Home Journals Progress in Chemistry
Progress in Chemistry

Abbreviation (ISO4): Prog Chem      Editor in chief: Jincai ZHAO

About  /  Aim & scope  /  Editorial board  /  Indexed  /  Contact  / 
12

In-situ Preparation Methods of Hydrogen Peroxide via Water Oxdation

  • Jingze Yu 1 ,
  • Tengfeng Xie , 2, *
Expand
  • 1 College of New Energy and Environment, Jilin University, Changchun 130021, China
  • 2 College of Chemistry, Jilin University, Changchun 130012, China
* Corresponding author e-mail:

Received date: 2023-06-19

  Revised date: 2023-11-11

  Online published: 2024-01-08

Supported by

National Natural Science Foundation of China(22172057)

Abstract

Hydrogen peroxide (H2O2) is an important chemical that may be used as a clean disinfectant. For scale application, H2O2 is produced primarily by the anthraquinone process. The necessary transportation and storage processes bring explosion risks, so it is urgent to develop in-situ preparation methods. Electrochemical and photocatalytic reduction of oxygen to product H2O2 have received wide attention, but these reactions are carried out at the gas-liquid-solid interface. This three-phase reaction requires complex equipment and sequentially limits large-scale production. Another equally important pathway for in-situ H2O2 production is the oxidation of water which needs only solid-liquid two-phase interface. This paper summarizes the common methods of oxidizing water to prepare H2O2, such as electrochemistry and photocatalysis, and focuses on the recent new methods of in-situ H2O2 preparation, including thermal catalysis, ultrasonic piezoelectricity, plasma and microdroplet method. These methods provide the references for in-situ H2O2 production and in particular its utilization in the field of disinfection.

Contents

1 Industrial process for the production of hydrogen peroxide

2 In-situ production of hydrogen peroxide via oxygen reduction reaction

3 In-situ production of hydrogen peroxide via water oxidation reaction

3.1 Electrochemical and photocatalytic hydrogen peroxide generation from water oxidation

3.2 Thermocatalytic hydrogen peroxide generation

3.3 Ultrasonic piezoelectrical hydrogen peroxide generation

3.4 Electrical discharge plasma hydrogen peroxide generation

3.5 Generation of hydrogen peroxide from aqueous microdroplets

4 Conclusion and outlook

Cite this article

Jingze Yu , Tengfeng Xie . In-situ Preparation Methods of Hydrogen Peroxide via Water Oxdation[J]. Progress in Chemistry, 2024 , 36(2) : 177 -186 . DOI: 10.7536/PC230613

1 Industrial production method of hydrogen peroxide

The standard reduction potential of hydrogen peroxide (H2O2) is 1.763 V (vs NHE). It has strong oxidizing properties. It can oxidize and destroy cell components, such as DNA, lipids and proteins of pathogen cell wall. It can kill viruses, spores, fungi and bacteria. It is often used as a disinfectant[1,2][3]. In the process of disinfection, H2O2 is decomposed into water and oxygen, which is harmless to human body, so it is often used for sterilization and disinfection of medical environment, respiratory tract and oral cavity, water and food.
Atomized H2O2 or H2O2steam are commonly used for disinfection of wards, ambulances, engine rooms and other environments. The nebulization system pressurizes the H2O2 solution with a concentration of 5% to 6% and sprays it through the nozzle to produce H2O2 aerosol with a diameter of 0.5 to 10 μm. The commonly used dose is 6 mL/m3, and the disinfection time is 2 H[4]. A variety of H2O2 gasification spray sterilizers have been sold in the market[1,5]. The H2O2 vapor system consists of a vapor generator and a H2O2 concentration monitor. The humidity of the space is reduced by a dehumidifier before disinfection, and then the H2O2 vapor is produced by evaporating a H2O2 solution with a concentration of more than 30% at a high temperature. The H2O2 vapor is circulated in the closed room by a high-speed air flow[4]. H2O2 spray can also be used for dental disinfection. Lazzarino's team recommends the use of “0.28 mL1.5%H2O2 nasal spray twice a day and gargling with 3%H2O2 solution for 1 min twice a day to reduce the risk of respiratory infectious diseases[6]. One of the important goals of food disinfection is to decompose aflatoxins. The use of H2O2 in food disinfection generally requires a combination of high temperature, ultraviolet radiation, alkali treatment and other methods. For example, heating 6% H2O2 aqueous solution to 80 ℃ and soaking defatted peanut powder for 30 min can remove 97% of aflatoxins[7]. When the concentration of H2O2 added to surface water reached 0. 3%, the phage could be killed by 6. 5 logs, while chlorine disinfection could only be completed by 3. 0 logs[8].
2-ethylanthraquinone process is the main method for large-scale centralized production of H2O2, accounting for 95% and 99% of global and domestic production of H2O2[9]. The anthraquinone process uses hydrogen and oxygen as raw materials, 2-ethylanthraquinone and palladium metal as catalysts, and heavy aromatics (1,2,4-trimethylbenzene) as solvent to produce H2O2. 2-Ethylanthraquinone is dissolved in heavy aromatics at a certain temperature and pressure, and then reacts with hydrogen under the action of a palladium catalyst to generate 2-ethylhydroanthraquinone, wherein the solubility of the 2-ethylanthraquinone is not high, and the solubility of the 2-ethylhydroanthraquinone increases after the 2-ethylanthraquinone is converted. 2-Ethylhydroanthraquinone reacts with oxygen at a certain temperature and pressure to produce 2-ethylanthraquinone and H2O2[10]. The aqueous solution of the H2O2 is extracted from the mixed solution by extraction, and finally the qualified aqueous solution of the H2O2 is obtained by heavy aromatics purification. The concentration of H2O2 solution directly produced by anthraquinone method is usually 27.5%, and higher concentrations (such as 35%, 50%, 70%) of H2O2 solution can be obtained by distillation and concentration[11].
H2O2 decomposes to produce water and oxygen and generates 98 kJ/mol of heat[12]. At room temperature, H2O2 decomposes slowly, about 1% per year[12]. When the temperature rises to 100 ℃, the decomposition rate increases sharply, the volume expands rapidly and the heat is released centrally, which is easy to cause explosion[13]. H2O2 with a concentration of more than 8% need to be packaged and transported in special containers. In order to save transportation costs, H2O2 are often concentrated to more than 50%, so there is a risk of explosion during transportation and storage. However, the concentration required in the actual disinfection process is only 3% to 7.5%, and high concentration H2O2 needs to be diluted. Obviously, in-situ production of H2O2 with concentration meeting disinfection requirements can not only omit the tedious steps of concentration, transportation, storage and dilution, but also avoid the risk of explosion.
Hydrogen and oxygen direct synthesis methods are also used to prepare H2O2 on a large scale. These methods not only require the use of dangerous or expensive raw materials such as H2 and metal catalysts, but also have problems in storage and transportation[14].

