Next Article in Journal
Recent Advances in Structured Catalysts Preparation and Use in Water-Gas Shift Reaction
Previous Article in Journal
The Roles of the Structure and Basic Sites of Sodium Titanates on Transesterification Reactions to Obtain Biodiesel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Strategies for Hydrogen Peroxide Production by Metal-Free Carbon Nitride Photocatalysts

by
André Torres-Pinto
,
Maria J. Sampaio
,
Cláudia G. Silva
*,
Joaquim L. Faria
and
Adrián M. T. Silva
Laboratory of Separation and Reaction Engineering-Laboratory of Catalysis and Materials (LSRE-LCM), Faculdade de Engenharia, Universidade do Porto, Rua Roberto Frias, 4200-465 Porto, Portugal
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(12), 990; https://doi.org/10.3390/catal9120990
Submission received: 23 October 2019 / Revised: 15 November 2019 / Accepted: 19 November 2019 / Published: 26 November 2019

Abstract

:
Hydrogen peroxide (H2O2) is a chemical which has gained wide importance in several industrial and research fields. Its mass production is mostly performed by the anthraquinone (AQ) oxidation reaction, leading to high energy consumption and significant generation of wastes. Other methods of synthesis found in the literature include the direct synthesis from oxygen and hydrogen. However, this H2O2 production process is prone to explosion hazard or undesirable by-product generation. With the growing demand of H2O2, the development of cleaner and economically viable processes has been under intense investigation. Heterogeneous photocatalysis for H2O2 production has appeared as a promising alternative since it requires only an optical semiconductor, water, oxygen, and ideally solar light irradiation. Moreover, employing a metal-free semiconductor minimizes possible toxicity consequences and reinforces the sustainability of the process. The most studied metal-free catalyst employed for H2O2 production is polymeric carbon nitride (CN). Several chemical and physical modifications over CN have been investigated together with the assessment of different sacrificial agents and light sources. This review shows the recent developments on CN materials design for enhancing the synthesis of H2O2, along with the proposed mechanisms of H2O2 production. Finally, the direct in situ generation of H2O2, when dealing with the photocatalytic synthesis of added-value organic compounds and water treatment, is discussed.

