the performance of Pt−Cu/TiO2 nanocatalysts for photo−thermal catalytic degradation of toluene [120]. The photothermal effect of PtCu alloy nanoparticles and the photocatalysis of TiO2 induced a synergistic effect (as shown in Figure 14), which resulted in a higher efficiency of toluene degradation than the thermocatalytic or photocatalytic process. At 110 °C, the toluene degradation rate is close to 100%. Li et al. prepared Z-scheme Ag/Ag3PO4/CeO2 heterojunction photothermal catalysts for gaseous benzene degradation [136]. At the initial benzene concentration of 600 ppm, the removal rate after 3 h was 90.18%. Ag3PO4 has better solar light absorption ability and CeO2 has a good thermal conductivity at lower temperatures. Under illumination, the photothermal effect promotes the efficient separation of photoexcited electrons and holes. Zhu et al. successfully prepared a photothermal catalyst Bi12CoO20 with wide spectrum adsorption (up to 1000 nm) [133]. The photothermal effect leads to the conversion of Co3+ and Bi3+ to Co2+ and Bi5+ at the octahedral site, which promotes charge separation. The catalyst has a high photothermal activity for the degradation of phenols, which is 2–8 times higher than the reported Bi25CoO40, P25-TiO2, g-C3N4, and Bi2WO6. A layered heterojunction photocatalyst Cu/TiO2-x/CoP was prepared by Xing et al. [125]. The LSPR effect of Cu NPs enables the catalyst to have a significant photothermal effect, thus exhibiting good photothermal-photocatalytic performance. It has a high photocatalytic degradation rate for 2,4-dichlorophenol (99.2%), which is 7.6 times that of the initial TiO2.
In addition to the above organic pollutants, thermal-coupled photocatalysis can also be used for the degradation of bisphenol A (BPA), formaldehyde, etc. Xing et al. fabricated flower-like hollow W18O49@ZnIn2S4 core-shell Z-type heterojunctions on carbon fiber cloth (CC) [130]. The catalyst showed good photocatalytic degradation performance for the degradation of bisphenol A (BPA) under simulated sunlight, and the removal rate of bisphenol A (BPA) was 95%. α-MnO2/graphene oxide nanohybrid was prepared by Rong et al. [137]. The catalyst can decompose 100% of formaldehyde at room temperature. On the one hand, the addition of graphene oxide is beneficial to improving the light absorption capacity and photothermal conversion efficiency; on the other hand, it is beneficial to electron transfer and separation of electrons and holes. These synergistic effects significantly enhance the catalytic activity of α-MnO2/graphene oxide nanohybrid.

3. Mechanical-Coupled Photocatalysis (MCP)

Noncentrosymmetric materials with spontaneous polarization, ferroelectric polarization, or piezoelectric polarization can be used to utilize mechanical energy. For example, when a piezoelectric material is subjected to compressive stress or tensile stress, the material will deform and a built-in electric field with different directions will be generated. This piezoelectric effect has been utilized in the field of photocatalysis to further improve photocatalytic efficiency. Generally, external forces are applied by ultrasound, stirring, and so on. Here, we will try to summarize the current research progress of MCP including the primary source of mechanical energy and materials for MCP. In addition, we also list the application results of mechanical-coupled photocatalysis in recent years in Table 2.

3.1. The Primary Source of Mechanical Energy

3.1.1. Ultrasound

Ultrasound is currently the most important source of mechanical energy in mechanical-coupled photocatalysis [19,20,188,189,190,191]. The photocatalytic reaction is carried out in a custom-made glass vessel exposed to the light and equipped with an ultrasonic generator machine. Ultrasonic waves will produce a cavitation effect in a liquid medium, which can generate microbubbles with a series of dynamic processes, such as contraction, expansion, oscillation, and implosion. After the microbubbles burst, the resulting local hot spot temperature can reach 5000 K, and the pressure can reach 500 atm. Therefore, as for MCP, the ultrasound is always coupled with a cooling system to remove the temperature effect caused by ultrasonic waves [192].
Zeng et al. studied the piezoelectric photocatalytic properties of Au-modified BiOBr, and the reaction setup is shown in Figure 15 [179]. Under the action of ultrasound, the removal rate of carbamazepine (CBZ) reached 95.8% within 30 min, and its rate constant was 1.73 times higher than the sum of the individual piezoelectric and photocatalytic properties. Zhai et al. synthesized BiOX/BaTiO3 (X = Cl, Br, Cl0.166Br0.834) composites for MO degradation [168]. Under the assistance of ultrasound, the piezoelectric photocatalytic performance (k = 0.15824 min−1) of the material far exceeds the sum of photocatalysis alone (k = 0.01159 min−1) and piezoelectric catalysis (k = 0.01208 min−1). This can be attributed to the synergistic effect of its chemical potential difference and piezoelectric potential difference to stop the recombination of photogenerated carriers. Wang et al. prepared oxygen vacancies (OVs)-dominated SrBi4Ti4O15 nanosheets for tetracycline degradation [173]. Under the synergistic effect of visible light and ultrasound, the degradation efficiency was greatly improved, and the oxidation rate constant reached 0.058 min−1, which was 2.15 times and 14.5 times higher than those under only visible light irradiation and only ultrasound, respectively. Xue et al. synthesized a core/shell BaTio3/TiO2 nanocomposite [160]. Under the co-excitation of ultrasound and UV irradiation, the oxidation rate constant of BaTiO3/TiO2 nanofibers in RhB dye can reach 0.0967 min−1, which is 3.22 times higher than that under UV irradiation alone.

3.1.2. Stir

The actuation of most piezoelectric materials requires an ultrasound, which undoubtedly greatly limits their applications. Therefore, researchers began to look for simpler driving methods, such as stirring. A new material design was proposed by Zhang et al. By combining semiconductors with a self-driven energy pad, they achieved piezo-photocatalysis under mild mechanical perturbation (stirring) without external energy input [171]. As shown in Figure 16, the energy pad consists of a porous composite graphene-piezoelectric polymer film that combines piezoelectric and dielectric power generation and stores electricity in situ. This energy buffering results in a large and prolonged electric field in response to mild mechanical perturbations. The experimental results show that the electric field enhances the photocatalytic performance of TiO2, BiOI, and CdS by 300%, 21%, and 400%, respectively. Bian et al. prepared core-shell lead zirconate titanate (PZT)/TiO2 catalysts [178]. By harvesting the mechanical energy of the water, an internal piezoelectric field is induced. Under stirring at 800 rpm, the transient photocurrent of this catalyst was approximately 1.7 times that at 400 rpm. The experimental results show that the photocatalytic degradation rates of RhB, bisphenol A and phenol are greatly improved. This indicates the facilitation of the stirring-induced piezoelectric field. Xue et al. synthesized ZnO/MoS2 nanoarrays on Ni foam [170]. The porous structure of the Ni substrate can effectively induce the applied deformation on the nanorod array and generate a built-in piezoelectric field by stirring the water flow, thereby enhancing the photocatalytic performance. Under illumination, the degradation rate of MO increased with the increase of stirring speed, and its highest degradation efficiency reached 92.7%. Fu et al. synthesized ternary ZnO/ZnS/MoS2 nanocomposites and realized piezoelectric photocatalytic performance [166]. Under illumination, the photocatalytic degradation efficiency of the catalyst for MB was 36.15%, and under the combined action of stirring and illumination, the degradation efficiency was increased to 87.14%. This excellent catalytic performance can be attributed to the stirring-induced piezoelectric field, which enhances charge separation.

3.2. Materials for Mechanical-Coupled Photocatalysis

3.2.1. Titanate-Based Materials

Titanate-based materials are widely used in mechanical-coupled photocatalysis, such as BaTiO3 [147,168,174,180,193,194,195], SrTiO3 [155,182,196], and PbTiO3 [153,183]. Wang et al. synthesized Ag2O−BaTiO3 hybrid photocatalysts [193]. Under periodic ultrasonic excitation, the spontaneous polarization potential of BaTiO3 nanocrystals to ultrasonic waves can be used as an alternating built-in electric field to continuously separate photo-induced carriers, which significantly improves the photocatalytic activity and cycling performance of the Ag2O−BaTiO3 hybrid structure. Li et al. prepared a cyclic Ti32-oxo cluster (CTOC)/BaTiO3/CuS three-layer heterojunction material for tetracycline degradation [174]. As shown in Figure 17, under the combined action of visible light irradiation and ultrasonic vibration, the catalyst reached 100% decomposition efficiency within 60 min. Under ultrasonic vibration, BaTiO3 generates a polarized electric field, which inhibits the recombination of photogenerated carriers and improves photocatalytic performance. Zhang et al. synthesized a BaTiO3/SrTiO3 nanocomposite [155]. The degradation rate of BaTiO3/SrTiO3 nanofibers to RhB within 30 min under the combined action of ultrasound and ultraviolet rays was 97.4%, which was 2.2 times that of pure SrTiO3 nanofibers. Cao et al. synthesized OH-modified SrTiO3 and used it as a catalyst for piezoelectric photocatalytic H2 production [182]. Because the catalyst has super-hydrophilic molecular contacts, as well as more oxygen vacancies, the carrier separation is effectively promoted. Under the action of ultrasound, the optimal piezoelectric-photocatalytic H2 evolution rate (701.2 μmol g−1 h−1) was 5.3 times that of the ordinary SrTiO3 photocatalytic H2 evolution rate. Jiang et al. studied the piezoelectric effect on the heterogeneous photocatalytic H2 production of SrTiO3 (STO) [196]. The piezoelectric effect led by the ultrasonic cavitation had a different influence on heterogeneous photocatalytic H2 production with or without sacrificial agents. The sacrificial agents with different viscosity and positions of the ultrasonic vibrators also affected the photocatalytic H2 production of STO. Through in situ photocatalysis and sonophotocatalysis deposition of Pt nanoparticles, it was found that Pt nanoparticles preferred to aggregate on the local surface of STO nanoparticles by ultrasonic cavitation. It was explained that a piezoelectric potential on the local surface of STO would be built by ultrasonic cavitation. Xu et al. combined PbTiO3 with g-C3N4 to construct a PbTiO3/g-C3N4 heterostructured photocatalyst with close contact [153]. By applying ultrasonic-induced mechanical waves, its built-in piezoelectric field was enhanced, which significantly facilitated charge separation. Under ultrasonic conditions, its photocatalytic performance was enhanced by 30%.

3.2.2. Sulfide-Based Materials

In mechanical-coupled photocatalysis, sulfide materials, such as CdS [176,183,197,198], MoS2 [158,174,175], and ZnS [162,187], have also received much attention. Liu et al. synthesized a novel PbTiO3/CdS hybrid material for photocatalytic H2 production [183]. The piezo-photocatalysis activity of PbTiO3/CdS (849.0 μmol g−1 h−1) was much higher than that of the photocatalysis alone (98.9 μmol g−1 h−1), after the alternating piezoelectric potential was introduced by ultrasonic waves. Specifically, as shown in Figure 18, when the direction of the applied periodic force is changed, the direction of the chemical potential difference is also changed and superimposed on the built-in piezoelectric field, thereby accelerating the carrier separation. Zhao et al. prepared CdS nanoparticles with a phase junction by controlling the hydrothermal temperature [197]. Under ultrasonic vibration, a dynamic piezoelectric field is excited inside the CdS nanoparticles. The presence of the piezoelectric field eliminates the shielding effect of carriers in the built-in electric field of the phase junction, enabling the continuous separation of photogenerated carriers, and thereby enhancing the photocatalytic efficiency of CdS. Cui et al. synthesized a MoS2@Ag2O heterostructured photocatalyst for efficient all-solar photocatalysis [199]. Under full sunlight and ultrasonic excitation, its MO degradation rate exceeds 95%. During this process, the spontaneous polarization potential of MoS2 forms a built-in electric field, and the ultrasonic wave as the driving force can continuously change the potential generated by the piezoelectric effect. Therefore, under ultrasonic excitation, the piezoelectric effect enhances the carrier separation, thereby enhancing the photocatalytic performance. Lin et al. developed a monolayer and few-layer MoS2 nanosheets (NSs) for bacterial inactivation [200]. After 60 min of mechanical vibration or visible light irradiation, MoS2 NSs can reduce Escherichia coli (E. coli) by 99.999%. After the deposition of Au nanoparticles (NPs) on MoS2 NSs, the time was shortened to 45 min. Furthermore, after synergizing mechanical vibration and near-infrared (NIR) light irradiation, Au-MoS2 eliminated E. coli in only 15 min. Zhai et al. modified In2S3 nanosheets with ZnS and BaTiO3(ZnS/In2S3/BTO) [187]. Under the combined action of visible light and ultrasound, H2O2 of approximately 378 μmol g−1 h−1 was generated in 100 min. The enhanced H2O2 yield on ZnS/In2S3/BTO under piezophotocatalysis can be attributed to the piezoelectric field promoted the separation of photogenerated carriers.

3.2.3. Polyvinylidene Fluoride

As a common piezoelectric material, polyvinylidene fluoride (PVDF) has great application potential in mechanical-coupled photocatalysis [161,169,171,181,185,201,202]. Sun et al. successfully synthesized a ZnS/Bi2S3-PVDF thin film photocatalyst containing S vacancies (ZnS-VS/Bi2S3-PVDF) [181]. As shown in Figure 19, the synergistic effect of the PVDF polarization electric field induced by ultrasonic cavitation and the built-in electric field formed between the ZnS-VS/Bi2S3 dissimilar junction structure is beneficial to suppressing the recombination of carriers. Its H2 evolution rate reaches 10.07 mmol g−1 h−1. Zhang et al. prepared a PVDF/MoS2 cavity/Au heterostructured photocatalyst for MB degradation [202]. The synergistic effect of the MoS2 cavity and PVDF fibers not only enhances the intrinsic piezoelectric activity of PVDF but also increases the catalytically active centers and free carrier density. The degradation efficiency was as high as 99.9% within 45 min. Lu et al. combined ZnO with PVDF thin films to obtain a bi-piezoelectric integration effect, resulting in enhanced photocatalytic performance [161]. In addition, the piezoelectric potential and photocatalytic activity of the as-prepared ZnO nanorods increased with the increase of the (100) crystal face ratio. Under flowing water conditions, its optimal photocatalytic activity is 10 times that of a single photocatalytic activity. Wang et al. synthesized a bi-piezoelectric ZnO nanorod (NR)/PVDF-HFP sponge-like film [169]. Under water flow, the catalyst can generate an integrated piezoelectric field that greatly drives the separation of photogenerated charge carriers. The results showed that the reaction rate constant for the degradation of MO increased three-fold when the flow rate was increased from 200 rpm to 1000 rpm.

4. Electromagnetism-Coupled Photocatalysis (ECP)

The electromagnetism-coupled photocatalysis is divided into two parts to review. They are electro-coupled photocatalysis and magnetism-coupled photocatalysis, respectively. In addition, Table 3 summarizes the application results of electromagnetism-coupled photocatalysis in recent years.

4.1. Electro-Coupled Photocatalysis

Adding an external electric field to the photocatalytic system can effectively promote the separation and migration of photogenerated carriers. At present, there are two main ways for researchers to add an electric field: one is to directly connect to an external power source during the photocatalytic process, and the other is to polarize the catalyst in the electric field and then perform photocatalytic experiments.
Directly adding an external electric field during the photocatalytic process is a common utilization method in electro-coupled photocatalysis. Li et al. synthesized rutile TiO2 nanograss on titanium mesh [222]. By connecting the titanium mesh to a power source, the photocatalytic removal capacity of rutile TiO2 nanograss for Cr(VI) ions in wastewater can reach 143.8 mg/g under an external electric field and sunlight. Studies have shown that electrons provided by an external electric field can effectively inhibit the recombination of photogenerated carriers, thereby enhancing the catalytic activity. Pan et al. synthesized layered TiO2 microspheres for the removal of tributyltin (TBT) in electro-coupled photocatalysis, and the reaction setup is shown in Figure 20 [214]. The bias potential provided by the applied electric field greatly facilitates the separation of photogenerated carriers. The results show that the removal reaction rate constant (k = 0.0488 min−1) under electro-coupled photocatalysis is nearly times higher than that under photocatalysis (k = 0.0052 min−1). Zhang et al. developed a new electro-assisted photocatalytic technology for liquid and gas phase reactions (such as water splitting and CO2 reduction) [220]. They applied a lower external voltage on the self-doped TiO2 nanotube films. With electrical assistance, the recombination of photogenerated carriers is greatly suppressed, and photoexcited electrons are ejected from the catalyst, enabling spatial separation of carriers. It is worth mentioning that the photocatalytic redox reaction is integrated on a thin film in this system, rather than taking place on traditional separate electrodes.
In addition to directly adding an external electric field, the researchers also developed another way of electrically coupled photocatalysis. They put the ferroelectric material into an external electric field for polarization before conducting the photocatalytic experiments. The strong ferroelectric field formed by the polarized material can effectively promote charge separation and improve photocatalytic efficiency. Nam et al. prepared polarized K0.5Na0.5NbO3 powder by a corona polarization method for H2 production [221]. The enlarged positively charged surface of the catalyst promotes the photocatalytic activity. Under optimal conditions, its photocatalytic activity was increased by 7.4 times. Yun et al. prepared well-polarized BiFeO3 nanoparticles by a simple electrical polarization method [210]. The photocatalytic degradation rate of RhB by the polarized BiFeO3 nanoparticles was increased by two times compared with the unpolarized. Its excellent photocatalytic performance can be attributed to the improved carrier separation by ferroelectric polarization. Tian et al. prepared ultrathin Bi2MoO6 nanosheets with strong ferroelectricity by a corona polarization method for photocatalytic CO2 reduction [223]. Its CO generation rate is 14.38 μmol g−1 h−1, which is more than 10 times higher than that of bulk Bi2MoO6. As shown in Figure 21, this combined strategy significantly promotes the separation of photogenerated electrons and holes, enriches the reaction sites for CO2 adsorption, and jointly improves the photocatalytic CO2 reduction performance.