2 In situ H2O2 process using oxygen as raw material

Both electrochemistry and photocatalysis can produce H2O2 in situ through the oxygen reduction route, the principle of which is that dissolved oxygen in water is reduced to H2O2 by electrons on the surface of the cathode or photocatalytic material, and the reaction formula is as follows :O2+2H++2e‾→H2O2[15,16]. The research goal of 2-electron oxygen reduction (2e ORR) to produce H2O2 is to continuously increase the concentration and volume of H2O2 aqueous solution to meet disinfection requirements. The two important ways are to regulate the adsorption energy of OOH groups on the surface of electrodes or photocatalytic materials and to inhibit the hydrogen production reaction of water decomposition.
There are two ways for oxygen to adsorb on the surface of the electrode. Oxygen connected by Bridge tends to get four electrons to produce water; The Pauling linked oxygen gets one electron to produce OOH group, the strongly adsorbed OOH group gets three electrons to produce water, and the moderately adsorbed OOH group gets one electron to produce H2O2. Therefore, the adsorption energy of OOH groups on the electrode surface is the key to determine the selectivity of H2O2[17]. The key to achieve this goal is to develop cathode materials with high 2e ORR selectivity, such as doped carbon electrodes[18]. In addition, the saturated dissolved oxygen concentration in water is only 8. 25 mg/L in air atmosphere at room temperature and normal pressure, which can only be converted into 000876 solution with a concentration of 8. 76 mg/L in theory, and the weight ratio is 0. 000876%, which is more than three orders of magnitude lower than the required concentration for disinfection. Therefore, hollow electrode aeration, pure oxygen aeration, pressurized oxygen supply, gas diffusion electrode and other ways are used to increase oxygen supply[19][20][21][22]. At present, the concentration of H2O2 in alkaline solution reported in the literature has exceeded 5%, which meets the requirements of practical disinfection[23].
In order to produce a H2O2 solution with a concentration and volume meeting the actual disinfection requirements in a neutral solution, the key technical problem to be solved is to develop a cathode material with high current density tolerance and high H2O2 selectivity. The electrochemical oxygen reduction process requires oxygen, aqueous solution and electrodes, that is, gas-liquid-solid three-phase interface. All aeration methods can not avoid the mass transfer problem caused by gas diffusion, which is also an inherent factor limiting the production of H2O2. Although the gas diffusion electrode does not need to be aerated in water, the hydrophobic layer of the gas diffusion electrode gradually becomes hydrophilic after being immersed in water for a long time, and finally the flooding phenomenon occurs and the gas diffusionelectrode fails, so how to prolong the service life of the gas diffusionelectrode becomes a difficult problem that must be solved to meet the application requirements[24][22].
The pH condition is an important factor for the preparation of H2O2. Some literatures have found that the yield of H2O2 under acidic condition is the highest, followed by neutral condition, and alkaline condition is the lowest[25]; However, it has also been reported that alkalinity is the best, followed by neutrality, and acidity is the worst[26]. The inconsistency of the effect of pH conditions on the performance of H2O2 production may be related to the properties of electrode materials, and revealing the general law plays an important role in further understanding the mechanism of H2O2 production by electrochemical oxygen reduction and further improving the yield.
Inhibition of side reactions must be considered to improve H2O2 production. There are several competitive reactions in the process of electrochemical oxygen reduction to produce H2O2, including cathodic water decomposition to produce hydrogen, 4-electron oxygen reduction to produce water, and 1-electron H2O2 reduction to produce · OH. Chemical oxygen reduction occurs at the cathode, and one of the products of the water splitting side reaction is OH-, which leads to a local pH rise, and alkaline conditions easily trigger the decomposition of H2O2 to produce O2 and H2O[27]. In addition, the H2O2 generated at the cathode diffuses to the anode and can be rapidly oxidized to produce O2 and H+. Therefore, to further increase the concentration of H2O2, it is necessary to effectively inhibit side reactions and avoid the decomposition of H2O2. The common idea is to increase the hydrogen overpotential of cathodic decomposition, as mentioned in the previous paragraph, to inhibit the 4E ORR and the reductive decomposition of H2O2 by heteroatom doping, to use a diaphragm to prevent the diffusion of the produced H2O2 to the anode, and to use a continuous flow form to avoid the accumulation of OH-.
Electrochemical oxygen reduction to produce H2O2 is currently the focus of attention in this field. There are many reports in the literature. For more comprehensive content, please refer to the published review articles on the principle, device and process[28][29][30].

3 Method for producing H2O2 in situ using water as raw material

3.1 Electrochemical and Photocatalytic Methods for H2O2 Production from Water Oxidation

Another way to electrochemically produce H2O2 is the anodic oxidation of water, which proceeds as follows :2H2O→H2O2+2H++2e‾. Water oxidation is a solid-liquid two-phase interfacial reaction, which only uses water as raw material without oxygen supply, and has mass transfer advantages over oxygen reduction reaction requiring gas-liquid-solid three-phase interface. The key to increase the concentration of H2O2 produced by water oxidation is to inhibit the competitive reaction. The competitive reaction of 2-electron oxidation of water (2e-WOR) to produce H2O2 is mainly 4-electron oxidation of water (4e-WOR) to produce oxygen. In order to improve the selectivity of H2O2, the 4-electron oxygen production reaction must be inhibited, and the main method is to use semiconductor electrodes with high oxygen production overpotential.Such as BiVO4, TiO2, WO3, SnO2, CaSnO3, ZnO, etc.,Commercial carbon felt, carbon cloth, carbon paper and boron-doped diamond as anodes also have high oxygen overpotential, which are often used to oxidize aquatic H2O2[1,31][32][33].
Another way to suppress the 4-electron oxidation reaction is to manipulate the adsorption energy of the intermediate. When the Gibbs adsorption free energy value (ΔG*OH) of * OH is between 1.6 and 2.4 eV, H2O2 is the dominant product, and ideally, 1.76 eV is the optimal thermodynamic value[34]. Strong adsorption of materials with ΔG*OH values below 1.6 eV leads to O2 precipitation, and weak adsorption of materials with ΔG*OH values above 2.4 eV leads to · OH production[34]. Therefore, reasonable regulation of OH adsorption is a feasible way to achieve efficient H2O2 production. At a sufficiently positive potential, the oxygen bubbles generated by 4 e water oxidation move upward from the electrode and then collapse and disappear. If these oxygen bubbles are confined near the active site, the accumulated O2 molecules can further interact with the OH on the surface of the catalyst, thus adjusting the binding strength in favor of 2-electron oxidation[34]. By coating the catalyst surface with a hydrophobic polymer to trap the in situ generated O2 gas close to the active site, Wang's group observed that once the generated O2 gas was confined to the surface, its H2O2 selectivity increased significantly[35]. Sun's group controlled the adsorption strength by doping a specific number of strongly OH-adsorbed ruthenium atoms into the lattice of weakly OH-adsorbed TiO2, and accumulated 20 mL of 422 mg/L H2O2 in 10 min under the optimal conditions[36].
At present, due to poor conductivity (metal oxide) or low selectivity (carbon material), the concentration of H2O2 aqueous solution produced by the reported electrode is only several hundred micromoles per liter, which can reach 24 ~ 33 mmol/L under strong alkaline conditions with high concentration of bicarbonate electrolyte, equivalent to 0. 1% of the mass concentration, which is still a certain gap from the actual disinfection requirement of 3%[32,37][38]. In order to increase the concentration, some reports adopt the idea of cathodic oxygen reduction-anodic water oxidation, and the two poles cooperate to produce H2O2. The reactor structure and H2O2 generation path are shown in Figure 1. In addition, the long-term stability of the electrode is also a problem to be solved in practical applications.
图1 电化学产H2O2反应器和反应路径示意图