Graphical Abstract

1. Introduction

H2O2 is considered an environmentally friendly, active, and safe chemical in a broad range of applications, acting as a powerful oxidizing agent. H2O2 has been reported in environmental remediation, textile whitening, paper bleaching, chemical synthesis, cleansing of electronic materials, processing of metals, energy storage, and electric energy generation in fuel cells [1,2,3].
H2O2 is frequently present in the textile industry, reacting with the colouring matter during the bleaching process [4,5]. In organosynthesis, H2O2 has the capability to accelerate oxidation reactions, being exploited for the production of many fine and bulk chemicals [6]. Another application for H2O2 is in the field of fuel cells as it is a promising alternative energy carrier [7,8,9,10], due to its high energy density and ease/safety of storage (contrasting with H2 that presents some storage issues [11]). H2O2 can play a crucial role in the removal of pollutants from aqueous or gaseous effluents, being commonly combined with catalysts, ozone, or a light source. Wet peroxide oxidation (with and without catalysts), Fenton and photo-Fenton oxidations, and sono-, electro-, or photo-chemical reactions are some of the main processes where H2O2 is employed for water and wastewater treatment [12,13,14,15,16,17,18,19]. Particularly, H2O2 is widely used in photocatalytic water treatment because the addition or even in situ generation of H2O2 has been established as a promoter of the degradation of pollutants [20,21,22,23,24,25,26,27,28,29,30,31,32].
In response to an increasing H2O2 demand, different production processes have been developed and are currently operating at industry and laboratory scale [1]. However, many of these technologies often involve security issues regarding H2/O2 gas mixture, the use of organic solvents, metal catalysts, and the generation of undesirable by-products and solvent wastes [3]. Thus, to reduce these disadvantages, numerous approaches have been studied and developed for the synthesis of H2O2. These methods include: (i) chemical, sonochemical, and electrochemical processes, (ii) catalytic, and (iii) photocatalytic (metal-based or -free) systems.
H2O2 is produced mainly by the anthraquinone (AQ) route, commonly known as Riedl-Pfleiderer process [2,33,34,35]. This process is based on redox reactions [2,36,37], carried out in the presence of a mixture of organic solvents, e.g., ester/hydrocarbon or octanol/methyl-naphthalene. An alkylanthraquinone reacts with hydrogen, in the presence of a noble metal catalyst (e.g., Pt or Pd), and the corresponding alkylanthraquinol. The latter is oxidized under oxygenated conditions (air or O2), generating H2O2 and the former alkylanthraquinone. The quinol species can be further hydrogenated and oxygenated under the same conditions producing H2O2 and tetrahydroanthraquinone. Afterwards, H2O2 is recovered from the organic solution with water to obtain a 30% H2O2 solution, which is then distilled under reduced pressure to remove impurities and increase its concentration. The anthraquinone process avoids implementing a H2/O2 gas mixture, which is known to be explosive, with this process being economically viable for large-scale production operations. However, there are still high operating costs associated with the acquisition of the raw materials, replacement of the catalysts, and the high energy requirement during the production procedure [2]. Another interesting technology for H2O2 generation consists on the application of sonochemistry which, by dissociation of oxygen and water, leads to the formation of H2O2 [38,39,40,41]. However, limitations of ultrasounds for chemical synthesis include long exposure time that can induce H2O2 decay and the difficulty in reactor design and process scale-up [42,43].
To reduce energy costs, the design of efficient and inexpensive catalytic processes has become imperative for the generation of H2O2. The use of metallic catalysts, such as Pd- and Au-based catalysts or metal oxides (e.g., Al2O3 or TiO2), has been thoroughly documented due to their high selectivity [44,45,46,47,48,49,50,51,52,53,54,55,56]. Although several catalytic technologies have been discussed in the literature, the focus of this review is on the application of heterogeneous photocatalysis for H2O2 production. Briefly, in heterogeneous photocatalysis, an optical semiconductor is irradiated by an appropriate light source for its activation, leading to the formation of photogenerated electron/hole pairs which, under certain conditions, have the ability to produce H2O2 [57]. Depending on the reaction medium, the electronic properties of specific optical semiconductors, and the wavelength and intensity of the radiation source, heterogeneous photocatalysis can be adjusted to different processes (e.g., degradation of aqueous contaminants and production of high-value chemicals, among others [58,59]). Furthermore, photocatalytic processes may be employed under ambient temperature and pressure conditions and can be activated by free and inexhaustible solar light, further reducing the energy costs in comparison with traditional thermally activated routes. In a careful look at the literature, it was possible to retrieve more than 11,000 scientific reports on the various approaches for H2O2 synthesis (source: Scopus database, October 2018). Although photocatalytic production of H2O2 can be considered a recent topic, in recent years, the number of publications has been increasing significantly, which indicates that this technology could be considered an interesting alternative to traditional routes.
Metal-metal oxide hybrid catalysts have also been widely studied as promising candidates in the context of photocatalytic production of H2O2, owing to their high chemical stability and low toxicity [60,61,62,63]. Among all the metal elements, Ru, Bi, Co, and Cd are the most commonly combined with optical semiconductors, such as TiO2 (the standard photocatalyst) [36,64,65,66,67,68,69,70,71]. Despite their high efficiency for several photocatalytic applications, metal catalysts present viability problems and poor sustainability due to their high cost, difficulty of extraction from their ores, and possibility of leaching, which can contaminate the reaction medium and lead to the generation of hazardous wastes [72,73,74,75].
In this way, the development and optimization of metal-free structures as heterogeneous photocatalysts has been attracting wide interest. In the scope of metal-free photocatalytic H2O2 generation, the most employed material is polymeric carbon nitride (g-C3N4, here denoted as CN). Polymeric CN generally shows a relatively narrow band gap enabling visible light absorption, yet fast recombination of the photogenerated electron/hole pairs generally occurs. A sustainable process for photocatalytic H2O2 synthesis, besides requiring a visible-light activated metal-free material, requires the use of water instead of organic solvents (e.g., alcohols as sacrificial agents). CN can catalyse water splitting and selectively produce H2O2 under visible light irradiation though oxygen reduction [76,77,78].
Depending of the CN precursor and preparation method, the band potentials of CN (typically −1.12 and 1.58 eV, for the valence and conduction levels, respectively) are thermodynamically suitable for several applications, especially organic synthesis, H2 production, and pollutants degradation [79,80,81,82,83]. Several reports have shown that pristine CN materials hold a small specific surface area and low chemisorption of oxygen, commonly leading to low photocatalytic efficiencies [79,84]. Thus, several modification strategies have been pursued to increase the efficiency of CN-based photocatalysts, such as soft templating approaches, exfoliation, elemental doping, or heterojunction formation [79,85]. However, some authors report that, besides material engineering, photochemical considerations are a rather important issue to discuss since proper charge separation and transfer kinetics are needed for efficient H2O2 production [86,87].
Recently, Haider et al. [88] briefly summarized some studies concerning the synthesis of hybrid CN-based photocatalysts for H2O2 production. Regardless of the several studies reported on the use of CN for H2O2 production by heterogeneous photocatalysis, a consolidated knowledge focused on the influence of both metal-free CN structure and the manipulation of the operating conditions for improving H2O2 productivity is still missing.
In the present review, the photocatalytic production of H2O2 using metal-free CN is explored. Herein, the influence of the catalyst tuning and process optimization is correlated and discussed. The capability of simultaneously producing H2O2 during the synthesis of value-added organic compounds and their ability to induce high photocatalytic degradation of organic pollutants present in waters is a crucial point taken into consideration in this study.