4.2. Magnetism-Coupled Photocatalysis

In recent years, more and more attention has been paid to the promotion of magnetic fields on photocatalytic performance. In a magnetism-coupled photocatalytic system, the magnetic-field-induced Lorentz force diverts the charge from its original recombination path, leading to more carriers participating in the reaction, thereby enhancing the photocatalytic activity [206,213,216,225,226]. In addition, for materials with electron spin properties, the magnetic field can also enhance the electron-spin parallel alignment, which also has a positive effect on photocatalytic performance [209,217,227].
Gao et al. reported a Lorentz force-enhanced photogenerated carrier separation strategy using highly crystalline TiO2 nanosheets with few defects as photocatalysts [225]. Studies have shown that the Lorentz force suppresses photogenerated carrier recombination, making more carriers available for transport in photocatalysts. The results showed that the photocatalytic degradation rate of MO was increased by 26% only by placing permanent magnets under the photocatalytic reactor. Nan et al. synthesized a 3D/2D Mn2O3/g-C3N4 photocatalyst for simultaneous removal of nitrate and ammonia from polluted water and developed a magnetic field-enhanced photocatalytic system [226]. Under the action of an external magnetic field, the removal efficiencies of nitrate and ammonia were 94.5% and 97.4%, respectively. The external magnetic field promotes the separation of carriers by generating the Lorentz force. Tsang et al. reported a photocatalytic water splitting system based on Fe3O4/N-TiO2 magnetic photocatalyst assisted by the local magnetic field effect [217]. At a magnetic field strength of 180 mT and 270 °C, the quantum efficiency at 437 nm reaches 88.7%. As shown in Figure 22, the synergistic effect of the Lorentz force and spin polarization greatly facilitates the charge separation process, thereby enhancing the photocatalytic efficiency. Sang et al. prepared a CdS/MoS2/Mo hybrid structured photocatalyst and studied the effect of magnetic field on photocatalytic activity [218]. The relative motion between the metal molybdenum sheet and the rotating magnetic field forms a kinematic electromotive force. This electromotive force significantly inhibits the photo-induced carrier recombination of CdS. Under the synergistic effect of MoS2 co-catalyst, its photocatalytic H2 production performance is improved by approximately 89% compared with single photocatalysis. Chen et al. synthesized a Mn3O4/γ-MnOOH photocatalyst for norfloxacin (NOR) degradation [215]. Under the action of magnetic field-assisted visible light, the degradation rate of NOR was 98.8% within 60 min. In neutral media, positively charged NOR and negatively charged catalysts align in the presence of a magnetic field, thereby enhancing the reactivity. Furthermore, the opposing Lorentz force contributed to the mutual attraction between NOR and catalyst, which accelerated NOR degradation.

5. Conclusions

In this review, we summarized the current research progress of thermal-coupled Photocatalysis (TCP), Mechanical-coupled Photocatalysis (MCP), and Electromagnetism-coupled Photocatalysis (ECP), including the primary sources of external energy, the related materials, and the main applications. These external physical fields can significantly improve the efficiency of the photocatalytic system, which have different mechanisms and respective advantages. In summary, the common feature of these external physical fields coupling to photocatalysis is that external physical fields can re-emit the trapped electrons and drive them rapidly to transfer to the surface of the photocatalyst. Though assisted external physical fields can overcome the inherent weakness of the photocatalytic reaction process, there are still some fundamental issues and challenges to be explored.
(i)
Evaluation of energy conversion efficiency. When external physical fields are applied to the photocatalytic reaction process, additional energy will inevitably be input into the system. Therefore, the solar energy conversion efficiency cannot only be considered when calculating energy conversion efficiency. The extra energy generated by external physical fields should be taken into account.
(ii)
Ubiquitous thermal effect. Thermal energy is low-quality energy. The external physical fields will eventually dissipate into heat energy. The special thermal effects can be generated by microwaves, ultrasonic waves, and electromagnetism waves. Thus, it is important to distinguish thermal effects and non-thermal effects on photocatalysis coupling by external physical fields.
(iii)
A single material response for multiple-physical fields. To realize photocatalysis coupling by external physical fields, composite materials combining photocatalysts and external field absorption materials are usually used. However, the composite materials are complicated both in the synthesis and photocatalytic reaction process. So, it is interesting to explore a single material that can respond to multiple-physical fields. For example, the spontaneous symmetry-breaking semiconductors can absorb multiple-physical fields at the same time, which might be used in photocatalysis coupling by external physical fields.
(iv)
Mechanisms for potential barrier formation by external physical fields. When external physical fields are applied to photocatalysts, a built-in electric field with different directions will be generated to separate the photogenerated carriers. However, mechanisms for potential barrier formation by external physical fields need to be uncovered. Thus, this built-in electric field can be controlled and optimized in favor of photocatalysis coupling by external physical fields.
(v)
Reactor design for photocatalysis coupling by external physical fields. The traditional photocatalytic reactor is not satisfied with photocatalysis coupling by external physical fields. The design principles of the reactor must be high-efficiency and well-adapted for different external physical fields.
(vi)
Horizontal comparison is neglected in the study of photocatalysis coupling by external physical fields. Different external physical fields should make different contributions to photocatalysis. Keeping a balance between photocatalytic efficiency and economic efficiency, the best assisted physical field for photocatalysis needs to be further studied.