Fig.1 Schematic diagram of electrochemical reactor and reaction route for H2O2 production

Similar to the electrochemical process, photocatalysis or photoelectrocatalysis can also generate H2O2 through the pathway of water oxidation[39]. Illuminated semiconductor photocatalytic materials, such as TiO2, g-C3N4, BiVO4, etc., can generate photogenerated electrons and holes. Photogenerated electrons are similar to the electrons provided by the cathode in the electrochemical process, which can reduce dissolved oxygen in water to produce H2O2. Holes are similar to the positive charges of the anode, which can oxidize aquatic H2O2. In the photocatalytic process, oxygen reduction and water oxidation, two ways of producing H2O2, can be carried out simultaneously[40]. The structure of the reactor for photoelectrocatalytic H2O2 production and the reaction path of the two-pole synergistic H2O2 production are shown in Fig. 2.
图2 光电化学产H2O2反应器和反应路径示意图

Fig.2 Schematic diagram of photo- electrochemical reactor and reaction route for H2O2 production

Obviously, photocatalysis saves electrical energy relative to electrochemical processes, but requires illumination. In addition, the electrochemical process requires aqueous solution to conduct electricity, that is, there must be electrolyte in water, while photocatalysis can produce H2O2 aqueous solution without other impurities using pure water as raw material[41]. The focus in the field of photocatalytic production of H2O2 is mainly on the development of new materials, but the cumulative concentration of H2O2 can rarely exceed 0.1%[42,43]. The main goal of photocatalytic methods is to further improve the concentration and volume of H2O2 produced, and the current exploration approach is still to improve the light absorption, promote the separation of photogenerated charges and enhance the efficiency of surface reaction, which is still emphasized by traditional photocatalysis[44].
In addition, there are also some methods for in-situ preparation of H2O2 with water as raw material, such as thermal catalysis method, ultrasonic piezoelectric catalysis method, plasma pyrolysis method, bioelectrochemical method, microdroplet method and so on. Although these methods have not been reported much, they provide new ideas or have application potential, and they are very promising to obtain H2O2 aqueous solution that meets the requirements of in-situ disinfection. This paper summarizes the latest reports in recent years and introduces these methods in turn.

3.2 Thermally catalyzed H2O2 production

The direct synthesis of H2 and O2 to produce H2O2 is a well-known thermocatalytic reaction[31,45]. Because of the explosion risk of mixing H2 and O2, it is necessary to find safe raw materials. Wate is an ideal hydrogen source, and that direct oxidation of (H2O+0.5O2=H2O2ΔG=116.7 kJ/mol) with O2 is a thermodynamically uphill reaction, thus requiring an energy input[46]. A common form of energy supply is thermal coupling, in which the heat released in an exothermic reaction drives an endothermic reaction, for example, the heat released in the oxidation of carbon monoxide to carbon dioxide provides heat for the gold-catalyzed reaction of water and oxygen to H2O2[47]. The product of liquid phase thermal coupling is a mixture of H2O2 and organic matter, which must be separated from the liquid phase. The high cost of the separation process is the disadvantage of this method. In addition, most of the energy of this process is used to heat the bulk of the diluent rather than to drive the desired reaction, which is energy inefficient for dilute solutions. Kung et al. Proposed to conduct the thermal coupling reaction in the vapor phase to produce H2O2, which can increase the reactant concentration and greatly improve the energy efficiency[48]. Their study showed that reactions such as epoxidation of alkenes and selective oxidation of alkanes to alcohols could not provide sufficient thermodynamic driving force, while the heat released from the oxidation of alcohols to aldehydes and acids could drive the oxidative aquatic H2O2 reaction. This approach requires consideration of the spontaneous decomposition of the resulting H2O2 at high temperatures.
Another form of thermal catalysis is that thermoelectric materials use the separation of positive and negative charges caused by temperature difference to convert heat energy into electrical energy to produce H2O2. Thermoelectric voltage is the key factor affecting the concentration and volume of H2O2 produced by thermoelectric materials. Bismuth telluride (Bi2Te3) has excellent physical properties and unique electronic structure at room temperature, which can produce higher thermoelectric voltage than other common thermoelectric materials such as antimony telluride (Sb2Te3) and lead telluride (PbTe). Lin et al. Prepared Bi2Te3@CFF materials by coating bismuth telluride nanoplates on carbon fiber fabrics[49]. Whether the temperature difference is positive or negative, the Bi2Te3@CFF shows significant antibacterial activity and good durability. Control experiments confirmed the generation of superoxide radicals during the thermocatalytic reaction, which favors the generation of H2O2, and also demonstrated that the disinfection ability comes from the in situ generated H2O2. A two-dimensional structure with a high surface area is responsible for that superior thermal catalytic performance of bismuth telluride nanoplate (NPs). The antibacterial filter based on Bi2Te3NPs is installed in the air conditioner, and the thermocatalytic reaction is triggered by the temperature difference between the inside of the air conditioner and the outside environment. The experimental results show that 5 mg bismuth telluride NPs can produce 30μM H2O2 within 20 min when the temperature difference is 30 K. The temperature difference was generated by a blower, and the sterilization rate of the 1×1 cm Bi2Te3@CFF against 1 mL of Escherichia coli with a concentration of 2×106CFU/mL reached 72% within 20 min, and the operation was stable for at least 30 days[49].