2. H2O2 Production by Carbon Nitride Photocatalysts

The photocatalytic production of H2O2 can occur through several pathways, which may differ depending on the use of metallic or non-metallic catalysts. The use of metal-based photocatalysts to generate H2O2 is well discussed in the literature [62,63,89,90,91,92,93]. However, using non-metals, such as CN materials, the reaction pathway for H2O2 generation is still under study. A general scheme is depicted in Figure 1, illustrating the main steps occurring after photoactivation of CN in the presence of molecular oxygen and water. After light absorption, the migration of electrons from the valence band (VB) to the conduction band (CB) occurs. Then, electrons (e) in the CB and photogenerated holes (h+) migrate to the surface of the photocatalyst and participate in reduction and oxidation reactions, respectively.
To improve H2O2 synthesis with photoactive CN, it is necessary to understand the reactions that take place at the photocatalyst surface. The pathway for H2O2 synthesis using CN photocatalysts is generally ascribed to the capacity of this material to drive two-electron oxygen reduction. Figure 2 represents the photoactivation of CN with visible light and the mechanism for H2O2 production suggested by Shiraishi et al. [94]. The authors propose that the electron/hole pairs are localized in the 1, 4 and 2, 6 positions highlighted in Figure 2, with the negatively charged sites attracting oxygen and later reacting with the trapped protons in nearby N atoms. Then, a selected alcohol is used as sacrificial agent. Photoexcited electrons react with O2, leading to the formation of the 1,4-endoperoxide species, which results in the liberation of H2O2. At the same time, a proton donor (e.g., an alcohol or water) undergoes oxidation and yields protons that contribute to generating H2O2 [76]. The efficient formation of the endoperoxide species suppresses one- and four-electron reduction of O2 (Equation (1), (2), respectively), improving the selectivity of two-electron reduction of O2 (Equation (3)). However, O2 on pristine CN preferably undergoes reduction to O2•‒ via one-electron reduction (Equation (1)), while on modified CN materials with more surface defects, a more facile production of H2O2 is achieved (Equation (3)) [95].
O2 + e → O2•–
O2 + 4 e + 4 H+ → 2 H2O
O2 + 2 e + 2 H+ → H2O2
Shiraishi et al. [76] reported Raman spectroscopy and electron spin resonance (ESR) studies that in their work, rationalizing the mechanism behind the selective two-electron reduction of O2 on photoexcited CN. After irradiation, the Raman spectrum of CN catalyst in an O2-saturated solution showed a new broad peak at 891 cm−1, ascribed to bond vibrations in the 1,4-endoperoxide species. Concerning the ESR analyses, products of one-electron reduction of O2 were not found, proving that O2 is being selectively reduced to H2O2 [76].
In another work, authors reported the formation of H2O2 by a two-step single-electron reduction of O2 (Equations (1) and (4)), via the reduction of O2•‒ (reduction product of O2) to H2O2 [96]. Additionally, under suitable conditions (i.e., with an appropriate VB energy), water oxidation may occur, generating H2O2 (Equation (5)) [96,97]. H2O2 can also evolve via the transformation of HO formed from the hole-oxidation of HO (Equation (6), (7)) [96]. These pathways enable the production of H2O2 by both oxidation and reduction routes (Figure 1). However, H2O2 production from water and O2 is hard to facilitate using the bulk CN photocatalyst, owing to the small thermodynamic driving force between the VB energy (1.4 V) and the water oxidation potential (0.8 V) [76,98,99,100].
O2•– + e + 2 H+ → H2O2
2 H2O + 2 h+ → H2O2 + 2 H+
OH + h+ → HO
2 HO → H2O2
H2O2 + e → OH + HO
H2O2 can drive the production of HO in the presence of light (Equation (8)), depending on the radiation wavelength and catalyst employed [20,22,23]. The degradation of H2O2 to form HO radicals has a potential of +0.39 V [101] and, therefore, is favourable to occur on the CB of the semiconductor. In this way, the presence of H2O2 and photoactivated CN can generate HO radicals. These radicals are desired for several applications in environmental remediation, such as in water treatment, e.g., abatement of phenols [102], dyes [103], and antibiotics [104].