Author Contributions

Conceptualization, W.F. and W.S.; writing—original draft preparation, Y.M., Y.Y. and Y.J.; writing—review and editing, W.F. and Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Basic Research and Development Program (2018YFB1502001), the National Natural Science Foundation of China (21773153), and the Postgraduate Research & Practice Innovation Program of Jiangsu Province (Yangzhou University, SJCX21_1576).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gong, E.; Ali, S.; Hiragond, C.B.; Kim, H.S.; Powar, N.S.; Kim, D.; Kim, H.; In, S.-I. Solar fuels: Research and development strategies to accelerate photocatalytic CO2 conversion into hydrocarbon fuels. Energ Environ. Sci. 2022, 15, 880–937. [Google Scholar] [CrossRef]
  2. Din, M.I.; Khalid, R.; Najeeb, J.; Hussain, Z. Fundamentals and photocatalysis of methylene blue dye using various nanocatalytic assemblies- a critical review. J. Clean Prod. 2021, 298, 126567. [Google Scholar] [CrossRef]
  3. Hisatomi, T.; Domen, K. Reaction systems for solar hydrogen production via water splitting with particulate semiconductor photocatalysts. Nat. Catal. 2019, 2, 387–399. [Google Scholar] [CrossRef]
  4. Wei, Z.; Liu, J.; Shangguan, W. A review on photocatalysis in antibiotic wastewater: Pollutant degradation and hydrogen production. Chin. J. Catal. 2020, 41, 1440–1450. [Google Scholar] [CrossRef]
  5. Shafiq, I.; Shafique, S.; Akhter, P.; Abbas, G.; Qurashi, A.; Hussain, M. Efficient catalyst development for deep aerobic photocatalytic oxidative desulfurization: Recent advances, confines, and outlooks. Catal. Rev. 2021, 1–46. [Google Scholar] [CrossRef]
  6. Shafiq, I.; Hussain, M.; Rashid, R.; Shafique, S.; Akhter, P.; Yang, W.; Ahmed, A.; Nawaz, Z.; Park, Y.-K. Development of hierarchically porous LaVO4 for efficient visible-light-driven photocatalytic desulfurization of diesel. Chem. Eng. J. 2021, 420, 130529. [Google Scholar] [CrossRef]
  7. Mahboob, I.; Shafiq, I.; Shafique, S.; Akhter, P.; Amjad, U.-e.-S.; Hussain, M.; Park, Y.-K. Effect of active species scavengers in photocatalytic desulfurization of hydrocracker diesel using mesoporous Ag3VO4. Chem. Eng. J. 2022, 441, 136063. [Google Scholar] [CrossRef]
  8. Lei, W.; Suzuki, N.; Terashima, C.; Fujishima, A. Hydrogel photocatalysts for efficient energy conversion and environmental treatment. Front. Energy 2021, 15, 577–595. [Google Scholar] [CrossRef]
  9. Bedia, J.; Muelas-Ramos, V.; Penas-Garzon, M.; Gomez-Aviles, A.; Rodriguez, J.J.; Belver, C. A review on the synthesis and characterization of metal organic frameworks for photocatalytic water purification. Catalysts 2019, 9, 52. [Google Scholar] [CrossRef]
  10. Chen, Y.L.; Bai, X. A review on quantum dots modified g-C3N4-based photocatalysts with improved photocatalytic activity. Catalysts 2020, 10, 142. [Google Scholar] [CrossRef] [Green Version]
  11. Kumaravel, V.; Imam, M.D.; Badreldin, A.; Chava, R.K.; Do, J.Y.; Kang, M.; Abdel-Wahab, A. Photocatalytic hydrogen production: Role of sacrificial reagents on the activity of oxide, carbon, and sulfide catalysts. Catalysts 2019, 9, 276. [Google Scholar] [CrossRef]
  12. Li, X.; Chen, Y.; Tao, Y.; Shen, L.; Xu, Z.; Bian, Z.; Li, H. Challenges of photocatalysis and their coping strategies. Chem. Catal. 2022, 2, 1315–1345. [Google Scholar] [CrossRef]
  13. Yuan, J.; Liu, H.; Wang, S.; Li, X. How to apply metal halide perovskites to photocatalysis: Challenges and development. Nanoscale 2021, 13, 10281–10304. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, H.; Li, X.; Zhao, X.; Li, C.; Song, X.; Zhang, P.; Huo, P.; Li, X. A review on heterogeneous photocatalysis for environmental remediation: From semiconductors to modification strategies. Chin. J. Catal. 2022, 43, 178–214. [Google Scholar] [CrossRef]
  15. Chen, S.; Yin, H.; Liu, P.; Wang, Y.; Zhao, H. Stabilisation and Performance Enhancement Strategies for Halide Perovskite Photocatalysts. Adv. Mater. 2022, e2203836. [Google Scholar] [CrossRef] [PubMed]
  16. Mai, H.; Chen, D.; Tachibana, Y.; Suzuki, H.; Abe, R.; Caruso, R.A. Developing sustainable, high-performance perovskites in photocatalysis: Design strategies and applications. Chem. Soc. Rev. 2021, 50, 13692–13729. [Google Scholar] [CrossRef]
  17. Tasleem, S.; Tahir, M. Current trends in strategies to improve photocatalytic performance of perovskites materials for solar to hydrogen production. Renew. Sustain. Energ Rev. 2020, 132, 110073. [Google Scholar] [CrossRef]
  18. Xu, Z.; Zhu, Q.; Xi, X.; Xing, M.; Zhang, J. Z-scheme CdS/WO3 on a carbon cloth enabling effective hydrogen evolution. Front. Energy 2021, 15, 678–686. [Google Scholar] [CrossRef]
  19. Li, X.; Wang, W.; Dong, F.; Zhang, Z.; Han, L.; Luo, X.; Huang, J.; Feng, Z.; Chen, Z.; Jia, G.; et al. Recent Advances in Noncontact External-Field-Assisted Photocatalysis: From Fundamentals to Applications. ACS Catal. 2021, 11, 4739–4769. [Google Scholar] [CrossRef]
  20. Hu, C.; Tu, S.; Tian, N.; Ma, T.; Zhang, Y.; Huang, H. Photocatalysis Enhanced by External Fields. Angew. Chem. Int. Ed. 2021, 60, 16309–16328. [Google Scholar] [CrossRef]
  21. Sun, M.; Zhao, B.; Chen, F.; Liu, C.; Lu, S.; Yu, Y.; Zhang, B. Thermally-assisted photocatalytic CO2 reduction to fuels. Chem. Eng. J. 2021, 408, 127280. [Google Scholar] [CrossRef]
  22. Song, C.; Wang, Z.; Yin, Z.; Xiao, D.; Ma, D. Principles and applications of photothermal catalysis. Chem. Catal. 2022, 2, 52–83. [Google Scholar] [CrossRef]
  23. Lu, Y.; Zhang, H.; Fan, D.; Chen, Z.; Yang, X. Coupling solar-driven photothermal effect into photocatalysis for sustainable water treatment. J. Hazard. Mater. 2022, 423, 127128. [Google Scholar] [CrossRef] [PubMed]
  24. Ma, R.; Sun, J.; Li, D.H.; Wei, J.J. Review of synergistic photo-thermo-catalysis: Mechanisms, materials and applications. Int. J. Hydrogen Energ. 2020, 45, 30288–30324. [Google Scholar] [CrossRef]
  25. Wang, Z.J.; Song, H.; Liu, H.; Ye, J. Coupling of Solar Energy and Thermal Energy for Carbon Dioxide Reduction: Status and Prospects. Angew. Chem. Int. Ed. 2020, 59, 8016–8035. [Google Scholar] [CrossRef] [PubMed]
  26. Du, S.; Bian, X.; Zhao, Y.; Shi, R.; Zhang, T. Progress and Prospect of Photothermal Catalysis. Chem. Res. Chin. Univ. 2022, 38, 723–734. [Google Scholar] [CrossRef]
  27. Keller, N.; Ivanez, J.; Highfield, J.; Ruppert, A.M. Photo-/thermal synergies in heterogeneous catalysis: Towards low-temperature (solar-driven) processing for sustainable energy and chemicals. Appl. Catal. B Environ. 2021, 296, 120320. [Google Scholar] [CrossRef]
  28. Cheng, C.; Wang, J.; Guo, X.; Xing, F.; Huang, C.; Song, M. Thermal-assisted photocatalytic H2 production over sulfur vacancy-rich Co0.85Se/Mn0.3Cd0.7S nanorods under visible light. Appl. Surf. Sci. 2021, 557, 149812. [Google Scholar] [CrossRef]
  29. Hu, S.; Shi, J.; Luo, B.; Ai, C.; Jing, D. Significantly enhanced photothermal catalytic hydrogen evolution over Cu2O-rGO/TiO2 composite with full spectrum solar light. J. Colloid Interface Sci. 2022, 608, 2058–2065. [Google Scholar] [CrossRef]
  30. Li, P.; Liu, L.; An, W.; Wang, H.; Cui, W. Efficient photothermal catalytic CO2 reduction to CH3CH2OH over Cu2O/g-C3N4 assisted by ionic liquids. Appl. Surf. Sci. 2021, 565, 150448. [Google Scholar] [CrossRef]
  31. Du, C.; Yan, B.; Yang, G. Promoting photocatalytic hydrogen evolution by introducing hot islands: SnSe nanoparticles on ZnIn2S4 monolayer. Chem. Eng. J. 2021, 404, 126477. [Google Scholar] [CrossRef]
  32. Lu, Y.; Li, Y.; Wang, Y.; Zhang, J. Two-photon induced NIR active core-shell structured WO3/CdS for enhanced solar light photocatalytic performance. Appl. Catal. B Environ. 2020, 272, 118979. [Google Scholar] [CrossRef]
  33. Tang, Y.; Zhou, W.; Shang, Q.; Guo, Y.; Hu, H.; Li, Z.; Zhang, Y.; Liu, L.; Wang, H.; Tan, X.; et al. Discerning the mechanism of expedited interfacial electron transformation boosting photocatalytic hydrogen evolution by metallic 1T-WS2-induced photothermal effect. Appl. Catal. B Environ. 2022, 310, 121295. [Google Scholar] [CrossRef]
  34. Zhang, X.; Liu, M.; Kang, Z.; Wang, B.; Wang, B.; Jiang, F.; Wang, X.; Yang, D.-P.; Luque, R. NIR-triggered photocatalytic/photothermal/photodynamic water remediation using eggshell-derived CaCO3/CuS nanocomposites. Chem. Eng. J. 2020, 388, 124304. [Google Scholar] [CrossRef]
  35. Ai, Z.; Yang, P.; Lu, X. Degradation of 4-chlorophenol by a microwave assisted photocatalysis method. J. Hazard. Mater. 2005, 124, 147–152. [Google Scholar] [CrossRef] [PubMed]
  36. Nüchter, M.; Müller, U.; Ondruschka, B.; Tied, A.; Lautenschläger, W. Microwave-Assisted Chemical Reactions. Chem. Eng. Technol. 2003, 26, 1207–1216. [Google Scholar] [CrossRef]
  37. Ling, L.; Feng, Y.; Li, H.; Chen, Y.; Wen, J.; Zhu, J.; Bian, Z. Microwave induced surface enhanced pollutant adsorption and photocatalytic degradation on Ag/TiO2. Appl. Surf. Sci. 2019, 483, 772–778. [Google Scholar] [CrossRef]
  38. Wang, X.; Jiang, L.; Li, K.; Wang, J.; Fang, D.; Zhang, Y.; Tian, D.; Zhang, Z.; Dionysiou, D.D. Fabrication of novel Z-scheme SrTiO3/MnFe2O4 system with double-response activity for simultaneous microwave-induced and photocatalytic degradation of tetracycline and mechanism insight. Chem. Eng. J. 2020, 400, 125981. [Google Scholar] [CrossRef]
  39. Gayathri, P.V.; Yesodharan, S.; Yesodharan, E.P. Microwave/Persulphate assisted ZnO mediated photocatalysis (MW/PS/UV/ZnO) as an efficient advanced oxidation process for the removal of RhB dye pollutant from water. J. Environ. Chem. Eng. 2019, 7, 103122. [Google Scholar] [CrossRef]
  40. Mateo, D.; Cerrillo, J.L.; Durini, S.; Gascon, J. Fundamentals and applications of photo-thermal catalysis. Chem. Soc. Rev. 2021, 50, 2173–2210. [Google Scholar] [CrossRef]
  41. Linic, S.; Christopher, P.; Ingram, D.B. Plasmonic-metal nanostructures for efficient conversion of solar to chemical energy. Nat. Mater. 2011, 10, 911–921. [Google Scholar] [CrossRef]
  42. Manthiram, K.; Alivisatos, A.P. Tunable localized surface plasmon resonances in tungsten oxide nanocrystals. J. Am. Chem. Soc. 2012, 134, 3995–3998. [Google Scholar] [CrossRef] [PubMed]
  43. Linic, S.; Aslam, U.; Boerigter, C.; Morabito, M. Photochemical transformations on plasmonic metal nanoparticles. Nat. Mater. 2015, 14, 567–576. [Google Scholar] [CrossRef] [PubMed]
  44. Chan, G.H.; Zhao, J.; Hicks, E.M.; Schatz, G.C.; Van Duyne, R.P. Plasmonic Properties of Copper Nanoparticles Fabricated by Nanosphere Lithography. Nano Lett. 2007, 7, 1947–1952. [Google Scholar] [CrossRef]
  45. Wang, F.; Li, C.; Chen, H.; Jiang, R.; Sun, L.D.; Li, Q.; Wang, J.; Yu, J.C.; Yan, C.H. Plasmonic harvesting of light energy for Suzuki coupling reactions. J. Am. Chem. Soc. 2013, 135, 5588–5601. [Google Scholar] [CrossRef]
  46. Yu, G.; Qian, J.; Zhang, P.; Zhang, B.; Zhang, W.; Yan, W.; Liu, G. Collective excitation of plasmon-coupled Au-nanochain boosts photocatalytic hydrogen evolution of semiconductor. Nat. Commun. 2019, 10, 4912. [Google Scholar] [CrossRef]
  47. Jian, C.-c.; Zhang, J.; He, W.; Ma, X. Au-Al intermetallic compounds: A series of more efficient LSPR materials for hot carriers-based applications than noble metal Au. Nano Energy 2021, 82, 105763. [Google Scholar] [CrossRef]
  48. Han, B.; Chen, L.; Jin, S.; Guo, S.; Park, J.; Yoo, H.S.; Park, J.H.; Zhao, B.; Jung, Y.M. Modulating Mechanism of the LSPR and SERS in Ag/ITO Film: Carrier Density Effect. J. Phys. Chem. Lett. 2021, 12, 7612–7618. [Google Scholar] [CrossRef]
  49. Li, P.; Liu, L.; An, W.; Wang, H.; Guo, H.; Liang, Y.; Cui, W. Ultrathin porous g-C3N4 nanosheets modified with AuCu alloy nanoparticles and C-C coupling photothermal catalytic reduction of CO to ethanol. Appl. Catal. B Environ. 2020, 266, 118618. [Google Scholar] [CrossRef]
  50. Yang, Y.; Cong, Y.; Lin, X.; Cao, B.; Dong, D.; Liu, K.; Xiao, Y.; Shang, J.; Bao, Y.; Liu, Y.; et al. Dual LSPR of Au/W18O49 heterostructures for upconversion enhancement and application of molecular detection. J. Mater. Chem. A 2020, 8, 4040–4048. [Google Scholar] [CrossRef]
  51. Kong, W.; Xing, Z.; Fang, B.; Cui, Y.; Li, Z.; Zhou, W. Plasmon Ag/Na-doped defective graphite carbon nitride/NiFe layered double hydroxides Z-scheme heterojunctions toward optimized photothermal-photocatalytic-Fenton performance. Appl. Catal. B Environ. 2022, 304, 120969. [Google Scholar] [CrossRef]
  52. Liao, J.; Xu, Y.; Zhao, Y.; Wang, C.-C.; Ge, C. Ag and Fe3O4 Comodified WO3-x Nanocomposites for Catalytic Photothermal Degradation of Pharmaceuticals and Personal Care Products. ACS Appl. Nano Mater. 2021, 4, 1898–1905. [Google Scholar] [CrossRef]
  53. Wei, T.; Liu, Y.; Dong, W.; Zhang, Y.; Huang, C.; Sun, Y.; Chen, X.; Dai, N. Surface-dependent localized surface plasmon resonances in CuS nanodisks. ACS Appl. Mater. Interfaces 2013, 5, 10473–10477. [Google Scholar] [CrossRef]
  54. Liu, Y.; Liu, M.; Swihart, M.T. Reversible Crystal Phase Interconversion between Covellite CuS and High Chalcocite Cu2S Nanocrystals. Chem. Mater. 2017, 29, 4783–4791. [Google Scholar] [CrossRef]
  55. Gao, H.; Chen, Y.; Li, H.; Zhang, F.; Tian, G. Hierarchical Cu7S4-Cu9S8 heterostructure hollow cubes for photothermal aerobic oxidation of amines. Chem. Eng. J. 2019, 363, 247–258. [Google Scholar] [CrossRef]
  56. Lin, H.; Wang, X.; Yu, L.; Chen, Y.; Shi, J. Two-Dimensional Ultrathin MXene Ceramic Nanosheets for Photothermal Conversion. Nano Lett. 2017, 17, 384–391. [Google Scholar] [CrossRef]
  57. Naik, G.V.; Schroeder, J.L.; Ni, X.; Kildishev, A.V.; Sands, T.D.; Boltasseva, A. Titanium nitride as a plasmonic material for visible and near-infrared wavelengths [erratum]. Opt. Mater. Express 2013, 3, 1658. [Google Scholar] [CrossRef]
  58. Lou, Z.; Gu, Q.; Liao, Y.; Yu, S.; Xue, C. Promoting Pd-catalyzed Suzuki coupling reactions through near-infrared plasmon excitation of WO3-x nanowires. Appl. Catal. B Environ. 2016, 184, 258–263. [Google Scholar] [CrossRef]
  59. Wang, Y.; Liu, M.; Wu, C.; Gao, J.; Li, M.; Xing, Z.; Li, Z.; Zhou, W. Hollow Nanoboxes Cu2-xS@ZnIn2S4 Core-Shell S-Scheme Heterojunction with Broad-Spectrum Response and Enhanced Photothermal-Photocatalytic Performance. Small 2022, 18, e2202544. [Google Scholar] [CrossRef]