3.3 Production of H2O2 by Piezoelectric Catalysis

Piezoelectric catalytic effect is a concept proposed in 2010[50]. Under the action of external force, piezoelectric materials will produce a polarizing electric field, which induces the separation of electron-hole pairs and shows catalytic ability. In recent years, there have been more and more reports on the production of H2O2 by piezoelectric catalysis. For example, by vibrating piezoelectric microfibers, water can be split piezocatalytically with an energy conversion efficiency of 18%[50]. Electron-hole pairs induced by piezoelectricity can react with O2 or water respectively, both of which can produce H2O2[51].
Perovskite is the main piezoelectric material. In order to enhance piezoelectric polarization, manufacturing defects, doping metal elements and constructing composite materials are effective ways. Perovskites are a class of oxides with the general molecular formula ABO3. A and B are rare earth (alkaline earth) elements and transition metal elements, respectively, and a and B can be replaced by other metals with similar radii without changing the perovskite crystal structure. In general, the net charge of each unit cell in perovskite crystal is zero, but when the titanium ions in the unit cell are slightly off the center, the electric polarity will be generated, and the unit cell will be converted into an electric dipole. When mechanical stress acts on the crystal, the position of titanium ions changes, which changes the polarization of the crystal and produces piezoelectric effect.
The most widely studied perovskite in the field of H2O2 production is lead zirconate titanate, with formula Pb(Tix, Zr1-x)O3, formed by solid solution of PbTiO3 and PbZrO3. The polarization degree of the lead zirconate titanate is adjusted by changing the external force, so that the intensity of a local electric field can be adjusted, and the adsorption capability of the surface of the lead zirconate titanate to OOH groups can be adjusted, and the aim of improving the yield of H2O2 is achieved[52].
To avoid the use of toxic lead, researchers have explored the production and H2O2 of perovskite piezoelectric catalytic materials such as barium titanate. In order to improve the piezoelectric properties, ternary perovskite barium zirconate titanate has also been developed. Due to the coexistence of three phases (orthorhombic, tetragonal and rhombic), the polarization energy barrier is reduced, resulting in a significant enhancement of piezoelectricity. Combined with the suitable concentration of oxygen vacancies, the H2O2 selectivity of barium zirconate titanate reaches 80% under ultrasonic irradiation. The research of Li's group further reveals that piezoelectric polarization in ternary perovskites promotes the generation, separation and transport of electron-hole pairs, thus promoting the catalytic reaction[53].
Piezoelectric catalytic performance depends on the intensity of piezoelectric polarization. Fabricating defects, doping metal elements and constructing composite materials are effective methods to enhance piezoelectric polarization. Huang's group fabricated ultrathin Bi4Ti3O12 nanosheets with a thickness of only 4.2 nm, which is only equivalent to 1 – 2 Bi4Ti3O12 unit cells, and has abundant surface oxygen defects (OVs) due to a large number of exposed surfaces[54]. Compared with ordinary Bi4Ti3O12, this ultrathin Bi4Ti3O12 nanosheet rich in OVs is more sensitive to external force, which can generate a stronger piezoelectric polarization field, promote the separation and migration of piezoelectric free charges, prolong the lifetime of piezoelectric free charges, and reduce the adsorption energy of O2 molecules to promote their activation into superoxide radicals. Benefiting from these advantages, the ultrathin Bi4Ti3O12 nanosheets with the optimal concentration of OVs showed excellent piezoelectric catalytic H2O2 generation performance, and 0. 05 G of catalyst added to 100 mL of ethanol – water solution could generate 161 μmol of H2O2 in 2 H under ultrasound, which was converted to 0. 0054% by weight, significantly higher than 86 μmol (0. 0029%) of ordinary Bi4Ti3O12 nanosheets with a thickness of 25. 8 nm[54].
In addition to construction defects, doping piezoelectric materials with metal elements can improve polarization. For example, doping Nb5+ into the lattice of BaTiO3 can introduce charge compensation, increase the carrier concentration and charge migration rate of BaTiO3, and thus obtain higher polarization electric field intensity under the same mechanical force. Coupling piezoelectric effect with photocatalysis can effectively improve the ability to produce H2O2, but perovskite materials generally have the photocatalytic ability of ultraviolet light response, which can not utilize the largest proportion of visible light in the solar spectrum, while loaded carbon quantum dots (CDs) can usually help perovskite to obtain visible light response ability. Zhai's group used niobium-doped barium titanate (BaTiO3:Nb) to enhance piezoelectric polarization, and then loaded CDs as visible light sensitizer and electron acceptor to increase the concentration of H2O2[55]. Under co-irradiation of visible light and ultrasound, 15 mg of CDs modified Nb-doped BaTiO3 catalyst was added into 30 mL of ethanol-water mixed solution to produce H2O2 at a rate of 1360μmol/(g<sub> catalyst </sub > · H) (0. 0023%), which was 27 times higher than that of BaTiO3 48μmol/(g<sub> catalyst </sub > · H)[55].
The third way to strengthen the polarization electric field is to construct composite materials. A composite constructed with spherical zinc sulfide and nanopiezoelectric barium titanate modified multilayer In2S3 nanosheets (ZnS/In2S3/BS3/BaTiO3) produced about 378 µM of H2O2(0.0013%) in 100 min under co-irradiation of visible light and ultrasound (piezo-photocatalysis), and the concentration of H2O2 produced by ZnS/In2S3/BaTiO3 could reach 1160 µM (0.0039%) in 600 min of continuous reaction. The enhanced yield of H2O2 under piezo-photocatalysis is attributed to the enhanced rate of the surface reaction due to the piezoelectric effect induced polarization electric field which promotes the separation of photogenerated charges[56].
Curie temperature and piezoelectric coefficient are important parameters reflecting the properties of piezoelectric materials. Once the service temperature exceeds the Curie temperature, the piezoelectric materials will lose their piezoelectric properties; The piezoelectric coefficient reflects the ability of a piezoelectric material to convert mechanical energy into electrical energy or electrical energy into mechanical energy. The higher the piezoelectric coefficient, the higher the energy conversion efficiency. The Curie temperature of perovskite RbBiNb2O7 exceeds 1000 ℃, which is higher than that of most piezoelectric materials and can operate in a wider temperature range. The piezoelectric coefficient of organic polytetrafluoroethylene (PTFE) is as high as 600×10-12C\N, and it is easy to obtain permanent polarization by ultrasonic irradiation. Li's group loaded RbBiNb2O7 on PTFE and obtained a synergistic effect[57]. Firstly, the hydrophobicity of PTFE is beneficial to the adsorption of oxygen, and the piezoelectric polarization decomposes water to produce H2O2 and oxygen, which can produce H2O2 by oxygen reduction because oxygen is bound on the surface of the material. Secondly, the built-in electric field formed by PTFE and RbBiNb2O7 promotes the separation and migration of polarization charges, and improves the energy efficiency of H2O2. The addition of 10 mg RbBiNb2O7\PTFE to 1 mL of ethanol-water solution produces H2O2 at a rate of 219.23μmol/(g<sub> catalyst </sub > · H), which is converted to 0.0074% by weight. This value is 12 and 3 times that of RbBiNb2O7 alone and PTFE alone, respectively[57]. In addition to perovskite, some metals, carbon materials, organics and so on can also be used as piezoelectric materials to produce H2O2.
Metal Pt also has the ability of piezoelectric polarization to produce H2O2. With the help of ionic liquid, a single atom of Pt was anchored on the surface of silica (Pt1/SiO2), and the resultant catalyst could generate 3027.1µmol/(L·h)H2O2( (0.01% by weight) under sonocatalysis. This was the highest concentration of sonocatalytic H2O2 production reported in the literature at that time. The high yield of H2O2 is attributed to the efficient bidirectional catalysis of Pt1. :Pt1 not only has high water adsorption energy and low water activation energy, but also can selectively activate O2, so it can simultaneously produce H2O2 through two pathways of oxygen reduction and water oxidation[58].
The metal-free graphite-phase carbon nitride (g-C3N4) is a layered two-dimensional structure, and the strain gradient generated between the layers leads to an obvious piezoelectric response when an external force is applied to the layered material[50]. The piezoelectric properties of two-dimensional g-C3N4 are caused by its nanoscale ferroelectric properties and the nanoscale triangular noncentrosymmetric holes in its lattice structure. In addition, g-C3N4 has abundant pyridinic nitrogen, which is the active site for adsorption and activation of oxygen[50]. Therefore, g-C3N4 is an ideal material for the production of H2O2 by piezocatalytic oxygen reduction. The above analysis was supported by experiments, and the production rate of H2O2 under ultrasonic irradiation reached 34 μmol/H when 50 mg of catalyst was put into 100 mL of pure water, which was converted to 0.0011% by weight[50]. Although this concentration is still far from the requirement of disinfection, this work opens up a new direction for the production of H2O2 from pure water and oxygen without the use of metal catalysts. The charge transfer and reaction path of the decomposition of aquatic H2O2 by piezoelectric effect under ultrasonic and g-C3N4 is shown in Figure 3.
图3 超声作用下g-C3N4产H2O2的示意图