3. Enhancing Photocatalytic Activity of Carbon Nitride

In the following sections, metal-free strategies to modify CN will be overviewed and discussed to understand their influence on the efficiency and productivity of H2O2 photosynthesis. Several approaches can be found in the literature for improving the efficiency of this material, including thermal treatments, chemical substitutions with other carbon materials, or with specific organic molecules [105,106,107]. All these approaches are summarized in the following sections along with the correspondent experimental conditions of the photocatalytic tests and their ensuing results. The H2O2 productivity is depicted in terms of the highest amount produced after a respective irradiation time and moles of H2O2 per catalyst load and time, i.e., production rate (µmol gcat−1 h−1).
Multiple articles have reported the preparation of CN using distinct precursors and their posterior application using different experimental conditions [76,107,108]. In Table 1, the results in terms of H2O2 production using several neat CN materials are listed. Metal-free CN materials modified by several approaches are also shown in this table and discussed below.
Shiraishi et al. [76] reported the application of a metal-free CN material for H2O2 production using various alcohols for improving H2O2 selectivity. In this study, high H2O2 selectivity (>90%) was achieved by using aqueous solutions of aliphatic or aromatic alcohols, namely ethanol, propan-2-ol, butan-2-ol, and benzyl alcohol. Concerning the proton donor, benzyl alcohol and propan-2-ol yielded the higher amounts of H2O2. Then, testing the system using solar irradiation with or without a light filter (λ > 420 nm), higher selectivity for H2O2 formation is obtained when CN is activated by visible light rather than using the total spectrum range. This is due to ultraviolet light leading to the unwanted decomposition of H2O2.
Comparing the same matrix, light source, and gas but changing the precursor used in the catalyst synthesis, a much higher rate is achieved when using melamine [107] instead of cyanamide precursor [76].
The same group investigated the use of silica templates for changing the textural properties of CN [94]. The increase of surface defects leads to the formation of a large number of primary amines, which act as active sites for four-electron reduction of O2. A decrease on the selectivity towards H2O2 formation was observed in this system with the highest production rate of 188 µmol gcat−1 h−1 and 60% selectivity.

3.1. Surface Chemistry Modulation

Li et al. [105] reported that H2O2 production can be improved up to 14 times in the absence of an organic electron scavenger in the presence of carbon-vacancies (Cv). This study revealed that a post-treatment with argon led to the destruction of the crystallinity of CN and, therefore, produce defects, i.e., carbon vacancies (Figure 3). The Cv enhanced the trapping of the photogenerated electrons. Moreover, amino groups were formed and promoted electron transfer and changed the H2O2 production pathway from two-electron direct reduction of O2 to sequential one-electron reduction of O2 (Figure 3). In addition, it was also found that Cv decreased the band gap energy but did not interfere with the CB potential. This study showed that the chemisorption of O2 on the catalyst was enhanced with this modification. The effect of nitrogen vacancies (Nv) was also assessed, with the H2O2 production being much lower than using materials with Cv. The highest H2O2 production rate obtained for CN-Cv and CN-Nv was 900 and 150 µmol gcat−1 h−1, respectively, under visible light (λ > 420 nm) irradiation.
In another study, the effect of the presence of nitrogen defects on a CN catalyst has been discussed in terms of their ability for reducing electron-hole recombination and improving the contact between the reactant and active sites on the catalyst surface [109].
Thermal post-treatment of bulk CN drives the cleavage between tri-s-triazine moieties forming nitride vacancies and increasing C≡N groups on the matrix (Figure 4) [110]. This is accomplished by the incorporation of electron-deficient or π-conjugated monomers that falls for low bands energy potentials and inhibit the recombination. Moreover, the incorporation of strong electron acceptor groups can reduce the band gap energy and positively shift both CB and VB. The same authors investigated the impact of the saturated gas by testing the catalyst activity in the presence of O2, air, and N2 at the same conditions. It was observed that the H2O2 production rate was enhanced by saturating the suspensions with these gases, following the order O2 > air > N2. These results seems to indicate that most H2O2 production derives from O2 reduction with a small contribution from water oxidation.
Finally, another approach was performed by a plasma treatment with a power input of high voltage under H2 atmosphere with the main aim of the inclusion of N vacancies on the CN matrix [111]. These vacancies act as active sites for O2 adsorption and also have the capability to promote electron transfer, thus accelerating the reduction step. Furthermore, this catalyst suppressed the decomposition of H2O2, which resulted in a very high production rate of 2167 µmol gcat−1 h−1 in the presence of ethanol and pure O2.