  60. Xie, Z.; Duo, Y.; Lin, Z.; Fan, T.; Xing, C.; Yu, L.; Wang, R.; Qiu, M.; Zhang, Y.; Zhao, Y.; et al. The Rise of 2D Photothermal Materials beyond Graphene for Clean Water Production. Adv. Sci. 2020, 7, 1902236. [Google Scholar] [CrossRef] [Green Version]
  61. Wang, J.; Li, Y.; Deng, L.; Wei, N.; Weng, Y.; Dong, S.; Qi, D.; Qiu, J.; Chen, X.; Wu, T. High-Performance Photothermal Conversion of Narrow-Bandgap Ti2O3 Nanoparticles. Adv. Mater. 2017, 29, 1603730. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, Z.; Yang, C.; Lin, T.; Yin, H.; Chen, P.; Wan, D.; Xu, F.; Huang, F.; Lin, J.; Xie, X.; et al. Visible-light photocatalytic, solar thermal and photoelectrochemical properties of aluminium-reduced black titania. Energy Environ. Sci. 2013, 6, 3007–3014. [Google Scholar] [CrossRef]
  63. Ye, M.; Jia, J.; Wu, Z.; Qian, C.; Chen, R.; O’Brien, P.G.; Sun, W.; Dong, Y.; Ozin, G.A. Synthesis of Black TiOx Nanoparticles by Mg Reduction of TiO2 Nanocrystals and their Application for Solar Water Evaporation. Adv. Energy Mater. 2017, 7, 1601811. [Google Scholar] [CrossRef]
  64. Low, J.; Zhang, L.; Zhu, B.; Liu, Z.; Yu, J. TiO2 Photonic Crystals with Localized Surface Photothermal Effect and Enhanced Photocatalytic CO2 Reduction Activity. ACS Sustain. Chem. Eng. 2018, 6, 15653–15661. [Google Scholar] [CrossRef]
  65. Pan, J.; Yu, X.; Dong, J.; Zhao, L.; Liu, L.; Liu, J.; Zhao, X.; Liu, L. Diatom-Inspired TiO2-PANi-Decorated Bilayer Photothermal Foam for Solar-Driven Clean Water Generation. ACS Appl. Mater. Interfaces 2021, 13, 58124–58133. [Google Scholar] [CrossRef]
  66. Chou, S.S.; Kaehr, B.; Kim, J.; Foley, B.M.; De, M.; Hopkins, P.E.; Huang, J.; Brinker, C.J.; Dravid, V.P. Chemically exfoliated MoS2 as near-infrared photothermal agents. Angew. Chem. Int. Ed. 2013, 52, 4160–4164. [Google Scholar] [CrossRef] [PubMed]
  67. Ghim, D.; Jiang, Q.; Cao, S.; Singamaneni, S.; Jun, Y.-S. Mechanically interlocked 1T/2H phases of MoS2 nanosheets for solar thermal water purification. Nano Energy 2018, 53, 949–957. [Google Scholar] [CrossRef]
  68. Ding, D.; Huang, W.; Song, C.; Yan, M.; Guo, C.; Liu, S. Non-stoichiometric MoO3-x quantum dots as a light-harvesting material for interfacial water evaporation. Chem. Commun. 2017, 53, 6744–6747. [Google Scholar] [CrossRef]
  69. Qi, Y.; Song, L.; Ouyang, S.; Liang, X.; Ning, S.; Zhang, Q.; Ye, J. Photoinduced Defect Engineering: Enhanced Photothermal Catalytic Performance of 2D Black In2O3-x Nanosheets with Bifunctional Oxygen Vacancies. Adv. Mater. 2020, 32, e1903915. [Google Scholar] [CrossRef]
  70. Zhang, Y.; Heo, Y.-J.; Son, Y.-R.; In, I.; An, K.-H.; Kim, B.-J.; Park, S.-J. Recent advanced thermal interfacial materials: A review of conducting mechanisms and parameters of carbon materials. Carbon 2019, 142, 445–460. [Google Scholar] [CrossRef]
  71. Li, J.; Ma, L.; Fu, C.; Huang, Y.; Luo, B.; Cao, J.; Geng, J.; Jing, D. Urchinlike Carbon-Coated TiO2 Microspheres with Enhanced Photothermal–Photocatalytic Hydrogen Evolution Performance for Full-Spectrum Solar Energy Conversion. Ind. Eng. Chem. Res. 2022, 61, 6436–6447. [Google Scholar] [CrossRef]
  72. Zhang, Q.; Xu, W.; Wang, X. Carbon nanocomposites with high photothermal conversion efficiency. Sci. China Mater. 2018, 61, 905–914. [Google Scholar] [CrossRef]
  73. Tong, X.; Li, N.; Zeng, M.; Wang, Q. Organic phase change materials confined in carbon-based materials for thermal properties enhancement: Recent advancement and challenges. Renew. Sustain. Energy Rev. 2019, 108, 398–422. [Google Scholar] [CrossRef]
  74. Guo, X.; Cheng, S.; Cai, W.; Zhang, Y.; Zhang, X.-a. A review of carbon-based thermal interface materials: Mechanism, thermal measurements and thermal properties. Mater. Design 2021, 209, 109936. [Google Scholar] [CrossRef]
  75. Chen, C.; Li, Y.; Song, J.; Yang, Z.; Kuang, Y.; Hitz, E.; Jia, C.; Gong, A.; Jiang, F.; Zhu, J.Y.; et al. Highly Flexible and Efficient Solar Steam Generation Device. Adv. Mater. 2017, 29, 1701756. [Google Scholar] [CrossRef]
  76. Yu, Y.; Chen, S.; Jia, Y.; Qi, T.; Xiao, L.; Cui, X.; Zhuang, D.; Wei, J. Ultra-black and self-cleaning all carbon nanotube hybrid films for efficient water desalination and purification. Carbon 2020, 169, 134–141. [Google Scholar] [CrossRef]
  77. Zhang, F.; Xu, D.; Zhang, D.; Ma, L.; Wang, J.; Huang, Y.; Chen, M.; Qian, H.; Li, X. A durable and photothermal superhydrophobic coating with entwinned CNTs-SiO2 hybrids for anti-icing applications. Chem. Eng. J. 2021, 423, 130238. [Google Scholar] [CrossRef]
  78. Robinson, J.T.; Tabakman, S.M.; Liang, Y.; Wang, H.; Casalongue, H.S.; Vinh, D.; Dai, H. Ultrasmall reduced graphene oxide with high near-infrared absorbance for photothermal therapy. J. Am. Chem. Soc. 2011, 133, 6825–6831. [Google Scholar] [CrossRef]
  79. Wang, Y.; Wang, C.; Song, X.; Megarajan, S.K.; Jiang, H. A facile nanocomposite strategy to fabricate a rGO–MWCNT photothermal layer for efficient water evaporation. J. Mater. Chem. A 2018, 6, 963–971. [Google Scholar] [CrossRef]
  80. Zhang, C.-R.; Cui, W.-R.; Niu, C.-P.; Yi, S.-M.; Liang, R.-P.; Qi, J.-X.; Chen, X.-J.; Jiang, W.; Zhang, L.; Qiu, J.-D. rGO-based covalent organic framework hydrogel for synergistically enhance uranium capture capacity through photothermal desalination. Chem. Eng. J. 2022, 428, 131178. [Google Scholar] [CrossRef]
  81. Zha, Z.; Yue, X.; Ren, Q.; Dai, Z. Uniform polypyrrole nanoparticles with high photothermal conversion efficiency for photothermal ablation of cancer cells. Adv. Mater. 2013, 25, 777–782. [Google Scholar] [CrossRef]
  82. Wu, X.; Jiang, Q.; Ghim, D.; Singamaneni, S.; Jun, Y.-S. Localized heating with a photothermal polydopamine coating facilitates a novel membrane distillation process. J. Mater. Chem. A 2018, 6, 18799–18807. [Google Scholar] [CrossRef]
  83. Wang, X.; Liu, Q.; Wu, S.; Xu, B.; Xu, H. Multilayer Polypyrrole Nanosheets with Self-Organized Surface Structures for Flexible and Efficient Solar-Thermal Energy Conversion. Adv. Mater. 2019, 31, e1807716. [Google Scholar] [CrossRef] [PubMed]
  84. Chen, Z.; Luo, Y.; Li, Q.; Chen, X. Microgroove-Structured PDA/PEI/PPy@PI-MS Photothermal Aerogel with a Multilevel Water Transport Network for Highly Salt-Rejecting Solar-Driven Interfacial Evaporation. ACS Appl. Mater. Interfaces 2021, 13, 40531–40542. [Google Scholar] [CrossRef] [PubMed]
  85. Yu, H.; Sun, C.; Xuan, Y.; Zhang, K.; Chang, K. Full solar spectrum driven plasmonic-assisted efficient photocatalytic CO2 reduction to ethanol. Chem. Eng. J. 2022, 430, 132940. [Google Scholar] [CrossRef]
  86. Lu, C.; Li, X.; Li, J.; Mao, L.; Zhu, M.; Chen, Q.; Wen, L.; Li, B.; Guo, T.; Lou, Z. Nonmetallic surface plasmon resonance coupling with pyroelectric effect for enhanced near-infrared-driven CO2 reduction. Chem. Eng. J. 2022, 445, 136739. [Google Scholar] [CrossRef]
  87. Huang, W.; Zhang, L.; Li, Z.; Zhang, X.; Dong, X.; Zhang, Y. Efficient CO2 reduction with H2O via photothermal chemical reaction based on Au-MgO dual catalytic site on TiO2. J. CO2 Util. 2022, 55, 101801. [Google Scholar] [CrossRef]
  88. Han, R.; Chen, L.; Xing, B.; Guo, Q.; Tian, J.; Sha, N.; Zhao, Z. Pr3+-doped La1-xPrxMn0.6Ni0.4O3-δ as efficient artificial photosynthesis catalysts for solar methanol. Catal. Commun. 2022, 165, 106440. [Google Scholar] [CrossRef]
  89. Fiorenza, R.; Bellardita, M.; Balsamo, S.A.; Spitaleri, L.; Gulino, A.; Condorelli, M.; D’Urso, L.; Scirè, S.; Palmisano, L. A solar photothermocatalytic approach for the CO2 conversion: Investigation of different synergisms on CoO-CuO/brookite TiO2-CeO2 catalysts. Chem. Eng. J. 2022, 428, 131249. [Google Scholar] [CrossRef]
  90. Bian, H.; Liu, T.; Li, D.; Xu, Z.; Lian, J.; Chen, M.; Yan, J.; Frank Liu, S. Unveiling the effect of interstitial dopants on CO2 activation over CsPbBr3 catalyst for efficient photothermal CO2 reduction. Chem. Eng. J. 2022, 435, 135071. [Google Scholar] [CrossRef]
  91. Zheng, Y.; Zhang, L.; Guan, J.; Qian, S.; Zhang, Z.; Ngaw, C.K.; Wan, S.; Wang, S.; Lin, J.; Wang, Y. Controlled Synthesis of Cu0/Cu2O for Efficient Photothermal Catalytic Conversion of CO2 and H2O. Acs Sustain. Chem. Eng. 2021, 9, 1754–1761. [Google Scholar] [CrossRef]
  92. Wang, Y.; Zhou, Q.; Zhu, Y.; Xu, D. High efficiency reduction of CO2 to CO and CH4 via photothermal synergistic catalysis of lead-free perovskite Cs3Sb2I9. Appl. Catal. B Environ. 2021, 294, 120236. [Google Scholar] [CrossRef]
  93. Li, Y.; Wen, M.; Wang, Y.; Tian, G.; Wang, C.; Zhao, J. Plasmonic Hot Electrons from Oxygen Vacancies for Infrared Light-Driven Catalytic CO2 Reduction on Bi2O3-x. Angew. Chem. Int. Ed. 2021, 60, 910–916. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, H.; Wang, Y.; Guo, L.; Zhang, X.; Ribeiro, C.; He, T. Solar-heating boosted catalytic reduction of CO2 under full-solar spectrum. Chin. J. Catal. 2020, 41, 131–139. [Google Scholar] [CrossRef]
  95. Yu, F.; Wang, C.; Li, Y.; Ma, H.; Wang, R.; Liu, Y.; Suzuki, N.; Terashima, C.; Ohtani, B.; Ochiai, T.; et al. Enhanced Solar Photothermal Catalysis over Solution Plasma Activated TiO2. Adv. Sci. 2020, 7, 2000204. [Google Scholar] [CrossRef]
  96. Yang, F.; Li, Z.; Xie, Y.; Wang, S.; Li, M.; Liao, L.; Zhou, W. Ag/polydopamine nanoparticles co-decorated defective mesoporous carbon nitride nanosheets assemblies for wide spectrum response and robust photothermal-photocatalytic performance. Appl. Surf. Sci. 2022, 598, 153895. [Google Scholar] [CrossRef]
  97. Li, B.; Ding, Y.; Li, Q.; Guan, Z.; Zhang, M.; Yang, J. The photothermal effect enhance visible light-driven hydrogen evolution using urchin-like hollow RuO2/TiO2/Pt/C nanomaterial. J. Alloys Compd. 2022, 890, 161722. [Google Scholar] [CrossRef]
  98. Guo, X.; Li, J.; Wang, Y.; Rui, Z. Photothermocatalytic water splitting over Pt/ZnIn2S4 for hydrogen production without external heat. Catal. Today 2022, 402, 210–219. [Google Scholar] [CrossRef]
  99. Cheng, X.; Wang, L.; Xie, L.; Sun, C.; Zhao, W.; Liu, X.; Zhuang, Z.; Liu, S.; Zhao, Q. Defect-driven selective oxidation of MoS2 nanosheets with photothermal effect for Photo-Catalytic hydrogen evolution reaction. Chem. Eng. J. 2022, 439, 135757. [Google Scholar] [CrossRef]
  100. Li, Y.; Xue, J.; Shen, Q.; Jia, S.; Li, Q.; Li, Y.; Liu, X.; Jia, H. Construction of a ternary spatial junction in yolk–shell nanoreactor for efficient photo-thermal catalytic hydrogen generation. Chem. Eng. J. 2021, 423, 130188. [Google Scholar] [CrossRef]
  101. Li, X.; Zhao, S.; Duan, X.; Zhang, H.; Yang, S.-z.; Zhang, P.; Jiang, S.P.; Liu, S.; Sun, H.; Wang, S. Coupling hydrothermal and photothermal single-atom catalysis toward excellent water splitting to hydrogen. Appl. Catal. B Environ. 2021, 283, 119660. [Google Scholar] [CrossRef]
  102. Li, M.; Sun, J.; Chen, G.; Yao, S.; Cong, B.; Liu, P. Construction photothermal/pyroelectric property of hollow FeS2/Bi2S3 nanostructure with enhanced full spectrum photocatalytic activity. Appl. Catal. B Environ. 2021, 298, 120573. [Google Scholar] [CrossRef]
  103. Guo, S.; Li, X.; Li, J.; Wei, B. Boosting photocatalytic hydrogen production from water by photothermally induced biphase systems. Nat. Commun. 2021, 12, 1343. [Google Scholar] [CrossRef] [PubMed]
  104. Ng, S.W.L.; Gao, M.; Lu, W.; Hong, M.; Ho, G.W. Selective Wavelength Enhanced Photochemical and Photothermal H2 Generation of Classical Oxide Supported Metal Catalyst. Adv. Funct. Mater. 2021, 31, 2104750. [Google Scholar] [CrossRef]
  105. Luo, X.; Li, R.; Homewood, K.P.; Chen, X.; Gao, Y. Hybrid 0D/2D Ni2P quantum dot loaded TiO2(B) nanosheet photothermal catalysts for enhanced hydrogen evolution. Appl. Surf. Sci. 2020, 505, 144099. [Google Scholar] [CrossRef]
  106. Ren, H.; Yang, J.-L.; Yang, W.-M.; Zhong, H.-L.; Lin, J.-S.; Radjenovic, P.M.; Sun, L.; Zhang, H.; Xu, J.; Tian, Z.-Q.; et al. Core–Shell–Satellite Plasmonic Photocatalyst for Broad-Spectrum Photocatalytic Water Splitting. ACS Mater. Lett. 2020, 3, 69–76. [Google Scholar] [CrossRef]
  107. Zhang, M.; Hu, Q.; Ma, K.; Ding, Y.; Li, C. Pyroelectric effect in CdS nanorods decorated with a molecular Co-catalyst for hydrogen evolution. Nano Energy 2020, 73, 104810. [Google Scholar] [CrossRef]
  108. Dai, B.; Fang, J.; Yu, Y.; Sun, M.; Huang, H.; Lu, C.; Kou, J.; Zhao, Y.; Xu, Z. Construction of Infrared-Light-Responsive Photoinduced Carriers Driver for Enhanced Photocatalytic Hydrogen Evolution. Adv. Mater. 2020, 32, e1906361. [Google Scholar] [CrossRef]
  109. Yang, F.; Wang, S.; Li, Z.; Xu, Y.; Yang, W.; Yv, C.; Yang, D.; Xie, Y.; Zhou, W. Polydopamine/defective ultrathin mesoporous graphitic carbon nitride nanosheets as Z-scheme organic assembly for robust photothermal-photocatalytic performance. J. Colloid Interface Sci. 2022, 613, 775–785. [Google Scholar] [CrossRef]
  110. Wang, R.; Shan, G.; Wang, T.; Yin, D.; Chen, Y. Photothermal enhanced photocatalytic activity based on Ag-doped CuS nanocomposites. J. Alloys Compd. 2021, 864, 158591. [Google Scholar] [CrossRef]
  111. Shi, S.; Han, X.; Liu, J.; Lan, X.; Feng, J.; Li, Y.; Zhang, W.; Wang, J. Photothermal-boosted effect of binary CuFe bimetallic magnetic MOF heterojunction for high-performance photo-Fenton degradation of organic pollutants. Sci. Total Environ. 2021, 795, 148883. [Google Scholar] [CrossRef] [PubMed]
  112. Liu, X.; Wang, W.; Lin, H.; Shen, Y.; Fang, Q.; Liu, F. Construction of hierarchical Prussian blue microcrystal with high sunlight absorption for efficient photo-thermal degradation of organic pollutants. Sep. Purif. Technol. 2021, 269, 118724. [Google Scholar] [CrossRef]
  113. Li, Y.; Chang, H.; Wang, Z.; Shen, Q.; Liu, X.; Xue, J.; Jia, H. A 3D C@TiO2 multishell nanoframe for simultaneous photothermal catalytic hydrogen generation and organic pollutant degradation. J. Colloid Interface Sci. 2022, 609, 535–546. [Google Scholar] [CrossRef]
  114. Li, J.; Zhang, J.; Jiao, X.; Shi, J.; Ye, J.; Song, K.; Bao, J.; Li, G.; Lei, K. NIR-driven PtCu-alloy nanocages via photothermal enhanced fenton catalytic degradation of pollutant dyes under neutral pH. J. Alloys Compd. 2022, 895, 162624. [Google Scholar] [CrossRef]
  115. Lin, E.; Huang, R.; Wu, J.; Kang, Z.; Ke, K.; Qin, N.; Bao, D. Recyclable CoFe2O4 modified BiOCl hierarchical microspheres utilizing photo, photothermal and mechanical energy for organic pollutant degradation. Nano Energy 2021, 89, 106403. [Google Scholar] [CrossRef]