Fig.3 Schematic diagram of ultrasonic-driven generation of H2O2 over g-C3N4

Organic polymers can also produce H2O2 as metal-free piezoelectric materials. Zhang's group has developed a new flexible polymer film composed of polyvinylidene fluoride-hexafluoropropylene (PVDF-HFP) matrix and silica nanoparticle filler[59]. PVDF-HFP has weak piezoelectric effect, while SiO2 has no piezoelectric effect, and the H2O2 of PVDF-HFP or silica alone under ultrasonic irradiation is less than 10 μmol/ (L · H). After forming the composite film, silica can not only enhance the polarization electric field of PVDF-HFP, but also adsorb the reaction intermediate, which can produce 246 μmol/ (L · H) of 00083 (0.00083% by weight) under ultrasonic irradiation. Computer simulation and experimental results show that the H2O2 generation ability originates from the enhancement effect of the piezoelectric polarization electric field on the gas-liquid-solid interface reaction, the hydrogen bond formed by the oxygen atom of silica and the hydrogen atom of PVDF enhances the piezoelectric response, and the adsorption energy of the surface hydroxyl group of silica to the oxygen reduction intermediate -OOH is moderate, which is beneficial to the selective generation of H2O2. The flexible film has the advantages of simple preparation method, environmental friendliness, long-term storage stability, easy recovery after use, and practical application potential[59].

3.4 Plasma method for producing H2O2.

Plasma is a globally electrically neutral assembly of electrons, ions, excited particles, and neutrals. Plasma contains a large number of highly reactive particles, which can trigger complex chemical reactions, so it is widely used in the fields of environment, energy and materials. There are three possible mechanisms for plasma to produce H2O2: positive and negative charges in plasma are transferred into water to produce H2O2 by oxidizing water or reducing oxygen, respectively; The gas-phase hydroxyl radical are combined to generate a H2O2 which is then dissolve into a liquid phase; H2O+ ions in water are dehydrogenated and oxidized to hydroxyl radicals, which combine to produce H2O2[60].
Plasma can be generated by glow, arc, corona and dielectric barrier discharge. Locke's group has developed a gliding arc plasma reactor based on the principle of arc discharge. The two curved electrodes of the reactor form a gradual gap, which gradually expands from 2 mm to 20 mm[61]. They apply a high voltage between the two stages to break down the air to form a plasma, and then spray small water droplets into the plasma between the electrodes through a nozzle. The small gap of millimeter level ensures that the water mist is fully contacted with the plasma, which produces H2O2 based on the principle mentioned in the previous paragraph. They used a 250 Hz pulse power supply to avoid the temperature rise caused by continuous power supply, prevent the evaporation and gasification of water droplets and inhibit the thermal decomposition of H2O2. The final achievement of a 1 mL/min influent flow rate yielded a H2O2( weight percent of 0.034%) of 10 mM. The concentration of H2O2 in the liquid phase product decreased with the increase of influent flow rate, and the concentration of H2O2 produced was 0. 9 mM when the influent flow rate was 20 mL/min. At this time, the energy produced per kWh was about 80 gH2O2[61]. Ranieri et al. Showed that compressed air was used to carry small droplets with a diameter of about 0. 3 μm into the microsecond pulse dielectric barrier discharge plasma device, which could generate H2O2 in the droplets and evenly cover the surface of agricultural products to complete sterilization and preservation[62]. Using a plasma discharge cross-section of 80 cm2, it took only 2 min to complete the disinfection of a 100 L space at a power density of 1 W/cm2 and a compressed air flow rate of 50 L/min[62].
Discharge time, discharge power, liquid phase conductance and other factors can affect the yield of H2O2. Takeuchi et al. Systematically studied the process mechanism of H2O2 generation by plasma discharge, and found that the H2O2 concentration in the liquid phase increased linearly with the plasma treatment time and discharge current, and was independent of the discharge gap distance[63]. The H2O2 is mainly generated at the gas-liquid interface, and the H2O2 generated in the gas phase and the H2O2 generated in the liquid film near the interface migrate into the solution through diffusion. Under the optimal conditions, about 22 mg/L of H2O2 was produced in 25 mL of pure water in 3 min, and the energy efficiency was 3.9 g H2O2/kWh[63]. The generation of H2O2 depends on the plasma volume and the interface area between plasma and liquid phase. In general, as the power increases, the plasma volume expands and the interface with the liquid phase increases. This provides more opportunities for water molecules to collide with electrons, producing more • OH and H2O2. Liquid conductivity is also a key factor affecting the H2O2 produced by plasma. Vor Voráč's group obtained aqueous solutions with conductivities of 0.01 ~ 36 mS/cm by adding KCl to deionized water[64]. For the microsecond pulse plasma, the conductivity increases from 0.01 mS/cm to 0.3 mS/cm, resulting in a decrease of 62% in the H2O2 generation rate from 0.12 μmol/s to 0.045 μmol/s. For nanosecond pulsed plasma in argon plasma, the generation rates of H2O2 corresponding to liquid films with conductivity of 0.01 mS/cm and 36 mS/cm are 0.16 μmol/s and 0.14 μmol/s, respectively, which are reduced by about 13%. The results show that the increase of liquid film conductivity is not conducive to the generation of H2O2, and the lower the pulse frequency is, the stronger the adverse effect is[64]. These works confirm that plasma discharge can also produce H2O2 in conductive solution, thus indicating that this method can be used for seawater or other electrical water sterilization.
In the background of plasma, the location of H2O2 is divided into gas phase and liquid phase. In order to study the contribution of different positions to the generation of H2O2, Li's group compared the concentration of H2O2 generated by bubble pulse discharge on the water surface and underwater. Regardless of the discharge mode, the yield and production rate of H2O2 increased with the increase of voltage, and the rate of H2O2 generated by bubble pulse discharge in 200 mL pure water was 0. 0196 mg/min, which was significantly higher than 0. 012 mg/min in water surface pulse discharge[65]. Considering that the bubbling process forms more gas-liquid interfaces within the liquid film, this result indicates that H2O2 is more easily formed in the gas phase[65]. Tachibana et al. Investigated the H2O2 production per unit energy consumption of a similar bubbling discharge system and a water surface discharge system. After 10 min of discharge, the bubbling system produced 270 mg/L of H2O2 in 12 mL of water and 1 gH2O2 per kWh of electricity, while the water surface discharge system produced about 93 mg/L of H2O2 in 15 mL of water and 0.1 gH2O2 per kWh of electricity[66]. This result further demonstrates that the plasma method is more economical to produce H2O2 in the gas phase[66]. Plasma requires a millimeter-scale discharge gap, and how to generate it on a large scale has been a research hotspot of this technology. Vlachos group has developed a new type of plasma reactor for distributed production of H2O2, which is based on the principle of dielectric barrier discharge and is characterized by filling a high-density gas-liquid interface[67]. It can be operated at 4 kV voltage and 0.7 W power to produce 2.2 mM H2O2, and the concentration of 33 mM H2O2 can be further obtained by increasing the power, which makes the technology easy to expand through the series-parallel connection of unit reactors, and is expected to achieve large-scale production of H2O2[67].