3.2. Functionalization

The most recent studies found in the scope of metal-free photocatalytic H2O2 production consisted on the combination of CN with carbon, via doping or bonding with carbon nanotubes (CNT). Using C-doped CN, a positive shift of the band potentials was found (Figure 5). This change on the VB accelerates water oxidation, and on the CB enhances oxygen reduction, improving H2O2 production owing to reduced kinetic barriers of the corresponding reactions. It was verified that H2O2 production is strongly dependent on the carbon content and the electronic structure of CN. The synthesized material with highest H2O2 production simultaneously presented the highest formation and lowest decomposition rate constants, yielding a maximum H2O2 rate of 365 µmol gcat−1 h−1 in a 5% propan-2-ol solution with O2 saturation [112]. This study strengthened the claim of the formation and decomposition of H2O2 being two competitive reactions, with the formation following a zero-order kinetics due to continuous O2 saturation, and the decomposition following a first-order kinetics.
The covalent combination between CNT and CN (CN-CNT) promotes the sequential two-step pathway (Figure 2), using formic acid or methanol as electron donors. However, CN-CNT catalyses both water oxidation and O2 reduction, forming H2O2 without the need of an electron donor. This was proved by the presence of benzoquinone, which depresses H2O2 production by inhibiting the sequential two-step O2 reduction. This material reached a maximum production rate of 487 µmol gcat−1 h−1 using a 5:95 formic acid:water solution in O2-saturated conditions [106].
Another strategy to improve the CN photocatalytic activity is the anchoring of organic compounds, such as AQ. Moon et al. [113] showed that H2O2 production was dependent on the concentration of AQ and that, for higher AQ loads, there is a light block effect. When AQ is physisorbed on CN, the H2O2 production is improved. The authors explained the results based on the reutilization studies in which it was found that AQ remain at the CN surface, thus achieving a highly stable photocatalyst. H2O2 production was enhanced not only due to the higher H2O2 formation but also as a consequence of lower H2O2 decomposition. The authors also tested several AQ sources which lead to the functionalization with different groups. The material with COOH groups showed the highest formation and lowest decomposition rates, enabling continuous production over extended irradiation time in the presence of an electron donor. Furthermore, the apparent quantum yield profile resembles its absorption spectrum, meaning that efficient optical absorption and charge collection are achieved; however, with some useless recombination remains.
Benzene doping was tested by Kim et al. [114], achieving an H2O2 production rate of 300 µmol gcat−1 h−1 that was obtained in O2-saturated conditions and using a 10% ethanol aqueous solution. The increase of photoactivity compared to the bulk material was mainly ascribed to the structure distortion (Figure 6), which was promoted by the substitution of the N atoms in the matrix by benzene with a much higher molecule size. In addition, the presence of benzene leads to an easier charge transfer and hinders the recombination of electron/hole pairs.
Several authors have reported doping CN with nitrogen, oxygen and phosphate as an efficient technique for increasing the efficiency for H2O2 production [115,116,117]. In the case of N-doping, it was described that it decreases O2 adsorption energy and enhances charge transfer [115]. Doping with oxygen promoted higher light absorption and efficient charge separation with a lower recombination rate [116]. The anchoring of phosphate on CN led to enhanced O2 adsorption, which was ascribed as the main reason for the increased H2O2 productivity [117].