  116. Chen, J.; Luo, W.; Yu, S.; Yang, X.; Wu, Z.; Zhang, H.; Gao, J.; Mai, Y.-W.; Li, Y.; Jia, Y. Synergistic effect of photocatalysis and pyrocatalysis of pyroelectric ZnSnO3 nanoparticles for dye degradation. Ceram. Int. 2020, 46, 9786–9793. [Google Scholar] [CrossRef]
  117. Tian, J.; Wu, S.; Liu, S.; Zhang, W. Photothermal enhancement of highly efficient photocatalysis with bioinspired thermal radiation balance characteristics. Appl. Surf. Sci. 2022, 592, 153304. [Google Scholar] [CrossRef]
  118. Okeil, S.; Yadav, S.; Bruns, M.; Zintler, A.; Molina-Luna, L.; Schneider, J.J. Photothermal catalytic properties of layered titanium chalcogenide nanomaterials. Dalton Trans. 2020, 49, 1032–1047. [Google Scholar] [CrossRef]
  119. Zhu, Q.; Song, J.; Liu, Z.; Wu, K.; Li, X.; Chen, Z.; Pang, H. Photothermal catalytic degradation of textile dyes by laccase immobilized on Fe3O4@SiO2 nanoparticles. J. Colloid. Interf. Sci. 2022, 623, 992–1001. [Google Scholar] [CrossRef]
  120. Qi, Y.; Yang, Z.; Jiang, Y.; Han, H.; Wu, T.; Wu, L.; Liu, J.; Wang, Z.; Wang, F. Platinum−Copper Bimetallic Nanoparticles Supported on TiO2 as Catalysts for Photo−thermal Catalytic Toluene Combustion. ACS Appl. Nano Mater. 2022, 5, 1845–1854. [Google Scholar] [CrossRef]
  121. Li, J.; Feng, J.; Guo, X.; Fang, H.; Chen, J.; Ma, C.; Li, R.; Wang, Y.; Rui, Z. Defect-band bridge photothermally activates Type III heterojunction for CO2 reduction and typical VOCs oxidation. Appl. Catal. B Environ. 2022, 309, 121248. [Google Scholar] [CrossRef]
  122. Li, A.; Zhang, Q.; Zhao, S.; Chong, Y.; Wu, P.; Li, Y.; Jin, X.; Chen, G.; Qiu, Y.; Yang, S.; et al. A dual plasmonic core—shell Pt/[TiN@TiO2] catalyst for enhanced photothermal synergistic catalytic activity of VOCs abatement. Nano Res. 2022, 15, 7071–7080. [Google Scholar] [CrossRef]
  123. Xie, A.; Wang, H.; Qi, S.; Li, X.; Zhu, Z.; Zhang, W.; Wang, Q.; Tang, Y.; Luo, S. Mesoporous SmMnO3/CuMnO catalyst for photothermal synergistic degradation of gaseous toluene. Ceram. Int. 2021, 47, 31485–31496. [Google Scholar] [CrossRef]
  124. Li, J.; Zhang, M.; Elimian, E.A.; Lv, X.; Chen, J.; Jia, H. Convergent ambient sunlight-powered multifunctional catalysis for toluene abatement over in situ exsolution of Mn3O4 on perovskite parent. Chem. Eng. J. 2021, 412, 128560. [Google Scholar] [CrossRef]
  125. Xiao, Y.; Wang, K.; Yang, Z.; Xing, Z.; Li, Z.; Pan, K.; Zhou, W. Plasma Cu-decorated TiO2-x/CoP particle-level hierarchical heterojunctions with enhanced photocatalytic-photothermal performance. J. Hazard. Mater. 2021, 414, 125487. [Google Scholar] [CrossRef]
  126. Jiang, H.; Xing, Z.; Zhao, T.; Yang, Z.; Wang, K.; Li, Z.; Yang, S.; Xie, L.; Zhou, W. Plasmon Ag nanoparticle/Bi2S3 ultrathin nanobelt/oxygen-doped flower-like MoS2 nanosphere ternary heterojunctions for promoting charge separation and enhancing solar-driven photothermal and photocatalytic performances. Appl. Catal. B Environ. 2020, 274, 118947. [Google Scholar] [CrossRef]
  127. Guo, M.; Zhao, T.; Xing, Z.; Qiu, Y.; Pan, K.; Li, Z.; Yang, S.; Zhou, W. Hollow Octahedral Cu2-xS/CdS/Bi2S3 p-n-p Type Tandem Heterojunctions for Efficient Photothermal Effect and Robust Visible-Light-Driven Photocatalytic Performance. ACS Appl. Mater. Interfaces 2020, 12, 40328–40338. [Google Scholar] [CrossRef]
  128. Guo, M.; Xing, Z.; Zhao, T.; Qiu, Y.; Tao, B.; Li, Z.; Zhou, W. Hollow flower-like polyhedral α-Fe2O3/Defective MoS2/Ag Z-scheme heterojunctions with enhanced photocatalytic-Fenton performance via surface plasmon resonance and photothermal effects. Appl. Catal. B Environ. 2020, 272, 118978. [Google Scholar] [CrossRef]
  129. Lee, H.; Park, S.H.; Park, Y.-K.; Kim, S.-J.; Seo, S.-G.; Ki, S.J.; Jung, S.-C. Photocatalytic reactions of 2,4-dichlorophenoxyacetic acid using a microwave-assisted photocatalysis system. Chem. Eng. J. 2015, 278, 259–264. [Google Scholar] [CrossRef]
  130. Li, M.; Xing, Z.; Zhang, Z.; Wang, Y.; Liu, M.; Li, Z.; Wang, N.; Zhou, W. Hollow core-shell Z-scheme heterojunction on self-floating carbon fiber cloth with robust photocatalytic-photothermal performance. J. Clean. Prod. 2022, 360, 132166. [Google Scholar] [CrossRef]
  131. Qiu, Y.; Xing, Z.; Guo, M.; Li, Z.; Wang, N.; Zhou, W. Hollow cubic Cu2-xS/Fe-POMs/AgVO3 dual Z-scheme heterojunctions with wide-spectrum response and enhanced photothermal and photocatalytic-fenton performance. Appl. Catal. B Environ. 2021, 298, 120628. [Google Scholar] [CrossRef]
  132. Yuan, Y.; Pan, W.-g.; Guo, R.-t.; Hong, L.-f.; Lin, Z.-d.; Ji, X.-y. Flower spherical-like Bi7O9I3/AgI S-scheme heterojunction for phenol photodegradation: The synergetic effect of dual surface plasmon resonance and photothermal property. Sep. Purif. Technol. 2022, 297, 121538. [Google Scholar] [CrossRef]
  133. Wang, Z.; Wang, Y.; Zhang, Y.; Sun, X.; Lou, Y.; Zhang, Y.; Dong, Y.; Pan, C.; Zhu, Y. Efficient photothermal degradation on Bi12CoO20 sillenite with a strong internal electric field induced by the thermal effect. Appl. Catal. B Environ. 2022, 313, 121452. [Google Scholar] [CrossRef]
  134. Zheng, Y.; Tan, J.; Zhang, G.; Ma, Y.; Liu, F.; Liu, M.; Wang, Y.; Zou, H.; Huang, H. Yolk-Shell AuAgPt Alloy Nanostructures with Tunable Morphologies: Plasmon-Enhanced Photothermal and Catalytic Properties. Adv. Energy Sustain. Res. 2022, 3, 2100222. [Google Scholar] [CrossRef]
  135. Wang, R.; Hao, Q.; Feng, J.; Wang, G.-C.; Ding, H.; Chen, D.; Ni, B. Enhanced separation of photogenerated charge carriers and catalytic properties of ZnO-MnO2 composites by microwave and photothermal effect. J. Alloys Compd. 2019, 786, 418–427. [Google Scholar] [CrossRef]
  136. Borjigin, B.; Ding, L.; Li, H.; Wang, X. A solar light-induced photo-thermal catalytic decontamination of gaseous benzene by using Ag/Ag3PO4/CeO2 heterojunction. Chem. Eng. J. 2020, 402, 126070. [Google Scholar] [CrossRef]
  137. Zeng, X.; Shan, C.; Sun, M.; Ding, D.; Rong, S. Graphene enhanced α-MnO2 for photothermal catalytic decomposition of carcinogen formaldehyde. Chin. Chem. Lett. 2022, 33, 4771–4775. [Google Scholar] [CrossRef]
  138. Zeng, Y.; Zhong, J.; Wang, H.; Fu, M.; Ye, D.; Hu, Y. Synergistic effect of tunable oxygen-vacancy defects and graphene on accelerating the photothermal degradation of methanol over Co3O4/rGO nanocomposites. Chem. Eng. J. 2021, 425, 131658. [Google Scholar] [CrossRef]
  139. Pang, Y.; Kong, L.; Lei, H.; Chen, D.; Yuvaraja, G. Combined microwave-induced and photocatalytic oxidation using zinc ferrite catalyst for efficient degradation of tetracycline hydrochloride in aqueous solution. J. Taiwan Inst. Chem. E 2018, 93, 397–404. [Google Scholar] [CrossRef]
  140. Barik, A.J.; Kulkarni, S.V.; Gogate, P.R. Degradation of 4-chloro 2-aminophenol using combined approaches based on microwave and photocatalysis. Sep. Purif. Technol. 2016, 168, 152–160. [Google Scholar] [CrossRef]
  141. Liu, X.; Chen, T.; Xue, Y.; Fan, J.; Shen, S.; Hossain, M.S.A.; Amin, M.A.; Pan, L.; Xu, X.; Yamauchi, Y. Nanoarchitectonics of MXene/semiconductor heterojunctions toward artificial photosynthesis via photocatalytic CO2 reduction. Coordin. Chem. Rev. 2022, 459, 214440. [Google Scholar] [CrossRef]
  142. Zhou, Y.; Wang, Z.; Huang, L.; Zaman, S.; Lei, K.; Yue, T.; Li, Z.A.; You, B.; Xia, B.Y. Engineering 2D Photocatalysts toward Carbon Dioxide Reduction. Adv. Energy Mater. 2021, 11, 2003159. [Google Scholar] [CrossRef]
  143. Yamaguchi, Y.; Kudo, A. Visible light responsive photocatalysts developed by substitution with metal cations aiming at artificial photosynthesis. Front. Energy 2021, 15, 568–576. [Google Scholar] [CrossRef]
  144. Fajrina, N.; Tahir, M. A critical review in strategies to improve photocatalytic water splitting towards hydrogen production. Int. J. Hydrogen Energy 2019, 44, 540–577. [Google Scholar] [CrossRef]
  145. Tahir, M.; Tasleem, S.; Tahir, B. Recent development in band engineering of binary semiconductor materials for solar driven photocatalytic hydrogen production. Int. J. Hydrogen Energy 2020, 45, 15985–16038. [Google Scholar] [CrossRef]
  146. Gopinath, C.S.; Nalajala, N. A scalable and thin film approach for solar hydrogen generation: A review on enhanced photocatalytic water splitting. J. Mater. Chem. A 2021, 9, 1353–1371. [Google Scholar] [CrossRef]
  147. Zheng, H.; Li, X.; Zhu, K.; Liang, P.; Wu, M.; Rao, Y.; Jian, R.; Shi, F.; Wang, J.; Yan, K.; et al. Semiconducting BaTiO3@C core-shell structure for improving piezo-photocatalytic performance. Nano Energy 2022, 93, 106831. [Google Scholar] [CrossRef]
  148. Shi, J.; Xie, Z.; Tang, X.; Wang, Y.; Yuan, G.; Liu, J.-M. Enhanced piezo-photocatalytic performance of Ag@Na0.5Bi0.5TiO3 composites. J. Alloys Compd. 2022, 911, 164885. [Google Scholar] [CrossRef]
  149. Shen, W.; Li, N.; Zuo, S.; Wu, M.; Sun, G.; Li, Q.; Shi, M.; Ma, J. Remarkably enhanced piezo-photocatalytic performance of Z-scheme Bi2WO6/Black TiO2 heterojunction via piezoelectric effect. Ceram. Int. 2022, 48, 15899–15907. [Google Scholar] [CrossRef]
  150. Kang, Z.; Ke, K.; Lin, E.; Qin, N.; Wu, J.; Huang, R.; Bao, D. Piezoelectric polarization modulated novel Bi2WO6/g-C3N4/ZnO Z-scheme heterojunctions with g-C3N4 intermediate layer for efficient piezo-photocatalytic decomposition of harmful organic pollutants. J. Colloid Interface Sci. 2022, 607, 1589–1602. [Google Scholar] [CrossRef]
  151. Gotipamul, P.P.; Vattikondala, G.; Rajan, K.D.; Khanna, S.; Rathinam, M.; Chidambaram, S. Impact of piezoelectric effect on the heterogeneous visible photocatalysis of g-C3N4/Ag/ZnO tricomponent. Chemosphere 2022, 287, 132298. [Google Scholar] [CrossRef] [PubMed]
  152. 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] [PubMed]
  153. Xu, T.; Niu, P.; Wang, S.; Li, L. High visible light photocatalytic activities obtained by integrating g-C3N4 with ferroelectric PbTiO3. J. Mater. Sci. Technol. 2021, 74, 128–135. [Google Scholar] [CrossRef]
  154. Sun, H.; Xu, Z.; Xie, X.; Niu, J.; Wang, M.; Zhang, X.; Chen, X.; Han, J. Enhanced photocatalytic activity of ferroelectric-based Ag2O/Bi4Ti3O12 hybrids by piezoelectric effect. J. Alloys Compd. 2021, 882, 160609. [Google Scholar] [CrossRef]
  155. Liu, X.; Shen, X.; Sa, B.; Zhang, Y.; Li, X.; Xue, H. Piezotronic-enhanced photocatalytic performance of heterostructured BaTiO3/SrTiO3 nanofibers. Nano Energy 2021, 89, 106391. [Google Scholar] [CrossRef]
  156. Liu, J.; Chen, J.; Wu, Z.; Zhu, K.; Wang, J.; Li, Z.; Tai, G.; Liu, X.; Lu, S. Enhanced visible-light photocatalytic performances of ZnO through loading AgI and coupling piezo-photocatalysis. J. Alloys Compd. 2021, 852, 156848. [Google Scholar] [CrossRef]
  157. Lei, H.; Wu, M.; Liu, Y.; Mo, F.; Chen, J.; Ji, S.; Zou, Y.; Dong, X. Built-in piezoelectric field improved photocatalytic performance of nanoflower-like Bi2WO6 using low-power white LEDs. Chin. Chem. Lett. 2021, 32, 2317–2321. [Google Scholar]
  158. Jiang, X.; Wang, H.; Wang, X.; Yuan, G. Synergetic effect of piezoelectricity and Ag deposition on photocatalytic performance of barium titanate perovskite. Sol. Energy 2021, 224, 455–461. [Google Scholar] [CrossRef]
  159. Dilly Rajan, K.; Gotipamul, P.P.; Khanna, S.; Chidambaram, S.; Rathinam, M. Piezo-photocatalytic effect of NaNbO3 interconnected nanoparticles decorated CuBi2O4 nanocuboids. Mater. Lett. 2021, 296, 129902. [Google Scholar] [CrossRef]
  160. Wu, J.; Wang, W.; Tian, Y.; Song, C.; Qiu, H.; Xue, H. Piezotronic effect boosted photocatalytic performance of heterostructured BaTiO3/TiO2 nanofibers for degradation of organic pollutants. Nano Energy 2020, 77, 105122. [Google Scholar]
  161. Wu, W.; Yin, X.; Dai, B.; Kou, J.; Ni, Y.; Lu, C. Water flow drived piezo-photocatalytic flexible films: Bi-piezoelectric integration of ZnO nanorods and PVDF. Appl. Surf. Sci. 2020, 517, 146119. [Google Scholar] [CrossRef]
  162. Ren, Z.; Li, X.; Guo, L.; Wu, J.; Li, Y.; Liu, W.; Li, P.; Fu, Y.; Ma, J. Facile synthesis of ZnO/ZnS heterojunction nanoarrays for enhanced piezo-photocatalytic performance. Mater. Lett. 2021, 292, 129635. [Google Scholar] [CrossRef]
  163. Dursun, S.; Akyildiz, H.; Kalem, V. PMN-PT nanoparticle/SnO2 nanofiber heterostructures: Enhanced photocatalytic degradation performance by ultrasonic wave induced piezoelectric field. J. Alloys Compd. 2021, 889, 161769. [Google Scholar] [CrossRef]
  164. Cheng, T.; Gao, H.; Li, R.; Wang, S.; Yi, Z.; Yang, H. Flexoelectricity-induced enhancement in carrier separation and photocatalytic activity of a photocatalyst. Appl. Surf. Sci. 2021, 566, 150669. [Google Scholar] [CrossRef]
  165. Yu, Y.; Yao, B.; He, Y.; Cao, B.; Ren, Y.; Sun, Q. Piezo-enhanced photodegradation of organic pollutants on Ag3PO4/ZnO nanowires using visible light and ultrasonic. Appl. Surf. Sci. 2020, 528, 146819. [Google Scholar] [CrossRef]
  166. Fu, Y.; Wang, Y.; Zhao, H.; Zhang, Z.; An, B.; Bai, C.; Ren, Z.; Wu, J.; Li, Y.; Liu, W.; et al. Synthesis of ternary ZnO/ZnS/MoS2 piezoelectric nanoarrays for enhanced photocatalytic performance by conversion of dual heterojunctions. Appl. Surf. Sci. 2021, 556, 149695. [Google Scholar] [CrossRef]
  167. Yu, C.; Tan, M.; Tao, C.; Hou, Y.; Liu, C.; Meng, H.; Su, Y.; Qiao, L.; Bai, Y. Remarkably enhanced piezo-photocatalytic performance in BaTiO3/CuO heterostructures for organic pollutant degradation. J. Adv. Ceram. 2022, 11, 414–426. [Google Scholar] [CrossRef]
  168. Zhou, X.; Yan, F.; Wu, S.; Shen, B.; Zeng, H.; Zhai, J. Remarkable Piezophoto Coupling Catalysis Behavior of BiOX/BaTiO3 (X = Cl, Br, Cl0.166 Br0.834) Piezoelectric Composites. Small 2020, 16, e2001573. [Google Scholar] [CrossRef]
  169. Wang, R.; Xie, X.; Xu, C.; Lin, Y.; You, D.; Chen, J.; Li, Z.; Shi, Z.; Cui, Q.; Wang, M. Bi-piezoelectric effect assisted ZnO nanorods/PVDF-HFP spongy photocatalyst for enhanced performance on degrading organic pollutant. Chem. Eng. J. 2022, 439, 135787. [Google Scholar] [CrossRef]
  170. Fu, Y.; Ren, Z.; Wu, J.; Li, Y.; Liu, W.; Li, P.; Xing, L.; Ma, J.; Wang, H.; Xue, X. Direct Z-scheme heterojunction of ZnO/MoS2 nanoarrays realized by flowing-induced piezoelectric field for enhanced sunlight photocatalytic performances. Appl. Catal. B Environ. 2021, 285, 119785. [Google Scholar] [CrossRef]