3.5 H2O2 production by microdroplet method

In humid environment, the surface of fine particles in soil and atmosphere will spontaneously produce H2O2. Because the micro-droplet with fine particles as the core is a necessary condition for the generation of H2O2, this process can be called micro-droplet process[68]. The microdroplet process produces H2O2 with only water and fine particles, without the application of voltage, catalyst, or added chemicals. It is generally believed that the H2O2 produced in the microdroplet process is produced by the recombination of two hydroxyl radicals. When micro-droplets and fine particles come into contact, water molecules collide with atoms or molecules on the solid surface, and electron transfer occurs at the solid-liquid interface. The hydroxide anion loses an electron to produce a hydroxyl radical, and the two radicals recombine to produce H2O2[68].
In order to fabricate a solid-liquid interface that can simulate the microdroplet process, Chen et al. Used polydimethylsiloxane to construct a closed straight channel (diameter 20 μm, length 100 μm) on glass, and the H2O2 could be detected by introducing deionized water[68]. In this process, the oxygen atoms required for the formation of H2O2 may come from the oxygen-containing groups on the surface of fine particles. To test this idea, silica was used as a substrate, and the substrate was activated by O2 plasma to increase the density of hydroxyl groups, which showed that the density of hydroxyl groups was positively correlated with the amount of H2O2 generated at the solid-liquid interface. Isotope experiments have confirmed that the oxygen atoms in the H2O2 come from the hydroxyl groups of the substrate, and it is difficult to produce H2O2 by using microdroplets made of water containing hydroxyl radical scavengers. The analysis shows that ion transfer, overlap of electron cloud and change of surface hydroxyl group can cause surface charge transfer when fine particles contact with water, which leads to the recombination of solid surface hydroxyl groups and the production of H2O2[69]. This method of microdroplet generation of H2O2 uses only water as raw material and does not require energy input, which is a new method worthy of study.
Will there be H2O2 at the gas-liquid interface? The gas-liquid interface of the droplet has a strong electric field, about 109V/m[70]. This electric field is strong enough to cause the hydroxide at the gas-liquid interface to lose an electron, ionizing the hydroxide ion to form a hydroxyl radical. Because of the higher curvature of the droplets, the charge density is greater and the electric field strength is greater, resulting in an increase in the efficiency of H2O2 generation. At droplet diameters below 20 μm, the amount of H2O2 production increased significantly. In addition, hydronium and hydroxide ions are separated and unevenly distributed in the droplet, which enhances the electric field strength on the surface of the droplet[71]. Although that electric field enhance the generation of H2O2 inside the droplet, no H2O2 was detected at the gas-liquid interface. This approach yielded H2O2 concentrations up to 30 μm.
This method, which uses neither chemicals nor added energy to synthesize H2O2, requires only water and equipment to produce the spray[72,73]. Ultrasonic atomization is the direct production of H2O2 by microdroplet method. When the ultrasonic energy is high enough, the tiny bubbles (cavitation nuclei) existing in the liquid vibrate, grow and continuously accumulate energy under the action of the ultrasonic field. When the energy reaches a certain threshold, the cavitation bubbles will collapse sharply. This phenomenon is called cavitation. The cavitation phenomenon can accelerate the diffusion of oxygen in water and increase the rate of H2O2 production, and the H2O2 spray produced can easily reach more than 30 cm[74].
In order to intensify the cavitation phenomenon and increase the H2O2 production, zinc flakes were introduced on the ultrasonic nebulizer[74]. The presence of surface defects, such as cracks, on the zinc sheet increases the nucleation sites of cavitation bubbles compared to cavitation in a homogeneous liquid phase. In turn, the microjets formed by the collapse of cavitation bubbles continuously impact and grind the zinc sheet, further increasing the number of surface defects (cracks, holes, surface roughness, etc.), thus exposing more active sites. The increase of nucleation sites can improve the cavitation effect. Therefore, the corrosion of zinc flake and the formation of H2O2 can promote each other. The reaction equation for the production of H2O2 by this method is as follows.
Zn−2e→Zn2+
e+O2→∙O2
e+∙O2+2H+→H2O2
Zn2++2OH→Zn(OH)2
The yield of H2O2 was 21 times higher than that without zinc flakes, and 4.75μg/mL Zn2+ was produced in the spray droplets[74]. The yield of H2O2 was positively correlated with zinc content and reaction time. 121.25μM H2O2 was obtained by treating E. coli contaminated feeding bottles with the produced microdroplets for 30 min, and the sterilization rate reached 93. 53%. This ultrasonic atomization using a functional zinc coating successfully increased the yield of H2O2 while generating Zn2+, providing a new idea for the development of microdroplet disinfection methods[74]. Although the concentration of H2O2 produced by microdroplet method does not meet the requirement of 3% disinfection, it has application potential and broad application scenarios, and is worthy of further study.

4 Conclusion and prospect

This paper introduces a new method to produce H2O2 in situ mainly using water as raw material by providing energy through light, electricity, heat, sound, plasma and microbial metabolism. Table 1 compares the typical concentration and general yield of H2O2 produced by various techniques, and there are differences in H2O2 concentration and yield due to the use of different key materials and equipment in these methods. It can be seen that both the concentration and the yield are different from the actual disinfection requirements.
表1 以水为原料现场产H2O2方法的对比

Table 1 Performance comparison of in-situ H2O2 preparation via water oxdation

Methods Typical H2O2 concentration(%) General yield (mL/min) Ref
electrochemicistry 0.04~0.1 ~2 1,32
photocatalysis
thermocatalysis ~0.1 0.05 49
piezoelectricity ≤0.01 ~1 54
plasma ~0.03 ~1 61
microdroplet ~0.1 unlimited 71,72
In addition, there are few data on equipment cost, operation energy consumption and key material life of unit weight H2O2 produced by various methods in the existing reports. Therefore, although these methods have many ideas and methods worthy of our reference, they are still difficult to meet the requirements of actual disinfection. The main problems and countermeasures are as follows:
First, most of the reported concentrations of accumulated H2O2 do not reach the 3% required for disinfection. The key to solve this problem is to develop efficient catalytic materials, which should have H2O2 selectivity and inhibit the competitive reaction of H2O2 formation. In addition, it is necessary to clarify the self-decomposition conditions and the lower limit of self-decomposition concentration of H2O2 by combining theoretical calculation and experimental research, and to avoid these conditions in the preparation of H2O2. If there is an upper limit of the theoretical accumulation concentration of H2O2 under the preparation of H2O2 and it does not meet the disinfection requirements, the study of in-situ concentration method should be carried out, and the post-treatment process should be established to increase the concentration rather than necessarily pursuing high concentration under the preparation conditions.
Second, the reported work produces high concentrations of H2O2 at a slow rate with high energy consumption, the volume of H2O2 solution produced per hour per unit volume reactor is usually only tens of milliliters, the H2O2 produced per unit electric energy input is not more than 1 G, and the cost is much higher than that of commercial products prepared by industrial concentration, which are difficult to cope with practical disinfection applications. New materials, reactor efficiency and reaction mechanism are of great significance to solve this problem. Establish a method to control the adsorption energy of key materials on OOH groups, design a monomer reactor filled with key materials in high density, and explore the efficient mechanism of H2O2 production by imitating natural photoreaction and biological processes.
Third, the reported new materials and experimental systems are small and short-lived, and application-oriented technologies require long-lived materials and scalable material preparation methods. It is suggested that new materials should be screened in advance from the aspects of raw material source and cost, preparation process and life cycle, and the technical route for batch preparation should be established; The corrosive effect of oxidizing H2O2 on key materials should not be ignored, and measures should be taken to accelerate the separation of H2O2 from reaction sites and improve the life of materials; Develop the series or parallel connection method of unit reactors, the expansion method of material area or volume, design the reaction device with the total volume equivalent to the actual application, investigate the practicability, find out the problems existing in the application process and explore the solutions.
[1]
Shi X J, Back S, Gill T M, Siahrostami S, Zheng X L. Chem, 2021, 7(1): 38.