3.3. Construction of Heterostructures

The combination of carbon materials, like fullerene (C60), graphene oxide (GO), and reduced graphene oxide (rGO), with CN have shown to promote a negative impact for H2O2 production owing to the higher affinity to one-electron O2 reduction route [108]. Even with O2 saturation and in the presence of a propan-2-ol solution (regarded as one of the best proton donors for this process [76]), the yield of H2O2 achieved with these hybrid materials was relatively low.
Aromatic diimides are n-type semiconductors with high electron mobility and stability. Therefore, their incorporation on the CN structure can lead to a positive shift on both VB and CB bands, owing to the high electron affinity [107].
Reports have been shown that pyromellitic diimide (PDI) units increase the rates of H2O2 formation as the valence band shifts promoting water oxidation to O2, facilitating H2O2 production (Figure 7). Shiraishi et al. [107] reported the use of a CN-PDI material using water and propan-2-ol as solvents. With this study, the authors obtained a much higher rate (573 µmol gcat−1 h−1) for H2O2 production when the alcohol was present, due its capacity of acting as a strong proton donor [107]. In another work, the CN material was modified with biphenyl diimide (BDI) [77], and the effect of polymerization temperature was evaluated. The authors found an optimal temperature of 653 K, which yielded 6.8 µmol of H2O2 after 24 h of irradiation corresponding to a rate of 5.7 µmol gcat−1 h−1. By increasing the polymerization temperature, the authors found a significant catalyst weight loss, followed by a decrease on the photocatalytic activity of the resulting materials. Moreover, the amount of BDI on the CN was studied, and among all the resulting materials, the best photocatalytic activity was achieved with a molar ratio of 50% BDI in the catalyst, yielding 41 µmol after 48 h of irradiation and a 9.7 µmol gcat−1 h−1 rate. According to the calculated apparent quantum yield, BDI doping is more effective than PDI. BDI doping leads to a positive shift on the VB and CB, enabling water oxidation and promoting H2O2 formation (Figure 7). Additionally, ab initio calculations suggest that there is a significant spatial charge separation (h+ in BDI and e on melem; on PDI both h+ and e are on melem) which hinders their recombination improving H2O2 formation.
The photoactivity of catalysts, such as CN, which are π-conjugated semiconductors, depends on the density and mobility of the photoformed charge carriers [118]. The same authors that used PDI also reported the combination of CN with mellitic triimide (MTI) [78]. The incorporation of MTI units and the subsequent stacking of melem layers can lead to efficient inter and intralayer charge transfer. Therefore, the CN-MTI catalyst showed improvements on the conductivity and charge transport, as well as a higher photoactivity, compared with the pristine CN towards H2O2 production.
PDI-, BDI-, and MTI-modified CN has also been combined with reduced graphene oxide (rGO). In general, rGO has the ability to trap photogenerated electrons from the CB of the CN-PDI material, acting as active sites for the two-electron reduction of O2. On the other hand, CN-PDI-rGO photocatalyst promoted slight decomposition of H2O2. However, using a physical mixture of CN-PDI and rGO, no significant effect was observed, which can be ascribed to the low interaction between CN-PDI and rGO materials [119].
Structures of CN coupled with boron nitride (BN) were prepared to further enhance electron transfer [120,121]. The composite made with BN quantum dots favours the acceleration of charge transfer and the decrease of recombination [120]. The material with BN nanosheets resulted in an elevated production (1400 µmol gcat−1 h−1) since the authors managed to decrease H2O2 decomposition while maintaining very high formation rates [121]. The addition of BN seems to promote the separation of holes hindering recombination. Kofuji et al. have explored the combination of a CN-PDI material with BN [122]. The photogenerated holes in the VB on the CN/PDI moiety move to the VB of BN leading to their entrapment, which enhances charge separation and promotes H2O2 production. Moreover, addition of rGO further inhibits the recombination of electrons and holes. The material consisting of CN-PDI combined with both BN and rGO resulted in the highest H2O2 production rates due to promoting both water oxidation and oxygen reduction (Figure 8), while increasing the charge transfer. Moreover, these studies also supported the impact of the presence of an alcohol as proton donor. The use of propan-2-ol as solvent increased the H2O2 production rate by a factor of 50, comparing with water.
Wang et al. [96] investigated the use of perylene imide (PI), to create a Z-scheme heterojunction with CN nanosheets for enhancing the oxidative power of photogenerated holes [123]. The CN-PI structures led to higher H2O2 production; however, for excessive PI amounts, light absorption by the nanosheets was compromised. PI seems to favour the separation of charge carriers and leads to faster interfacial charge transfer (Figure 9). The addition of PI changes the production of H2O2 from a direct two-electron oxygen reduction to a two-channel pathway (Equations (3) and (4)). Moreover, the presence of PI inhibits the subsequent decomposition of H2O2 [96].
The combination of CN and black phosphorus (BP) was reported by Zheng et al. [124]. This composite allowed for a higher H2O2 productivity than the lone CN, owing to BP being highly reactive to oxygen.
To compare the different structures and experimental conditions by an unbiased parameter, the apparent quantum yield (AQY) and the solar-to-chemical conversion (SCC) of H2O2 production are typically applied. These coefficients were calculated by the respective authors and were collected in Table 2. The AQY gives the information about the formation of H2O2 relative to the number of incident photons and relates the stoichiometric amount of H2O2 formed with a specific light intensity and emission wavelength [125]. The SCC is related to the performance of a catalyst to yield H2O2, relating the free energy of H2O2 formation and the total incident energy [125]. Therefore, higher AQY and SCC values can demonstrate the photoactivity efficiency and proneness to selectively generate H2O2. For instance, Kofuji et al. [122] reported the hybrid CN-PDI-BN-rGO, which yielded the highest reported values of AQY and SCC compared to other works, as well as the highest rates for H2O2 production.

4. Photocatalytic Application with In Situ H2O2 Generation

H2O2 is used in numerous applications and, due to its oxidizing power, is commonly added in many systems, namely in the abatement of organic contaminants and in fine chemistry, to improve conversion and accelerate the reaction. For instance, the application of sonochemistry has been reported for the simultaneous in situ generation of H2O2 and degradation of organics [39,40]. The coupling of ultrasounds and photocatalysts (sonophotocatalysis) has been employed for the degradation of phenol and 4-chlorophenol with titania-based materials, where the presence of in situ evolved H2O2 markedly improved the degradation process [126]. Particularly, in photocatalysis, the presence of H2O2 is reported to increase the mineralization or removal rates of several organics [12,13,14,15,16,17,18,19] and enhance the selectivity of photochemical synthesis [6]. Therefore, the in situ generation of H2O2 in applications in which it is used as reactant is not only an advantage in terms of process design but also in terms of cost reduction, as described in the next sections.