  171. Tong, W.; Zhang, Y.; Huang, H.; Xiao, K.; Yu, S.; Zhou, Y.; Liu, L.; Li, H.; Liu, L.; Huang, T.; et al. A highly sensitive hybridized soft piezophotocatalyst driven by gentle mechanical disturbances in water. Nano Energy 2018, 53, 513–523. [Google Scholar] [CrossRef]
  172. Yao, Z.; Sun, H.; Xiao, S.; Hu, Y.; Liu, X.; Zhang, Y. Synergetic piezo-photocatalytic effect in a Bi2MoO6/BiOBr composite for decomposing organic pollutants. Appl. Surf. Sci. 2021, 560, 150037. [Google Scholar] [CrossRef]
  173. Zhu, Q.; Dar, A.A.; Zhou, Y.; Zhang, K.; Qin, J.; Pan, B.; Lin, J.; Patrocinio, A.O.T.; Wang, C. Oxygen Vacancies Promoted Piezoelectricity toward Piezo-Photocatalytic Decomposition of Tetracycline over SrBi4Ti4O15. ACS EST Eng. 2022, 2, 1365–1375. [Google Scholar] [CrossRef]
  174. Zhou, L.; Dai, S.; Xu, S.; She, Y.; Li, Y.; Leveneur, S.; Qin, Y. Piezoelectric effect synergistically enhances the performance of Ti32-oxo-cluster/BaTiO3/CuS p-n heterojunction photocatalytic degradation of pollutants. Appl. Catal. B Environ. 2021, 291, 120019. [Google Scholar] [CrossRef]
  175. Chen, J.; Liao, B.; Liao, X.; Xie, H.; Yu, Y.; Hou, S.; Wang, C.; Fan, X. Strain-Driven Polarized Electric Field-Promoted Photocatalytic Activity in Borate-Based CsCdBO3 Bulk Materials. ACS Appl. Mater. Interfaces 2021, 13, 34202–34212. [Google Scholar] [CrossRef]
  176. Zhang, C.; Li, N.; Chen, D.; Xu, Q.; Li, H.; He, J.; Lu, J. The ultrasonic-induced-piezoelectric enhanced photocatalytic performance of ZnO/CdS nanofibers for degradation of bisphenol A. J. Alloys Compd. 2021, 885, 160987. [Google Scholar] [CrossRef]
  177. Zhang, C.; Fei, W.; Wang, H.; Li, N.; Chen, D.; Xu, Q.; Li, H.; He, J.; Lu, J. p-n Heterojunction of BiOI/ZnO nanorod arrays for piezo-photocatalytic degradation of bisphenol A in water. J. Hazard. Mater. 2020, 399, 123109. [Google Scholar] [CrossRef]
  178. Feng, Y.; Li, H.; Ling, L.; Yan, S.; Pan, D.; Ge, H.; Li, H.; Bian, Z. Enhanced Photocatalytic Degradation Performance by Fluid-Induced Piezoelectric Field. Environ. Sci. Technol. 2018, 52, 7842–7848. [Google Scholar] [CrossRef]
  179. Hu, J.; Chen, Y.; Zhou, Y.; Zeng, L.; Huang, Y.; Lan, S.; Zhu, M. Piezo-enhanced charge carrier separation over plasmonic Au-BiOBr for piezo-photocatalytic carbamazepine removal. Appl. Catal. B Environ. 2022, 311, 121369. [Google Scholar] [CrossRef]
  180. Li, Y.; Li, R.; Zhai, Y.; Huang, Y.; Lee, S.; Cao, J. Improved photocatalytic activity of BaTiO3/La2Ti2O7 heterojunction composites via piezoelectric-enhanced charge transfer. Appl. Surf. Sci. 2021, 570, 151146. [Google Scholar] [CrossRef]
  181. Li, M.; Sun, J.; Chen, G.; Yao, S.; Cong, B. Construction double electric field of sulphur vacancies as medium ZnS/Bi2S3-PVDF self-supported recoverable piezoelectric film photocatalyst for enhanced photocatalytic performance. Appl. Catal. B Environ. 2022, 301, 120792. [Google Scholar] [CrossRef]
  182. Jiang, Y.; Xie, J.; Lu, Z.; Hu, J.; Hao, A.; Cao, Y. Insight into the effect of OH modification on the piezo-photocatalytic hydrogen production activity of SrTiO3. J. Colloid Interface Sci. 2022, 612, 111–120. [Google Scholar] [CrossRef]
  183. Huang, X.; Lei, R.; Yuan, J.; Gao, F.; Jiang, C.; Feng, W.; Zhuang, J.; Liu, P. Insight into the piezo-photo coupling effect of PbTiO3/CdS composites for piezo-photocatalytic hydrogen production. Appl. Catal. B Environ. 2021, 282, 119586. [Google Scholar] [CrossRef]
  184. Hu, C.; Chen, F.; Wang, Y.; Tian, N.; Ma, T.; Zhang, Y.; Huang, H. Exceptional Cocatalyst-Free Photo-Enhanced Piezocatalytic Hydrogen Evolution of Carbon Nitride Nanosheets from Strong In-Plane Polarization. Adv. Mater. 2021, 33, e2101751. [Google Scholar] [CrossRef]
  185. Cui, Y.; Wang, Z.; Li, B.; Yan, Y.; Xu, R.; Meng, M.; Yan, Y. Fluid-induced piezoelectric field enhancing photocatalytic hydrogen evolution reaction on g-C3N4/LiNbO3/PVDF membrane. Nano Energy 2022, 99, 107429. [Google Scholar] [CrossRef]
  186. Tang, R.; Gong, D.; Zhou, Y.; Deng, Y.; Feng, C.; Xiong, S.; Huang, Y.; Peng, G.; Li, L.; Zhou, Z. Unique g-C3N4/PDI-g-C3N4 homojunction with synergistic piezo-photocatalytic effect for aquatic contaminant control and H2O2 generation under visible light. Appl. Catal. B Environ. 2022, 303, 120929. [Google Scholar] [CrossRef]
  187. Zhou, X.; Shen, B.; Zhai, J.; Conesa, J.C. High Performance Generation of H2O2 under Piezophototronic Effect with Multi-Layer In2S3 Nanosheets Modified by Spherical ZnS and BaTiO3 Nanopiezoelectrics. Small Methods 2021, 5, e2100269. [Google Scholar] [CrossRef] [PubMed]
  188. Mohanty, R.; Mansingh, S.; Parida, K.; Parida, K. Boosting sluggish photocatalytic hydrogen evolution through piezo-stimulated polarization: A critical review. Mater. Horiz. 2022, 9, 1332–1355. [Google Scholar] [CrossRef]
  189. Yan, X.; Li, G.; Wang, Z.; Yu, Z.; Wang, K.; Wu, Y. Recent progress on piezoelectric materials for renewable energy conversion. Nano Energy 2020, 77, 105180. [Google Scholar] [CrossRef]
  190. Pan, L.; Sun, S.; Chen, Y.; Wang, P.; Wang, J.; Zhang, X.; Zou, J.J.; Wang, Z.L. Advances in Piezo-Phototronic Effect Enhanced Photocatalysis and Photoelectrocatalysis. Adv. Energy Mater. 2020, 10, 2000214. [Google Scholar] [CrossRef]
  191. Liu, Z.; Yu, X.; Li, L. Piezopotential augmented photo- and photoelectro-catalysis with a built-in electric field. Chin. J. Catal. 2020, 41, 534–549. [Google Scholar] [CrossRef]
  192. Lops, C.; Ancona, A.; Di Cesare, K.; Dumontel, B.; Garino, N.; Canavese, G.; Hernandez, S.; Cauda, V. Sonophotocatalytic degradation mechanisms of Rhodamine B dye via radicals generation by micro- and nano-particles of ZnO. Appl. Catal. B Environ. 2019, 243, 629–640. [Google Scholar] [CrossRef] [PubMed]
  193. Li, H.; Sang, Y.; Chang, S.; Huang, X.; Zhang, Y.; Yang, R.; Jiang, H.; Liu, H.; Wang, Z.L. Enhanced ferroelectric-nanocrystal-based hybrid photocatalysis by ultrasonic-wave-generated piezophototronic effect. Nano Lett. 2015, 15, 2372–2379. [Google Scholar] [CrossRef] [PubMed]
  194. Xu, S.; Guo, L.; Sun, Q.; Wang, Z.L. Piezotronic Effect Enhanced Plasmonic Photocatalysis by AuNPs/BaTiO3 Heterostructures. Adv. Funct. Mater. 2019, 29, 1808737. [Google Scholar] [CrossRef]
  195. Liu, Q.; Zhao, W.; Ao, Z.; An, T. Photo-piezoelectric synergistic degradation of typical volatile organic compounds on BaTiO3. Chin. Chem. Lett. 2022, 33, 410–414. [Google Scholar] [CrossRef]
  196. Jiang, Y.; Li, M.; Mi, Y.; Guo, L.; Fang, W.; Zeng, X.; Zhou, T.; Liu, Y. The influence of piezoelectric effect on the heterogeneous photocatalytic hydrogen production of strontium titanate nanoparticles. Nano Energy 2021, 85, 105949. [Google Scholar] [CrossRef]
  197. Zhao, J.; Yang, H.; Li, Y.; Lu, K. Photocatalytic activity of CdS nanoparticles enhanced by the interaction between piezotronic effect and phase junction. J. Alloys Compd. 2020, 815, 152494. [Google Scholar] [CrossRef]
  198. Zhao, Y.; Huang, X.; Gao, F.; Zhang, L.; Tian, Q.; Fang, Z.B.; Liu, P. Study on water splitting characteristics of CdS nanosheets driven by the coupling effect between photocatalysis and piezoelectricity. Nanoscale 2019, 11, 9085–9090. [Google Scholar] [CrossRef]
  199. Li, Y.; Wang, Q.; Wang, H.; Tian, J.; Cui, H. Novel Ag2O nanoparticles modified MoS2 nanoflowers for piezoelectric-assisted full solar spectrum photocatalysis. J. Colloid Interface Sci. 2019, 537, 206–214. [Google Scholar] [CrossRef]
  200. Chou, T.-M.; Chan, S.-W.; Lin, Y.-J.; Yang, P.-K.; Liu, C.-C.; Lin, Y.-J.; Wu, J.-M.; Lee, J.-T.; Lin, Z.-H. A highly efficient Au-MoS2 nanocatalyst for tunable piezocatalytic and photocatalytic water disinfection. Nano Energy 2019, 57, 14–21. [Google Scholar] [CrossRef]
  201. Venkatesan, M.; Chen, W.-C.; Cho, C.-J.; Veeramuthu, L.; Chen, L.-G.; Li, K.-Y.; Tsai, M.-L.; Lai, Y.-C.; Lee, W.-Y.; Chen, W.-C.; et al. Enhanced piezoelectric and photocatalytic performance of flexible energy harvester based on CsZn0.75Pb0.25I3/CNC–PVDF composite nanofibers. Chem. Eng. J. 2022, 433, 133620. [Google Scholar] [CrossRef]
  202. Pan, J.; Li, Y.; Guo, G.; Zhao, X.; Yu, J.; Li, Z.; Xu, S.; Man, B.; Wei, D.; Zhang, C. Synergizing piezoelectric and plasmonic modulation of PVDF/MoS2 cavity/Au for enhanced photocatalysis. Appl. Surf. Sci. 2022, 577, 151811. [Google Scholar] [CrossRef]
  203. Sittipol, W.; Sronsri, C.; U-yen, K. Effect of magnetic fields on the efficiency of the photocatalytic degradation of methylene blue in a dynamic fluid system. J. Clean Prod. 2021, 325, 129284. [Google Scholar] [CrossRef]
  204. Silva, E.C.; Bonacin, J.A.; Passos, R.R.; Pocrifka, L.A. The effect of an external magnetic field on the photocatalytic activity of CoFe2O4 particles anchored in carbon cloth. J. Photoch. Photobio. A 2021, 416, 113317. [Google Scholar] [CrossRef]
  205. Tank, C.M.; Sakhare, Y.S.; Kanhe, N.S.; Nawale, A.B.; Das, A.K.; Bhoraskar, S.V.; Mathe, V.L. Electric field enhanced photocatalytic properties of TiO2 nanoparticles immobilized in porous silicon template. Solid State Sci. 2011, 13, 1500–1504. [Google Scholar] [CrossRef]
  206. Gao, W.; Zhao, X.; Cui, C.; Su, X.; Zhang, S.; Wang, X.; Zhang, X.L.; Sang, Y.; Liu, H. High-efficiency separation and transfer of photo-induced charge carrier in graphene/TiO2 via heterostructure in magnetic field. J. Alloys Compd. 2021, 862, 158283. [Google Scholar] [CrossRef]
  207. Bian, Y.; Zheng, G.; Ding, W.; Hu, L.; Sheng, Z. Magnetic field effect on the photocatalytic degradation of methyl orange by commercial TiO2 powder. RSC Adv. 2021, 11, 6284–6291. [Google Scholar] [CrossRef]
  208. Zhao, C.; Zhou, L.; Zhang, Z.; Gao, Z.; Weng, H.; Zhang, W.; Li, L.; Song, Y.Y. Insight of the Influence of Magnetic-Field Direction on Magneto-Plasmonic Interfaces for Tuning Photocatalytical Performance of Semiconductors. J. Phys. Chem. Lett. 2020, 11, 9931–9937. [Google Scholar] [CrossRef]
  209. Pan, L.; Ai, M.; Huang, C.; Yin, L.; Liu, X.; Zhang, R.; Wang, S.; Jiang, Z.; Zhang, X.; Zou, J.J.; et al. Manipulating spin polarization of titanium dioxide for efficient photocatalysis. Nat. Commun. 2020, 11, 418. [Google Scholar] [CrossRef]
  210. Huang, G.; Zhang, G.; Gao, Z.; Cao, J.; Li, D.; Yun, H.; Zeng, T. Enhanced visible-light-driven photocatalytic activity of BiFeO3 via electric-field control of spontaneous polarization. J. Alloys Compd. 2019, 783, 943–951. [Google Scholar] [CrossRef]
  211. Wang, Y.; Wang, S.; Wu, Y.; Wang, Z.; Zhang, H.; Cao, Z.; He, J.; Li, W.; Yang, Z.; Zheng, L.; et al. A α-Fe2O3/rGO magnetic photocatalyst: Enhanced photocatalytic performance regulated by magnetic field. J. Alloys Compd. 2021, 851, 156733. [Google Scholar] [CrossRef]
  212. Lu, Y.; Ren, B.; Chang, S.; Mi, W.; He, J.; Wang, W. Achieving effective control of the photocatalytic performance for CoFe2O4/MoS2 heterojunction via exerting external magnetic fields. Mater. Lett. 2020, 260, 126979. [Google Scholar] [CrossRef]
  213. Huang, H.; Chen, Y.; Shi, H. Boosting Separation of Charge Carriers in 2D/0D BiOBr Nanoflower Sheets/BN Quantum Dots with the Lorentz Force via Magnetic Field. Energ. Fuel 2022. [Google Scholar] [CrossRef]
  214. Zhao, Y.; Huang, Z.; Chang, W.; Wei, C.; Feng, X.; Ma, L.; Qi, X.; Li, Z. Microwave-assisted solvothermal synthesis of hierarchical TiO2 microspheres for efficient electro-field-assisted-photocatalytic removal of tributyltin in tannery wastewater. Chemosphere 2017, 179, 75–83. [Google Scholar] [CrossRef]
  215. Li, N.; He, M.; Lu, X.; Liang, L.; Li, R.; Yan, B.; Chen, G. Enhanced norfloxacin degradation by visible-light-driven Mn3O4/gamma-MnOOH photocatalysis under weak magnetic field. Sci. Total Environ. 2021, 761, 143268. [Google Scholar] [CrossRef]
  216. Wu, Y.; Li, T.; Ren, X.; Fu, Y.; Zhang, H.; Feng, X.; Huang, H.; Xie, R. Magnetic field assisted alpha-Fe2O3/Zn1-xFexO heterojunctions for accelerating antiviral agents degradation under visible-light. J. Environ. Chem. Eng. 2022, 10, 106990. [Google Scholar] [CrossRef]
  217. Li, Y.; Wang, Z.; Wang, Y.; Kovács, A.; Foo, C.; Dunin-Borkowski, R.E.; Lu, Y.; Taylor, R.A.; Wu, C.; Tsang, S.C.E. Local magnetic spin mismatch promoting photocatalytic overall water splitting with exceptional solar-to-hydrogen efficiency. Energy Environ. Sci. 2022, 15, 265–277. [Google Scholar] [CrossRef]
  218. Zhao, X.; Gao, W.; Liu, Q.; Cui, C.; Zhou, W.; Wang, X.; Zhang, X.L.; Zhao, L.; Sang, Y.; Liu, H. Enhanced photo-induced carrier separation of CdS/MoS2 via micro-potential of Mo microsheet derived from electromagnetic induction. Chem. Eng. J. 2021, 404, 126972. [Google Scholar] [CrossRef]
  219. Gao, W.; Liu, Q.; Zhang, S.; Yang, Y.; Zhang, X.; Zhao, H.; Qin, W.; Zhou, W.; Wang, X.; Liu, H.; et al. Electromagnetic induction derived micro-electric potential in metal-semiconductor core-shell hybrid nanostructure enhancing charge separation for high performance photocatalysis. Nano Energy 2020, 71, 104624. [Google Scholar] [CrossRef]
  220. Pan, H.; Sun, M.; Wang, X.; Zhang, M.; Murugananthan, M.; Zhang, Y. A novel electric-assisted photocatalytic technique using self-doped TiO2 nanotube films. Appl. Catal. B Environ. 2022, 307, 121174. [Google Scholar] [CrossRef]
  221. Park, S.; Lee, C.W.; Kang, M.-G.; Kim, S.; Kim, H.J.; Kwon, J.E.; Park, S.Y.; Kang, C.-Y.; Hong, K.S.; Nam, K.T. A ferroelectric photocatalyst for enhancing hydrogen evolution: Polarized particulate suspension. Phys. Chem. Chem. Phys. 2014, 16, 10408–10413. [Google Scholar] [CrossRef] [PubMed]
  222. Xu, S.C.; Ding, H.L.; Pan, S.S.; Luo, Y.Y.; Li, G.H. Electric-Field-Enhanced Photocatalytic Removal of Cr(VI) under Sunlight of TiO2 Nanograss Mesh with Nondestructive Regeneration and Feasible Collection for Cr(III). Acs Sustain. Chem. Eng. 2016, 4, 6887–6893. [Google Scholar] [CrossRef]
  223. Li, S.; Bai, L.; Ji, N.; Yu, S.; Lin, S.; Tian, N.; Huang, H. Ferroelectric polarization and thin-layered structure synergistically promoting CO2 photoreduction of Bi2MoO6. J. Mater. Chem. A 2020, 8, 9268–9277. [Google Scholar] [CrossRef]
  224. Zhao, Z.; Wang, D.; Gao, R.; Wen, G.; Feng, M.; Song, G.; Zhu, J.; Luo, D.; Tan, H.; Ge, X.; et al. Magnetic-Field-Stimulated Efficient Photocatalytic N2 Fixation over Defective BaTiO3 Perovskites. Angew. Chem. Int. Ed. 2021, 60, 11910–11918. [Google Scholar] [CrossRef]
  225. Gao, W.; Lu, J.; Zhang, S.; Zhang, X.; Wang, Z.; Qin, W.; Wang, J.; Zhou, W.; Liu, H.; Sang, Y. Suppressing Photoinduced Charge Recombination via the Lorentz Force in a Photocatalytic System. Adv. Sci. 2019, 6, 1901244. [Google Scholar] [CrossRef]