[2]
Brillas E, Sirés I, Oturan M A. Chem. Rev., 2009, 109(12): 6570.

[3]
Ali S, Muzslay M, Bruce M, Jeanes A, Moore G, Wilson A P R. J. Hosp. Infect., 2016, 93(1): 70.

[4]
Otter J A, Yezli S, Perl T M, Barbut F, French G L. J. Hosp. Infect., 2013, 83: 1.

[5]
Sun S X, Fan X X, Sun M J, Zhang H B, Tian L L, Zhang P P, Sun S F. Chin. J. Disinfect., 2022, 39(7): 494.

孙世雄, 樊晓旭, 孙明军, 张皓博, 田莉莉, 张培培, 孙淑芳. 中国消毒学杂志, 2022, 39(7): 494.)

[6]
Caruso A A, Del Prete A, Lazzarino A I. Med. Hypotheses, 2020, 144: 109910.

[7]
Shen M-H, Singh R K. LWT-Food Sci. Technol., 2021, 142: 110986.

[8]
Silva K J S, Sabogal-Paz L P. Water Air Soil Pollut., 2021, 232(12): 483.

[9]
Yang M F, Feng B, Bai L G, Zhao X D. Chem. Propellants Polym. Mater., 2024, 22(1): 25.

杨盟飞, 冯彬, 白立光, 赵晓东. 化学推进剂与高分子材料, 2024, 22(1): 25.)

[10]
Ingle A A, Ansari S Z, Shende D Z, Wasewar K L, Pandit A B. Environ. Sci. Pollut. Res., 2022, 29(57): 86468.

[11]
Li H B, Zheng B, Pan Z Y, Zong B N, Qiao M H. Front. Chem. Sci. Eng., 2018, 12(1): 124.

[12]
Ciriminna R, Albanese L, Meneguzzo F, Pagliaro M. Chem. Sus. Chem., 2016, 9(24): 3374.

[13]
Liu H, Xue F F. China Pulp & Paper Industry., 2022, 43(6): 24.

(刘嚆, 薛飞飞. 中华纸业, 2022, 43(6): 24.)

[14]
Liang H R, Wang L, Liu G Z. Chemical Industry and Engineering Progress., 2021, 40(4): 2060.

(梁海瑞, 王涖, 刘国柱. 化工进展, 2021, 40(4): 2060.)

[15]
Wang N, Ma S B, Zuo P J, Duan J Z, Hou B R. Adv. Sci., 2021, 8(15): 2100076.

[16]
Li N, Huang C C, Wang X, Feng Y J, An J K. Chem. Eng. J., 2022, 450: 138246.

[17]
Shen P K. Fundamental and Aplications of Electrochemical Oxygen Reduction. Nanning: Guangxi Science and technology press, 2018. 11.

(沈培康. 电化学氧还原的理论基础和应用技术. 南宁: 广西科学技术出版社, 2018. 11.)

[18]
Ali I, Van Eyck K, De Laet S, Dewil R. Chemosphere, 2022, 308: 136127.

[19]
Li D, Zheng T, Liu Y L, Hou D, Yao K K, Zhang W, Song H R, He H Y, Shi W, Wang L, Ma J. J. Hazard. Mater., 2020, 396: 122591.

[20]
Lima V N, Rodrigues C S D, Sampaio E F S, Madeira L M. J. Environ. Manag., 2020, 265: 110501.

[21]
Pérez J F, Galia A, Rodrigo M A, Llanos J, Sabatino S, Sáez C, Schiavo B, Scialdone O. Electrochim. Acta, 2017, 248: 169.

[22]
Santos G O S, Cordeiro-Junior P J M, Sánchez-Montes I, Souto R S, Kronka M S, Lanza M R V. Curr. Opin. Electrochem., 2022, 36: 101124.

[23]
Cao P K, Quan X, Zhao K, Zhao X Y, Chen S, Yu H T. ACS Catal., 2021, 11(22): 13797.

[24]
Sui W B, Li W Z, Zhang Z S, Wu W X, Xu Z Y, Wang Y. J. Power Sources, 2023, 556: 232438.

[25]
Iglesias D, Giuliani A, Melchionna M, Marchesan S, Criado A, Nasi L, Bevilacqua M, Tavagnacco C, Vizza F, Prato M, Fornasiero P. Chem., 2018, 4(1): 106.

[26]
Sun Y Y, Sinev I, Ju W, Bergmann A, Dresp S, Kühl S, Spöri C, Schmies H, Wang H, Bernsmeier D, Paul B, Schmack R, Kraehnert R, Roldan Cuenya B, Strasser P. ACS Catal., 2018, 8(4): 2844.

[27]
Zhu W Y, Chen S W. Electroanalysis, 2020, 32(12): 2591.

[28]
Li S L, Ma J X, Xu F, Wei L Y, He D. Chem. Eng. J., 2023, 452: 139371.

[29]
Tian Y H, Deng D J, Xu L, Li M, Chen H, Wu Z Z, Zhang S Q. Nano Micro. Lett., 2023, 15(1): 122.

[30]
Shin H, Lee S, Sung Y E. Curr. Opin. Electrochem., 2023, 38: 101224.

[31]
Mavrikis S, Perry S C, Leung P K, Wang L, Ponce de León C. ACS Sustainable Chem. Eng., 2021, 9(1): 76.

[32]
Pangotra D, Csepei L I, Roth A, Ponce de León C, Sieber V, Vieira L. Appl. Catal. B Environ., 2022, 303: 120848.

[33]
Espinoza-Montero P J, Alulema-Pullupaxi P, Frontana-Uribe B A, Barrera-Diaz C E. Curr. Opin. Solid State Mater. Sci., 2022, 26(3): 100988.

[34]
Shi X J, Siahrostami S, Li G L, Zhang Y R, Chakthranont P, Studt F, Jaramillo T F, Zheng X L, Nørskov J K. Nat. Commun., 2017, 8: 701.