4.1. Pollutant Degradation

The photodegradation of organic molecules and removal of biological contaminants using metal-free CN materials has been extensively investigated. Some works report the addition of H2O2, resulting in the enhancement of the degradation process [22,127,128,129,130,131]. In most cases, the CN photocatalyst acts as a Fenton-mimic since it turns H2O2 into HO which attack the pollutants, leading to increased mineralization [20,132,133]. This interesting duality of CN is worth of being explored, and several authors have already verified the in situ evolution of H2O2 to enhance the oxidative abatement of contaminants. As previously discussed, metal-free CN, under visible-light, leads to the formation of H2O2, and is used for the removal of several organic compounds by photocatalysis. Many studies report the formation of H2O2 simultaneously to the degradation of the contaminant molecules [20,23,120,134,135,136,137,138,139,140,141]. The degradation is accompanied by the formation of reactive oxygen species (ROS), H2O2 being detected during the photocatalytic experiments. Two studies recently followed the time-dependent H2O2 concentration along the photocatalytic degradation reaction of phenol [95,142]. Zhang et al. detected H2O2 using CN nanosheets under visible light irradiation with the production being markedly dependent in the structure of CN, yielding larger amounts for more exfoliated materials which promote selective two-electron O2 reduction [95]. Additionally, the formation of highly reactive oxygen species promoted phenol degradation, such as HO originated from H2O2 decomposition. The presence of O2 in an aqueous solution can lead to the formation of H2O2 and other reactive oxygen species which aid the oxidation of organic molecules. H2O2 is a fast reacting molecule; however, its stabilization and production is very dependent of the medium [1]. For instance, the pH dependency of H2O2 is known and, in more acidic media, H2O2 is more stable than in an alkaline environment [143]. However, CN has been proven to act efficiently in all pH range since this material presents amphoteric properties [144,145]. Relative to H2O2 formation, it is observed that many other factors have to be taken into consideration, namely the content in dissolved oxygen, pollutant initial concentration, catalyst load and light source. The work developed by Svoboda et al. [142] somewhat showed the impact of these parameters on H2O2 production since quenching experiments lead to changes on the degradation of phenol (sacrificial agent for H2O2 formation). However, the use of scavenging species, to study H2O2 formation and decomposition, may suffer interference owing to their degradation by different reactive oxy-species. This can hinder or facilitate the generation of H2O2 leading to ambiguous results. Svoboda et al. [142] investigated the degradation of phenol using CN nanosheets obtained from the thermal post-treatment of melamine-derived bulk CN. This allowed for an increase of the surface area of the material, leading to much higher photoactivity. In this work, an impressive H2O2 production rate of 3300 µmol gcat−1 h−1 was reported, using visible-LEDs with a maximum emission wavelength of λ = 416 nm, continuous air purging and a phenol initial concentration of 20 mg L−1.
Furthermore, using differently-substituted phenolic compounds and exfoliated CN it is possible to obtain high production rates between 633 and 3103 µmol gcat−1 h−1 [146], using visible-light emitting diodes and continuous air saturation. This study contemplates the relation between O2 concentration and H2O2 production, establishing that as the dissolved oxygen content increases up to 21% there is an increase in H2O2 formation. Moreover, the exfoliated CN in this work was tested in a propan-2-ol aqueous solution, achieving a H2O2 production rate of 19,200 µmol gcat−1 h−1.
To date, the combination of oxidation driven by in situ evolved H2O2 in CN photocatalysts with other AOPs for water treatment has been reported for ozonation and persulfate activation [147,148]. These two studies investigate the in situ evolution of H2O2 during the reaction and discuss the synergic effect that promoted the removal of the contaminant molecules. However, it is interesting to point out that H2O2 is fundamental in the Fenton reaction. In addition, in a homogeneous Fenton system, the treated water matrix remains with dissolved iron which has to be separated. Iron-doped CN photocatalysts result on the combination of traditional Fenton and CN photocatalysis which enhances oxidation by promoting a two-channel pathway of H2O2 reduction to generate HO. In this way, many authors have synthesized iron-doped CN to try combat the disadvantage of dissolved iron in the mineralized waters [149,150,151,152].
CN has been applied as metal-free photocatalyst for the degradation of several organic pollutants, but the generation of H2O2 was not monitored in the publications [153,154]; thus, they are not considered in this review.

4.2. Fine Chemistry

In the case of selective organic synthesis, there are reports of H2O2 addition while employing metal-free CN photocatalysts [155,156]. The presence of H2O2 improves the oxidation of selected molecules, such as the conversion of toluene to benzaldehyde [155] or of cyclic olefins to the respective epoxides [156]. However, there have been reports where H2O2 formation was observed in the presence of both visible light and a CN catalyst. Lopes et al. [157] achieved very high production rates of ca. 5000 µmol gcat−1 h−1 using nanosheets of CN in an anisyl alcohol solution. In that work, H2O2 was formed as a by-product of the oxidation of aromatic alcohols into the corresponding aldehydes. The simultaneous formation of H2O2 is a further advantage to the selective organic synthesis owing to the oxidizing power of H2O2, such as Zhang et al. [158] describes with an oxygen-enriched CN material employed for the transformation of amines into imines.

5. Conclusions and Future Prospects

This review article summarizes the state-of-the-art on modified metal-free carbon nitride photocatalysts for the selective evolution of H2O2, a high-value and multi-faceted chemical. Conventional processes for H2O2 synthesis are generally characterized by high energy consumption and waste generation. The latest studies, employing sustainable metal-free carbon nitride materials and clean aqueous matrices as solvents, show an emergent photocatalytic technology for H2O2 generation, including the smart tailoring of these materials for optimal conversion. Moreover, the ambivalence of the photocatalytic process, with simultaneous production and direct application of H2O2, has already been explored for pollutant degradation and fine chemical synthesis. The referred studies provide a favourable starting point to achieve sustainability in the industry of H2O2 production, albeit more research has to be performed to develop the necessary scale up operation, productivity enhancement, and overall process optimization.