  226. Zhao, J.; Li, N.; Yu, R.; Zhao, Z.; Nan, J. Magnetic field enhanced denitrification in nitrate and ammonia contaminated water under 3D/2D Mn2O3/g-C3N4 photocatalysis. Chem. Eng. J. 2018, 349, 530–538. [Google Scholar] [CrossRef]
  227. Peng, C.; Fan, W.; Li, Q.; Han, W.; Chen, X.; Zhang, G.; Yan, Y.; Gu, Q.; Wang, C.; Zhang, H.; et al. Boosting photocatalytic activity through tuning electron spin states and external magnetic fields. J. Mater. Sci. Technol. 2022, 115, 208–220. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of three types of external field-assisted photocatalysis.
Figure 1. Schematic diagram of three types of external field-assisted photocatalysis.
Catalysts 12 01042 g001
Figure 2. (a) H2-evolution rate of 7 wt% Co0.85Se/Mn0.3Cd0.7S at various reaction temperatures (5–25 °C) under visible light for 1 h (Reprinted/adapted with permission from Ref. [28]. Copyright 2021, copyright Elsevier). (b) H2 production rates under light, thermal, and photothermal conditions (10 mg of 1.0%Cu2O-rGO/TiO2 was used.) (Reprinted/adapted with permission from Ref. [29]. Copyright 2022, copyright Elsevier). (c) The achieved product yield of C2H6O over Cu2O/g-C3N4 at different temperatures (Reprinted/adapted with permission from Ref. [30]. Copyright 2021, copyright Elsevier).
Figure 2. (a) H2-evolution rate of 7 wt% Co0.85Se/Mn0.3Cd0.7S at various reaction temperatures (5–25 °C) under visible light for 1 h (Reprinted/adapted with permission from Ref. [28]. Copyright 2021, copyright Elsevier). (b) H2 production rates under light, thermal, and photothermal conditions (10 mg of 1.0%Cu2O-rGO/TiO2 was used.) (Reprinted/adapted with permission from Ref. [29]. Copyright 2022, copyright Elsevier). (c) The achieved product yield of C2H6O over Cu2O/g-C3N4 at different temperatures (Reprinted/adapted with permission from Ref. [30]. Copyright 2021, copyright Elsevier).
Catalysts 12 01042 g002
Figure 3. Schematic illustration of the photocatalytic H2 generation process over SnSe/ZnIn2S4 (Reprinted/adapted with permission from Ref. [31]. Copyright 2021, copyright Elsevier).
Figure 3. Schematic illustration of the photocatalytic H2 generation process over SnSe/ZnIn2S4 (Reprinted/adapted with permission from Ref. [31]. Copyright 2021, copyright Elsevier).
Catalysts 12 01042 g003
Figure 4. Schematic diagram of microwave-assisted photocatalytic degradation (Reprinted/adapted with permission from Ref. [37]. Copyright 2019, copyright Elsevier).
Figure 4. Schematic diagram of microwave-assisted photocatalytic degradation (Reprinted/adapted with permission from Ref. [37]. Copyright 2019, copyright Elsevier).
Catalysts 12 01042 g004
Figure 5. Schematic diagram of three photothermal effects (Reprinted/adapted with permission from Ref. [22]. Copyright 2022, copyright Elsevier).
Figure 5. Schematic diagram of three photothermal effects (Reprinted/adapted with permission from Ref. [22]. Copyright 2022, copyright Elsevier).
Catalysts 12 01042 g005
Figure 6. (a) Photocatalytic H2 evolution activity of pure Zn0.67Cd0.33S and Zn0.67Cd0.33S supported by Au nanoparticles with different spatial arrangements (Au@Zn0.67Cd0.33S) under visible light irradiation (≥420 nm); blue columns represent the enhancement promoted by plasmonic-Au nanostructures. (b) Illustration for the proximity of semiconductors to the electromagnetic fields of plasmonic Au nanoparticles [46].
Figure 6. (a) Photocatalytic H2 evolution activity of pure Zn0.67Cd0.33S and Zn0.67Cd0.33S supported by Au nanoparticles with different spatial arrangements (Au@Zn0.67Cd0.33S) under visible light irradiation (≥420 nm); blue columns represent the enhancement promoted by plasmonic-Au nanostructures. (b) Illustration for the proximity of semiconductors to the electromagnetic fields of plasmonic Au nanoparticles [46].
Catalysts 12 01042 g006
Figure 7. (a) Comparison of experimental and theoretical UV-vis-NIR absorption spectra of CuS nanodisks dispersed in tetrachloroethylene. (b) LSPR curves (curves 1–4) shift with increasing surface-treatment cycles (Reprinted/adapted with permission from Ref. [53]. Copyright 2013, copyright American Chemical Society).
Figure 7. (a) Comparison of experimental and theoretical UV-vis-NIR absorption spectra of CuS nanodisks dispersed in tetrachloroethylene. (b) LSPR curves (curves 1–4) shift with increasing surface-treatment cycles (Reprinted/adapted with permission from Ref. [53]. Copyright 2013, copyright American Chemical Society).
Catalysts 12 01042 g007
Figure 8. (a) Solar spectrum and UV–Vis–NIR absorption spectra of TiOx nanoparticles with different colors (Reprinted/adapted with permission from Ref. [63]. Copyright 2017, copyright Wiley-VCH). (b) UV-vis absorption spectra for TiO2 NTAs and TiO2 PCs and the optical photograph of the TiO2 NTAs, TiO2 PCs, and Papalio nireus (Reprinted/adapted with permission from Ref. [64]. Copyright 2018, copyright American Chemical Society).
Figure 8. (a) Solar spectrum and UV–Vis–NIR absorption spectra of TiOx nanoparticles with different colors (Reprinted/adapted with permission from Ref. [63]. Copyright 2017, copyright Wiley-VCH). (b) UV-vis absorption spectra for TiO2 NTAs and TiO2 PCs and the optical photograph of the TiO2 NTAs, TiO2 PCs, and Papalio nireus (Reprinted/adapted with permission from Ref. [64]. Copyright 2018, copyright American Chemical Society).
Catalysts 12 01042 g008
Figure 9. (a) Experimental absorption spectra of samples synthesized with different content of chitosan and simulated result by LSPR method (the dashed line) (Reprinted/adapted with permission from Ref. [68]. Copyright 2017, copyright Royal Society of Chemistry). (b) UV–vis–NIR spectra, inset: photograph (left, bulk In2O3-x(OH)y; right, 2D black In2O3-x) (Reprinted/adapted with permission from Ref. [69]. Copyright 2020, copyright Wiley-VCH).
Figure 9. (a) Experimental absorption spectra of samples synthesized with different content of chitosan and simulated result by LSPR method (the dashed line) (Reprinted/adapted with permission from Ref. [68]. Copyright 2017, copyright Royal Society of Chemistry). (b) UV–vis–NIR spectra, inset: photograph (left, bulk In2O3-x(OH)y; right, 2D black In2O3-x) (Reprinted/adapted with permission from Ref. [69]. Copyright 2020, copyright Wiley-VCH).
Catalysts 12 01042 g009
Figure 10. (a) Light absorption spectra for the F-Wood and F-Wood/CNTs membranes (Reprinted/adapted with permission from Ref. [75]. Copyright 2017, copyright Wiley-VCH). (b) UV-vis absorption curves of nano-GO and nano-rGO, respectively. The inset shows a zoom-in view of the curves in the 800 nm region (Reprinted/adapted with permission from Ref. [78]. Copyright 2011, copyright American Chemical Society). (c) UV–vis–NIR extinction spectra were recorded for 25 μg mL−1 PVA stabilized PPy nanoparticles dispersed in water and stored at 4 °C for six months. The inset shows the prepared sample (Reprinted/adapted with permission from Ref. [81]. Copyright 2013, copyright Wiley-VCH).
Figure 10. (a) Light absorption spectra for the F-Wood and F-Wood/CNTs membranes (Reprinted/adapted with permission from Ref. [75]. Copyright 2017, copyright Wiley-VCH). (b) UV-vis absorption curves of nano-GO and nano-rGO, respectively. The inset shows a zoom-in view of the curves in the 800 nm region (Reprinted/adapted with permission from Ref. [78]. Copyright 2011, copyright American Chemical Society). (c) UV–vis–NIR extinction spectra were recorded for 25 μg mL−1 PVA stabilized PPy nanoparticles dispersed in water and stored at 4 °C for six months. The inset shows the prepared sample (Reprinted/adapted with permission from Ref. [81]. Copyright 2013, copyright Wiley-VCH).
Catalysts 12 01042 g010
Figure 11. (a) Photothermal catalytic performances at T = 120 °C on the TiO2-CeO2-based samples. (b) Photothermal catalytic mechanism with the CoO-CuO/TiO2-CeO2 catalyst (Reprinted/adapted with permission from Ref. [89]. Copyright 2022, copyright Elsevier).
Figure 11. (a) Photothermal catalytic performances at T = 120 °C on the TiO2-CeO2-based samples. (b) Photothermal catalytic mechanism with the CoO-CuO/TiO2-CeO2 catalyst (Reprinted/adapted with permission from Ref. [89]. Copyright 2022, copyright Elsevier).
Catalysts 12 01042 g011
Figure 12. The schematic diagram illustrates the mechanism for enhanced photocatalytic H2 generation due to photothermal and photochemical effects (Reprinted/adapted with permission from Ref. [104]. Copyright 2021, copyright Wiley-VCH).
Figure 12. The schematic diagram illustrates the mechanism for enhanced photocatalytic H2 generation due to photothermal and photochemical effects (Reprinted/adapted with permission from Ref. [104]. Copyright 2021, copyright Wiley-VCH).
Catalysts 12 01042 g012
Figure 13. Schematic diagram of the photothermal catalytic mechanism of CFO-BiOCl (Reprinted/adapted with permission from Ref. [115]. Copyright 2021, copyright Elsevier).
Figure 13. Schematic diagram of the photothermal catalytic mechanism of CFO-BiOCl (Reprinted/adapted with permission from Ref. [115]. Copyright 2021, copyright Elsevier).
Catalysts 12 01042 g013
Figure 14. Mechanistic diagram of the photo−thermal catalytic degradation of the toluene reaction on Pt−Cu/TiO2 catalysts (Reprinted/adapted with permission from Ref. [120]. Copyright 2022, copyright American Chemical Society).
Figure 14. Mechanistic diagram of the photo−thermal catalytic degradation of the toluene reaction on Pt−Cu/TiO2 catalysts (Reprinted/adapted with permission from Ref. [120]. Copyright 2022, copyright American Chemical Society).
Catalysts 12 01042 g014
Figure 15. Schematic of the piezo-photocatalytic reactor setup (Reprinted/adapted with permission from Ref. [179]. Copyright 2022, copyright Elsevier).
Figure 15. Schematic of the piezo-photocatalytic reactor setup (Reprinted/adapted with permission from Ref. [179]. Copyright 2022, copyright Elsevier).
Catalysts 12 01042 g015
Figure 16. Schematic illustration of the design concept of the hybrid piezophotocatalyst (Reprinted/adapted with permission from Ref. [171]. Copyright 2018, copyright Elsevier).
Figure 16. Schematic illustration of the design concept of the hybrid piezophotocatalyst (Reprinted/adapted with permission from Ref. [171]. Copyright 2018, copyright Elsevier).
Catalysts 12 01042 g016
Figure 17. Mechanism for the photo/piezo-catalysis process of CTOC/BaTiO3/CuS composites under visible-light irradiation (Reprinted/adapted with permission from Ref. [174]. Copyright 2021, copyright Elsevier).
Figure 17. Mechanism for the photo/piezo-catalysis process of CTOC/BaTiO3/CuS composites under visible-light irradiation (Reprinted/adapted with permission from Ref. [174]. Copyright 2021, copyright Elsevier).
Catalysts 12 01042 g017
Figure 18. Schematic diagram of (a) photocatalysis, (b) piezocatalysis, and (c) piezo-photocatalysis for PbTiO3/CdS composites (Reprinted/adapted with permission from Ref. [183]. Copyright 2021, copyright Elsevier).
Figure 18. Schematic diagram of (a) photocatalysis, (b) piezocatalysis, and (c) piezo-photocatalysis for PbTiO3/CdS composites (Reprinted/adapted with permission from Ref. [183]. Copyright 2021, copyright Elsevier).
Catalysts 12 01042 g018
Figure 19. (a) Polarization schematic diagram of PVDF under ultrasonic conditions. (b) Schematic diagram of the photocatalytic mechanism of ZnS-VS/Bi2S3-PVDF under visible light irradiation (Reprinted/adapted with permission from Ref. [181]. Copyright 2022, copyright Elsevier).
Figure 19. (a) Polarization schematic diagram of PVDF under ultrasonic conditions. (b) Schematic diagram of the photocatalytic mechanism of ZnS-VS/Bi2S3-PVDF under visible light irradiation (Reprinted/adapted with permission from Ref. [181]. Copyright 2022, copyright Elsevier).
Catalysts 12 01042 g019
Figure 20. Diagrammatic sketch of the electro-field-assisted-photocatalytic system (Reprinted/adapted with permission from Ref. [214]. Copyright 2017, copyright Elsevier).
Figure 20. Diagrammatic sketch of the electro-field-assisted-photocatalytic system (Reprinted/adapted with permission from Ref. [214]. Copyright 2017, copyright Elsevier).
Catalysts 12 01042 g020
Figure 21. Schematic diagram of ferroelectric field-assisted photocatalysis of Bi2MoO6 (Reprinted/adapted with permission from Ref. [223]. Copyright 2020, copyright Royal Society of Chemistry).
Figure 21. Schematic diagram of ferroelectric field-assisted photocatalysis of Bi2MoO6 (Reprinted/adapted with permission from Ref. [223]. Copyright 2020, copyright Royal Society of Chemistry).
Catalysts 12 01042 g021
Figure 22. Schematic diagram of magnetic field-promoted photocatalytic water splitting system (Reprinted/adapted with permission from Ref. [217]. Copyright 2022, copyright Royal Society of Chemistry).
Figure 22. Schematic diagram of magnetic field-promoted photocatalytic water splitting system (Reprinted/adapted with permission from Ref. [217]. Copyright 2022, copyright Royal Society of Chemistry).
Catalysts 12 01042 g022
Table 1. Summary of thermal-coupled photocatalysis in recent years.
Table 1. Summary of thermal-coupled photocatalysis in recent years.
MaterialConditionApplicationActivityBlank ControlRef.
SrTiO3/Cu@Ni/TiNXe lamp
(600 mW/cm2)
artificial photosynthesis21.33 μmol g−1 h1
(C2H6O)
380 °C:
1.44 μmol g−1 h−1
(C2H6O)
[85]
BP/WO300 W Xe lamp (UV-vis-NIR)artificial photosynthesis26.1 μmol g−1 h−1
(CO)
BP:
1.47 μmol g−1 h−1
(CO)
[86]
Au-MgO/TiO2300 W Xe lamp (UV-vis-NIR)artificial photosynthesis6.624 μmol g−1 h−1
(CH4)
P25:
2.07 μmol g−1 h−1
(CH4)
[87]
La0.4Pr0.6Mn0.6
Ni0.4O3-δ
300 W Xe lamp
ET: 300 °C
artificial photosynthesis3970 μmol g−1 h−1
(CH4O)
PC:
0.24 μmol g−1 h−1
(CH4O)
[88]
CoO-CuO/TiO2-CeO2300 W Xe lampartificial photosynthesis1.84 μmol g−1 h−1
(CH4)
PC:
0.5 μmol g−1 h−1
(CH4)
[89]
Cu:CsPbBr3300 W Xe lampartificial photosynthesis14.72 μmol g−1 h−1
(CH4)
CsPbBr3:
3.62 μmol g−1 h−1
(CH4)
[90]
Cu0/Cu2O300 W Xe lamp
(λ > 400 nm)
artificial photosynthesis2.6 μmol g−1 h−1
(CH4O)
Cu2O:
0 μmol g−1 h−1
(CH4O)
[91]
Cs3Sb2I9Xe lamp
(200 mW/cm2)
ET: 235 °C
artificial photosynthesis95.7 μmol g−1 h−1
(CO)
PC:
1.1 μmol g−1 h−1
(CO)
[92]
Bi2O3−xLED light
(365–940 nm)
artificial photosynthesisAQY = 0.113%
(940 nm)
AOY = 0.028%
(450 nm)
[93]