[35]
Xia C, Back S, Ringe S, Jiang K, Chen F H, Sun X M, Siahrostami S, Chan K R, Wang H T. Nat. Catal., 2020, 3(2): 125.

[36]
Wang Z L, Xu W H, Tan G Y, Duan X X, Yuan B C, Sendeku M G, Liu H, Li T S, Wang F M, Kuang Y, Sun X M. Sci. Bull., 2023, 68(6): 613.

[37]
Fan L, Bai X W, Xia C, Zhang X, Zhao X H, Xia Y, Wu Z Y, Lu Y Y, Liu Y Y, Wang H T. Nat. Commun., 2022, 13: 2668.

[38]
Dantas L R, Wollmann L C, Suss P H, Kraft L, Ribeiro V S T, Tuon F F. Cell Tissue Bank., 2021, 22(4): 643.

[39]
Sun Y Y, Han L, Strasser P. Chem. Soc. Rev., 2020, 49(18): 6605.

[40]
Cheng H, Cheng J, Wang L, Xu H X. Chem. Mater., 2022, 34(10): 4259.

[41]
Zhang E H, Zhu Q H, Huang J H, Liu J, Tan G Q, Sun C J, Li T, Liu S, Li Y M, Wang H Z, Wan X D, Wen Z H, Fan F T, Zhang J T, Ariga K. Appl. Catal. B Environ., 2021, 293: 120213.

[42]
Wang L X, Zhang J J, Zhang Y, Yu H G, Qu Y H, Yu J G. Small, 2022, 18(8): 2104561.

[43]
Hou H L, Zeng X K, Zhang X W. Angew. Chem. Int. Ed., 2020, 59(40): 17356.

[44]
Chen Z, Yao D C, Chu C C, Mao S. Chem. Eng. J., 2023, 451: 138489.

[45]
Hargreaves J S J, Chung Y M, Ahn W S, Hisatomi T, Domen K, Kung M C, Kung H H. Appl. Catal. A: Gen., 2020, 594: 117419.

[46]
Hargreaves J S J, Chung Y M, Ahn W S, Hisatomi T, Domen K, Kung M C, Kung H H. Appl. Catal. A: Gen., 2020, 594: 117419.

[47]
Ketchie W C, Murayama M, Davis R J. Top. Catal., 2007, 44(1): 307.

[48]
Kung M C, Kung H H. Chin. J. Catal., 2019, 40(11): 1673.

[49]
Lin Y J, Khan I, Saha S, Wu C C, Barman S R, Kao F C, Lin Z H. Nat. Commun., 2021, 12: 180.

[50]
Wang K F, Shao D K, Zhang L, Zhou Y Y, Wang H P, Wang W Z. J. Mater. Chem. A, 2019, 7(35): 20383.

[51]
Hong K S, Xu H F, Konishi H, Li X C. J. Phys. Chem. C, 2012, 116(24): 13045.

[52]
Yang H Y, Wu J, Chen Z R, Zou K, Liang R H, Kang Z H, Menezes P W, Chen Z L. Energy Environ. Sci., 2023, 16(1): 210.

[53]
Wang K, Zhang M Q, Li D G, Liu L H, Shao Z P, Li X Y, Arandiyan H, Liu S M. Nano Energy, 2022, 98: 107251.

[54]
Wang C Y, Chen F, Hu C, Ma T Y, Zhang Y H, Huang H W. Chem. Eng. J., 2022, 431: 133930.

[55]
Zhou X F, Yan F, Lyubartsev A, Shen B, Zhai J W, Conesa J C, Hedin N. Adv. Sci., 2022, 9(18): 2105792.

[56]
Zhou X F, Shen B, Zhai J W, Conesa J C. Small Meth., 2021, 5(6): 2100269.

[57]
Ma Y L, Wang B, Zhong Y Z, Gao Z Y, Song H L, Zeng Y J, Wang X Y, Huang F, Li M R, Wang M Y. Chem. Eng. J., 2022, 446: 136958.

[58]
Yan W, Sun J X, Hu T, Tian S H, Feng J X, Xiong Y. Appl. Catal. B Environ., 2023, 323: 122143.

[59]
Wang L C, Chen Z S, Zhang Y H, Liu C, Yuan J P, Liu Y L, Ge W Y, Lin S, An Q, Feng Z G. Chem., 2022, 17(15): e202200278.

[60]
Xi W H, Lan Y, Shen J, Han W, Cheng C. J. Anhui Univ. Nat. Sci. Ed., 2022, 46(3): 3.

奚文灏, 兰彦, 沈洁, 韩伟, 程诚. 安徽大学学报(自然科学版), 2022, 46(3): 3.)

[61]
Burlica R, Shih K Y, Locke B R. Ind. Eng. Chem. Res., 2010, 49(14): 6342.

[62]
Ranieri P, McGovern G, Tse H, Fulmer A, Kovalenko M, Nirenberg G, Miller V, Rabinovich A, Fridman A, Fridman G. IEEE Trans. Plasma Sci., 2019, 47(1): 395.

[63]
Takeuchi N, Ishibashi N. Plasma Sources Sci. Technol., 2018, 27(4): 045010.

[64]
Wang H H, Wandell R J, Tachibana K, Voráč J, Locke B R. J. Phys. D: Appl. Phys., 2019, 52(7): 075201.

[65]
Shang K F, Li J, Wang X J, Yao D, Lu N, Jiang N, Wu Y. Jpn. J. Appl. Phys., 2016, 55(1S): 01AB02.

[66]
Tachibana K, Nakamura T. Jpn. J. Appl. Phys., 2019, 58(4): 046001.

[67]
Cameli F, Dimitrakellis P, Chen T Y, Vlachos D G. ACS Sustainable Chem. Eng., 2022, 10(5): 1829.

[68]
Chen B L, Xia Y, He R X, Sang H Q, Zhang W C, Li J, Chen L F, Wang P, Guo S S, Yin Y G, Hu L G, Song M Y, Liang Y, Wang Y W, Jiang G B, Zare R N. Proc. Natl. Acad. Sci. U. S. A., 2022, 119(32): e2209056119.

[69]
Ben-Amotz D. Science, 2022, 376(6595): 800.

[70]
Kathmann S M, Kuo I F W, Mundy C J. J. Am. Chem. Soc., 2008, 130(49): 16556.

[71]
Wei H R, Vejerano E P, Leng W N, Huang Q S, Willner M R, Marr L C, Vikesland P J. Proc. Natl. Acad. Sci. U. S. A., 2018, 115(28): 7272.

[72]
Lee J K, Walker K L, Han H S, Kang J, Prinz F B, Waymouth R M, Nam H G, Zare R N. Proc. Natl. Acad. Sci. U. S. A., 2019, 116(39): 19294.

[73]
Zhu C Q, Francisco J S. Proc. Natl. Acad. Sci. U. S. A., 2019, 116(39): 19222.

[74]
Cao T T, Tong W S, Feng F, Zhang S T, Li Y N, Liang S J, Wang X, Chen Z S, Zhang Y H. Chem. Eng. J., 2022, 431: 134005.

Outlines

/