Au/rutile300 W Xe lamp (UV-vis-NIR)artificial photosynthesis∼5 μmol g−1 h−1
(CO)
rutile:
∼2.25 μmol g−1 h−1
(CO)
[94]
TiO2(AB)150 W Xe lamp (UV-vis-NIR)artificial photosynthesis11.93 μmol g−1 h−1
(CH4)
PC:
0.09 μmol g−1 h−1
(CH4)
[95]
AuCu/g-C3N4300 W Xe lamp
ET: 120 °C
artificial photosynthesis0.89 mmol g−1 h−1
(C2H6O)
PC:
0.21 mmol g−1 h−1
(C2H6O)
[49]
Cu2O/g-C3N4300 W Xe lamp
(λ > 420 nm)
ET: 100 °C
artificial photosynthesis0.71 mmol g−1 h−1 (C2H6O)PC:
0.37 mmol g−1 h−1
(C2H6O)
[30]
Ag-PDA/DCN300 W Xe lamp
(λ > 420 nm)
H2 production3840 μmol g−1 h−1PDA/DCN:
548 μmol g−1 h−1
[96]
1T-WS2/CdS300 W Xe lamp
(λ > 420 nm)
H2 production70.9 mmol g−1 h−1Pt0.02-CdS:
20.2 mmol g−1 h−1
[33]
TiO2@CS300 W Xe lamp
ET: 120 °C
H2 production22772.6
μmol g−1 h−1
120 °C:
∼1100 μmol g−1 h−1
[71]
RuO2/TiO2/Pt/C300 W Xe lamp
(λ > 420 nm)
H2 production81.62 μmol g−1 h−1RuO2-Pt/TiO2/C:
37.45 μmol g−1 h−1
[97]
Cu2O-rGO/TiO2300 W Xe lamp
ET: 90 °C
H2 production17.8 mmol g−1 h−1PC:
3.8 mmol g−1 h−1
[29]
Pt/ZnIn2S4300 W Xe lamp
(200 mW/cm2)
H2 production19.4 mmol g−1 h−1Co9S8 @ZnIn2S4:
6.25 mmol g−1 h−1
[98]
oxide-MoS2Xe lampsH2 production7.85 mmol g−1 h−12H-MoS2:
2.52 mmol g−1 h−1
[99]
C@TiO2/TiO2-x300 W Xe lamp
ET: 80 °C
H2 production3667 μmol g−1 h−1TiO2-x:
262 μmol g−1 h−1
[100]
SAAg-g-CN300 W Xe lamp
ET: 55 °C
H2 production498 μmol g−1 h−125 °C:
248 μmol g−1 h−1
[101]
FeS2/Bi2S3300 W Xe lampH2 production16.8 mmol g−1 h−1FeS2:
0.52 mmol g−1 h−1
[102]
wood/CuS–MoS2AM 1.5 G
(100 mW/cm2)
H2 production85604 μmol g−1 h−1MoSx-TiO2:
11090 μmol g−1 h−1
[103]
Ag@SiO2@TiO2/AuUV LED
(365 nm)
(150 mW/cm2)
H2 production30.2 mmol g−1 h−1SiO2@TiO2:
0.41 mmol g−1 h−1
[104]
SnSe/ZnIn2S4300 W Xe lamp
(UV-vis-NIR)
H2 production5058 μmol g−1 h1ZnIn2S4:
1691 μmol g−1 h−1
[31]
Co0.85Se/Mn0.3Cd0.7S300 W Xe lamp
(λ > 420 nm)
ET: 25 °C
H2 production79.7 mmol g−1 h−15 °C:
46.7 mmol g−1 h−1
[28]
Ni2P/TiO2(B)300 W Xe lamp
ET: 90 °C
H2 production20.129 mmol g−1 h−150 °C:
6.752 mmol g−1 h−1
[105]
Au/SiO2/CdS/AgLEDs
(400–800 nm)
H2 production130 mmol g−1 h−1SiO2/CdS/Ag:
37.53 mmol g−1 h−1
[106]
WO3/CdS300 W Xe lamp
(UV-vis-NIR)
H2 production65.98 mmol g−1 h−110 °C:
20.82 mmol g−1 h−1
[32]
Zn NPshalogen lamp
(4.67 W/cm2)
H2 production200 μmol g−1 h−1Zn powder:
20 μmol g−1 h−1
[107]
PVDF-HFP/CdS/CNT280 W Xe lampH2 production451 μmol g−1 h−1PVDF-CTFE/CdS:
136 μmol g−1 h−1
[108]
PDA/DCNXe lamp
(λ > 420 nm)
MB degradationDE = 98%
(70 min)
DCN:
DE = 48%
[109]
Cu0.75Ag0.5S300 W Xe lampMB degradationDE = 93.8%
(30 min)
CuS/rGO:
DE = 80%
(140 min)
[110]
CuFe2O4@MIL-100(Fe, Cu)300 W Xe lamp
(λ > 400 nm)
MB degradationk = 0.075 min−1CuFe2O4:
k = 0.023 min−1
[111]
Prussian blue (PB) microcrystalssolar simulator
(100 mW/cm2)
MB degradationk = 0.0430 min−1PC:
k = 0.0231 min−1
[112]
C@TiO2300 W Xe lampRhB degradationk = 0.045 min−1TiO2 MNF:
k = 0.011 min−1
[113]
OPtCu-NCs808 nm
NIR laser
RhB degradationDE = 91.87%
(120 min)
[114]
CoFe2O4-BiOCl300 W Xe lampRhB degradationk = 1.16 min−1Vis:
k = 0.74 min−1
[115]
ZnSnO3UV light
(293–338 K)
RhB degradationDE = 98.1%
(80 min)
UV:
DE = 76.8%
[116]
ZnO6 W mercury arc lampRhB degradationDE ≈ 75%
(30 min)
PC:
DE ≈ 50%
[39]
Au/TiO2_PW300 W Xe lamp
(UV–vis light)
MO degradationDE = 74%
(30 min)
Au-TiO2:
DE = 10%
[117]
TiXn150 W Xe lampMO degradationDE = 90%
(120 min)
TiS3:
DE = 34%
(180 min)
[118]
Ag/TiO2300 W Hg lampMO degradationDE = 100%
(60 min)
-[37]
Fe3O4@SiO2-laccase500 W Xe lamp
(λ > 400 nm)
MG degradationDE = 99.6%
(60 min)
alizarin red:
DE = 79.3%
[119]
Pt−Cu/TiO2300 W Xe lamptoluene degradationDE = ∼100%
(120 min)
110 °C:
DE = ∼1%
[120]
WO3-x-R/GdCrO3Xe lamp
(400 mW/cm2)
toluene degradationk = 0.029 min−1SrTiO3/TiO2:
k = 0.0054 min−1
[121]
Pt/[TiN@TiO2]Xe lamp
(500 mW/cm2)
toluene degradationDE = 100%
(24 min)
Pt/TiO2:
DE = ∼45%
[122]
SmMnO3/CuMnOx400 W Xe lamp
(λ > 420 nm)
toluene degradationDE = 100%
(275 °C)
SmMnO3:
DE = 100%
(300 °C)
[123]
LaMn1.3O3300 W Xe lamp
(94 mW/cm2)
toluene degradationDE = 98%
(120 min)
-[124]
Cu/TiO2-x/CoP300 W Xe lamp2,4-D degradationDE = 99.2%
(180 min)
TiO2-x/CoP:
DE = 75.8%
[125]
Ag/Bi2S3/MoS2500 W Xe lamp
(λ > 420 nm)
2,4-D degradationk = 0.02334 min−1Bi2S3/MoS2:
k = 0.00638 min−1
[126]
Cu2-xS/CdS/Bi2S3300 W Xe lamp
(λ > 420 nm)
2,4-D degradationk = 0.03193 min−1Cu2-xS/Bi2S3:
k = 0.00422 min−1
[127]
α-Fe2O3/Defective MoS2/AgXe lamp
(λ > 420 nm)
2,4-D degradationk = 0.043 min−1α-Fe2O3:
k = 0.0013 min−1
[128]
TiO2200 W
mercury lamp
2,4-D degradationk ≈ 0.007 min−1
(O3)
PC:
k ≈ 0.001 min−1
[129]
W18O49@ZnIn2S4/CC300 W Xe lamp
(λ > 420 nm)
BPA degradationDE = 95%
(150 min)
CC:
DE = 15%
[130]
Ag/NaCNN/NiFe-LDH500 W Xe lamp
(λ > 420 nm)
BPA degradationk = 0.0432 min−1NiFe-LDH:
k=0.00292 min−1
[51]
Cu2-xS/Fe-POMs/AgVO3300 W Xe lamp
(λ > 420 nm)
BPA degradationDE = 98.6%
(150 min)
AgVO3:
DE = 19.6%
[131]
Bi7O9I3/AgIXe lamp
(λ > 420 nm)
phenol degradationDE = 95.38%
(80 min)
Bi7O9I3:
DE = 30.1%
[132]
Bi12CoO20300 W Xe lamp
ET: 90 °C
phenol degradationk = 0.113 min−115 °C:
k = 0.009 min−1
[133]
AuAgPt-12 YSNSs500 W Xe lamp4-NP degradationk = 0.155 min−1AuAgPt-6 YSNSs:
k=0.023 min−1
[134]
CaCO3/CuSNIR laser
(2.5 W/cm2)
4-NP degradationDE = 98%
(15 min)
CuS NPs:
DE = 0%
[34]
SrTiO3/MnFe2O4200 W
low-pressure
Hg lamp
TC
degradation
DE = 100%
(20 min)
PC:
DE = 38.2%
(25 min)
[38]
ZnO-MnO2300 W Xe lampTC
degradation
k ≈ 0.23 min−1MnO2:
k ≈ 0.2 min−1
[135]
Ag/Ag3PO4/CeO2300 W Xe lampbenzene degradationDE = 90.18%
(180 min)
CeO2:
DE ≈ 70%
[136]
α-MnO2/GO300 W Xe lampformaldehyde degradationDE = 100%
(60 min)
-[137]
Co3O4/rGO500 W Xe lamp
(UV-vis-NIR)
fethanol degradationDE = 96%
(90 min)
UV-vis:
DE = 41%
[138]
ZnFe2O4MEDL
MW oven
(100 W)
TCH
degradation
DE = 91.6%
(250 s)
PC:
DE ≈ 20%
[139]
TiO2UV lamps4C2AP
degradation
DE = 93.23%
(30 min)
PC:
DE = 85.28%
[140]
ET: external heating temperature; PC: photocatalysis; DE: degradation efficiency; k: degradation rate constant; MEDL: microwave electrodeless discharge lamp; MW: microwave; MB: methylene blue; RhB: rhodamine B; MO: methyl orange; MG: malachite green; 2,4-D: 2,4-dichlorophenol; BPA: bisphenol A; 4-NP:4-nitrophenol; TC: tetracycline; TCH: tetracycline hydrochloride; 4C2AP: 4-chloro 2-aminophenol.
Table 2. Summary of mechanical-coupled photocatalysis in recent years.
Table 2. Summary of mechanical-coupled photocatalysis in recent years.
MaterialConditionApplicationActivityBlank ControlRef.
BaTiO3@C300 W Xe lamp
40 kHz US cleaner
RhB
degradation
k = 0.03585 min−1PZC:
k = 0.00085 min−1
[147]
Ag@Na0.5Bi0.5TiO3300 W Xe lamp
40 kHz US cleaner
RhB
degradation
k = 0.146 min−1Na0.5Bi0.5TiO3(PC):
k = 0.019 min−1
[148]
Bi2WO6/Black TiO2220 W Xe lamp
US cleaner
RhB
degradation
DE = 98.43%
(60 min)
PC:
DE = 54.23%
(60 min)
[149]
Bi2WO6/g-C3N4/ZnO300 W Xe lamp
40 kHz US cleaner
RhB
degradation
k = 0.231 min−1PC:
k = 0.097 min−1
[150]
g-C3N4/Ag/ZnO50 W LED
US cleaner
RhB
degradation
DE = 89%
(180 min)
PC:
DE = 70%
[151]
PAN/TiO2350 W Xe lamp
100 kHz EPR spectrometer
RhB
degradation
k = 0.036 min−1PC:
k = 0.015 min−1
[152]
PbTiO3/g-C3N4300 W Xe lamp (λ > 420 nm)
US vibration
RhB
degradation
k = 0.1357 min−1PC:
k = 0.1044 min−1
[153]
Ag2O/Bi4Ti3O12400 W
metal halide lamp
US vibration
RhB
degradation
k = 0.1557 min−1PC:
k = 0.0363 min−1
[154]
BaTiO3/SrTiO3LED UV lamp
(30 W, 365 nm)
40 kHz US cleaner
RhB
degradation
DE = 97.4%
(30 min)
SrTiO3:
DE = 44.3%
[155]
AgI/ZnO250 W Xe lamp (λ > 400 nm)
US vibration
RhB
degradation
k = 0.037 min−1ZnO:
k = 0.002 min−1
[156]
Bi2WO69 W LED
120 W US cleaner
RhB
degradation
k = 0.141 min−1PC:
k = 0.008 min−1
[157]
Ag/BaTiO3500 W Xe lamp
150 W US vibration
RhB
degradation
DE = 70%
(120 min)
PC:
DE = 25%
[158]
NaNbO3/CuBi2O450 W LED
35 W US vibration
RhB
degradation
DE = 75%
(90 min)
NaNbO3:
DE = 40%
[159]
BaTiO3/TiO2250 W Xe lamp
40 kHz US cleaner
RhB
degradation
k = 0.0967 min−1TiO2:
k = 0.0275 min−1
[160]
ZnO@PVDF300 W Xe lamp
magnetic stirrer
RhB
degradation
DE ≈ 95%
(100 min)
PC:
DE ≈ 55%
[161]
ZnO/ZnS300 W Xe lamp
180 W US vibration
MB
degradation
DE = 53.8%
(50 min)
PC:
DE = 19.1%
[162]
PMN-PT@SnO2250 W
metal halide lamp
45 kHz US cleaner
MB
degradation
DE = 97%
(120 min)
SnO2:
DE = 87%
[163]
Ag2MoO4300 W Xe lamp
40 kHz US vibration
MB
degradation
DE = 96.2%
(40 min)
PC:
DE = 82%
[164]
Ag3PO4/ZnO300 W Xe lamp
(λ > 420 nm)
40 kHz US cleaner
MB
degradation
DE = 98.16%
(30 min)
PC:
DE = 90.18%
[165]
ZnO/ZnS/MoS2300 W Xe lamp
magnetic stirrer
(1000 rpm)
MB
degradation
k = 0.0411 min−1PC:
k = 0.0089 min−1
[166]
BaTiO3/CuO200 W Xe lamp
US cleaner
MO
degradation
k = 0.05 min−1PZC:
k = 0.007 min−1
[167]
BiOBr/BaTiO3Xe lamp
(100 mW cm–2)
40 kHz US vibration
MO
degradation
k = 0.1123 min−1BaTiO3:
k = 0.001 min−1
[168]
ZnO NR/PVDF-HFPXe lamp
(180 mW/cm2)
magnetic stirrer
(1000 rpm)
MO
degradation
k = 0.0399 min−1PC(200 rpm):
k = 0.0101 min−1
[169]
ZnO/MoS2300 W Xe lamp
magnetic stirrer
(1000 rpm)
MO
degradation
DE = 92.7%
(50 min)
PC:
DE = 50.6%
[170]
TiO2@rGO-F/PVDF-HFP300 W Xe lamp
(UV light)
magnetic stirrer
MO degradationDE = 99%
(100 min)
-[171]
Bi2MoO6/BiOBr400 W
metal halide lamp
US cleaner
MV degradationk = 0.0284 min−1Bi2MoO6:
k = 0.0082 min−1
[172]
SrBi4Ti4O15300 W Xe lamp
40 kHz US cleaner
TC degradationk = 0.058 min−1PC:
k = 0.004 min−1
[173]
Ti32-oxo-cluster/BaTiO3/CuS300 W Xe lamp
120 W US vibration
TC degradationDE = 100%
(60 min)
PC:
DE = 55.67%
[174]
CsCdBO3300 W Xe lamp
(λ > 420 nm)
40 kHz US cleaner
TC degradationDE = 92%
(30 min)
-[175]
ZnO/CdS300 W Xe lamp
150 W US cleaner
BPA
degradation
k = 0.1557 min−1PC:
k = 0.0135 min−1
[176]
BiOI/ZnO300 W Xe lamp
40 kHz US cleaner
BPA
degradation
DE = 100%
(30 min)
PC:
DE = 25%
[177]
PZT/TiO2LED light
(15 mW/cm2)
magnetic stirrer
(800 rpm)
BPA
degradation
DE ≈ 90%
(40 min)
PC(200 rpm):
DE ≈ 70%
[178]
Au-BiOBr300 W Xe lamp
40 kHz US cleaner
CBZ
degradation
k = 0.091 min−1k = 0.00516 min−1[179]
BaTiO3/La2Ti2O7300 W Xe lamp
(λ > 420 nm)
40 kHz US cleaner
CIP
degradation
k = 0.0844 min−1BaTiO3:
k = 0.0469 min−1
[180]
ZnS/
Bi2S3-PVDF
300 W Xe lamp
(λ > 420 nm)
US cleaner
H2 production10.07 mmol g−1 h−1ZnS/Bi2S3:
1.77 mmol g−1 h−1
[181]
OH-modified SrTiO3300 W Xe lamp
40 kHz US cleaner
H2 production701.2 μmol g−1 h−1PC:
295.4 μmol g−1 h−1
[182]
PbTiO3/CdS300 W Xe lamp
US vibration
H2 production849 μmol g−1 h−1PC:
98.9 μmol g−1 h−1
[183]
g-C3N4300 W Xe lamp
(λ > 420 nm)
US cleaner
H2 production12.16 mmol g−1 h−1PZC:
8.35 mmol g−1 h−1
[184]
g-C3N4/LiNbO3/PVDF300 W Xe lamp
magnetic stirrer
H2 production136.02 μmol g−1 h−1PC:
87.71 μmol g−1 h−1
[185]
g-C3N4/PDI-g-C3N4300 W Xe lamp
40 kHz US cleaner
H2O2
production
625.54 μmol g−1 h−1PC:
149.85 μmol g−1 h−1
[186]
ZnS/In2S3/BaTiO3300 W Xe lamp
(λ > 400 nm)
40 kHz US horn
H2O2
production
228 μmol g−1 h−1 72 μmol g−1 h−1 [187]
US: ultrasonic; PC: photocatalysis; PZC: piezocatalysis; DE: degradation efficiency; k: degradation rate constant; MB: methylene blue; RhB: rhodamine B; MO: methyl orange; MV: methyl violet; TC: tetracycline; BPA: bisphenol A; CBZ: carbamazepine; CIP: ciprofloxacin.
Table 3. Summary of electromagnetism-coupled photocatalysis in recent years.
Table 3. Summary of electromagnetism-coupled photocatalysis in recent years.
MaterialConditionApplicationActivityBlank ControlRef.
ZnOUV light
magnet (600 mT)
MB
degradation
DE = 60%
(3 min)
PC:
DE = 24%
[203]
CoFe2O4UV light
(100 mW/cm2)
magnet (200 mT)
MB
degradation
DE = 80%
(60 min)
PC:
DE = 28%
[204]
n-TiO2/PS400 W UV light
power supply (3 V)
MB
degradation
DE = 38%
(30 min)
PC:
DE = 20%
[205]
rGO/TiO240 W UV lamp
Nd2Fe14B magnet
MO
degradation
DE = 91%
(120 min)
TiO2:
DE = 68%
[206]
TiO24 W mercury lamp
magnet (280 mT)
MO
degradation
DE ≈ 94%
(75 min)
PC:
DE ≈ 70%
[207]
α-Fe2O3/TiO2Diode Green Laser
magnet (400 mT)
RhB
degradation
DE ≈ 70%
(60 min)
PC:
DE ≈ 45%
[208]
Ti0.936O2300 W Xe lamp
electromagnet
(80 mT)
RhB
degradation
k ≈ 0.075 min−1PC:
k ≈ 0.05 min−1
[209]
BiFeO3500 W Xe lamp
(λ > 420 nm)
electrical poling
RhB
degradation
k = 0.035 min−1PC:
k = 0.016 min−1
[210]
α-Fe2O3/rGO300 W Xe lamp
magnet
CR
degradation
DE = 87%
(30 min)
PC:
DE = 60%
[211]
CoFe2O4/MoS2300 W Xe lamp
electromagnet
(150 mT)
CR
degradation
DE = 96.6%
(60 min)
PC (50 mT):
DE = 82%
[212]
BiOBr/BNQDs300 W Xe lamp
(λ > 420 nm)
magnet
TC
degradation
DE = 81%
(60 min)
BiOBr:
DE = 60%
[213]
hierarchical TiO2
microspheres
500 W Xe lamp
power supply (2 V)
TBT
degradation
k = 0.0488 min−1PC:
k = 0.0052 min−1
[214]
Mn3O4/γ-MnOOH300 W Xe lamp
magnet (60 mT)
NOR
degradation
DE = 98.8%
(160 min)
PC:
DE = 90.3%
[215]
α-Fe2O3/Zn1-xFexOXe lamp
(λ > 420 nm)
ten magnets
(20 mT)
RIB
degradation
k = 0.0125 min−1PC:
k = 0.0072 min−1
[216]
Au/Fe3O4/N-TiO270 W tungsten light
magnet (180 mT)
H2 production21230 μmol g−1 h−1PC:
7600 μmol g−1 h−1
[217]
CdS/MoS2/Mo300 W Xe lamp
(100 mW/cm2)
rotating magnet
H2 production1.97 mmol g−1 h−1PC:
1.04 mmol g−1 h−1
[218]
Au-CdS300 W Xe lamp
rotating magnet
H2 production105 μmol g−1 h−1PC:
223 μmol g−1 h−1
[219]
Pt/TiO23 W LED lamp
DC power (1 V)
H2 production3242.6 μmol g−1 h−1PC:
1102.8 μmol g−1 h−1
[220]
K0.5Na0.5NbO3300 W Xe lamp
corona poling
(690 kV/cm)
H2 production470 μmol g−1 h−1PC:
63 μmol g−1 h−1
[221]
Rutile TiO2 nanograsssunlight
power supply (2 V)
Cr ion removal143.8 mg/g-[222]
Bi2MoO6300 W Xe lamp
corona poling
(20 kV/cm)
CO2 reduction14.38 μmol g−1 h−1
(CO)
PC:
4.08 μmol g−1 h−1
[223]
BaTiO3300 W Xe lamp
magnet
nitrogen fixation1.93 mgL−1 h−1PC:
1.35 mgL−1 h−1
[224]
PC: photocatalysis; DE: degradation efficiency; k: degradation rate constant; MB: methylene blue; RhB: rhodamine B; MO: methyl orange; CR: Congo red; TC: tetracycline; TBT: tributyltin; NOR: norfloxacin; RIB: ribavirin.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mi, Y.; Fang, W.; Jiang, Y.; Yang, Y.; Liu, Y.; Shangguan, W. Recent Advancements in Photocatalysis Coupling by External Physical Fields. Catalysts 2022, 12, 1042. https://doi.org/10.3390/catal12091042

AMA Style

Mi Y, Fang W, Jiang Y, Yang Y, Liu Y, Shangguan W. Recent Advancements in Photocatalysis Coupling by External Physical Fields. Catalysts. 2022; 12(9):1042. https://doi.org/10.3390/catal12091042

Chicago/Turabian Style

Mi, Yan, Wenjian Fang, Yawei Jiang, Yang Yang, Yongsheng Liu, and Wenfeng Shangguan. 2022. "Recent Advancements in Photocatalysis Coupling by External Physical Fields" Catalysts 12, no. 9: 1042. https://doi.org/10.3390/catal12091042

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