Author Contributions

Conceptualization, all authors; methodology, A.T.-P. and M.J.S.; investigation, A.T.-P.; resources, J.L.F.; C.G.S. and A.M.T.S.; writing—original draft preparation, A.T.-P. and M.J.S.; writing—review and editing, all authors; supervision, C.G.S.; J.L.F. and A.M.T.S.

Funding

This work was financially supported by project NORTE-01-0145-FEDER-031049 (InSpeCt, PTDC/EAM-AMB/31049/2017) funded by the European Regional Development Fund (ERDF) through NORTE 2020 - Programa Operacional Regional do NORTE and by national funds (PIDDAC) through FCT-Fundação para a Ciência e a Tecnologia, and by projects POCI-01-0145-FEDER-030674, POCI-01-0145-FEDER-031398 and POCI-01-0145-FEDER-029600, funded by ERDF through COMPETE2020 – Programa Operacional Competitividade e Internacionalização (POCI) – and by national funds through FCT. We would also like to thank the scientific collaboration under project “AIProcMat@N2020 - Advanced Industrial Processes and Materials for a Sustainable Northern Region of Portugal 2020”, with the reference NORTE-01-0145-FEDER-000006, supported by NORTE 2020 under the Portugal 2020 Partnership Agreement through ERDF, and project Associate Laboratory LSRE-LCM - UID/EQU/50020/2019 funded by national funds through FCT/MCTES (PIDDAC). C.G.S. acknowledges the FCT Investigator Programme (IF/00514/2014) with financing from the European Social Fund (ESF) and the Human Potential Operational Programme.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hess, W.T. Hydrogen Peroxide. In Kirk-Othmer Encyclopedia of Chemical Technology, 5th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2007; Volume 14, pp. 35–79. [Google Scholar]
  2. Samanta, C. Direct synthesis of hydrogen peroxide from hydrogen and oxygen: An overview of recent developments in the process. Appl. Catal. A Gen. 2008, 350, 133–149. [Google Scholar] [CrossRef]
  3. Campos-Martin, J.M.; Blanco-Brieva, G.; Fierro, J.L.G. Hydrogen Peroxide Synthesis: An Outlook beyond the Anthraquinone Process. Angew. Chem. Int. Ed. 2006, 45, 6962–6984. [Google Scholar] [CrossRef]
  4. Asghar, A.; Raman, A.A.A.; Wan Daud, W.M.A. Advanced oxidation processes for in-situ production of hydrogen peroxide/hydroxyl radical for textile wastewater treatment: A review. J. Clean. Prod. 2015, 87, 826–838. [Google Scholar] [CrossRef]
  5. Karmakar, S.R. Chapter 6—Bleaching of textiles. In Chemical Techonology in the Pre-Treatment Processes of Textiles; Elsevier: Amsterdam, The Netherlands, 1999; Volume 12, pp. 160–216. [Google Scholar]
  6. Sato, K.; Aoki, M.; Noyori, R. A “Green” Route to Adipic Acid: Direct Oxidation of Cyclohexenes with 30 Percent Hydrogen Peroxide. Science 1998, 281, 1646–1647. [Google Scholar] [CrossRef]
  7. Yamazaki, S.-I.; Siroma, Z.; Senoh, H.; Ioroi, T.; Fujiwara, N.; Yasuda, K. A fuel cell with selective electrocatalysts using hydrogen peroxide as both an electron acceptor and a fuel. J. Power Sour. 2008, 178, 20–25. [Google Scholar] [CrossRef]
  8. Shaegh, S.A.M.; Nguyen, N.-T.; Ehteshami, S.M.M.; Chan, S.H. A membraneless hydrogen peroxide fuel cell using Prussian Blue as cathode material. Energy Environ. Sci. 2012, 5, 8225–8228. [Google Scholar] [CrossRef]
  9. Yamada, Y.; Yoneda, M.; Fukuzumi, S. A Robust One-Compartment Fuel Cell with a Polynuclear Cyanide Complex as a Cathode for Utilizing H2O2 as a Sustainable Fuel at Ambient Conditions. Chem. Eur. J. 2013, 19, 11733–11741. [Google Scholar] [CrossRef] [PubMed]
  10. Yang, F.; Cheng, K.; Wu, T.; Zhang, Y.; Yin, J.; Wang, G.; Cao, D. Preparation of Au nanodendrites supported on carbon fiber cloth and its catalytic performance to H2O2 electroreduction and electrooxidation. RSC Adv. 2013, 3, 5483–5490. [Google Scholar] [CrossRef]
  11. Fukuzumi, S. Bioinspired Energy Conversion Systems for Hydrogen Production and Storage. Eur. J. Inorg. Chem. 2008, 2008, 1351–1362. [Google Scholar] [CrossRef]
  12. Perathoner, S.; Centi, G. Wet hydrogen peroxide catalytic oxidation (WHPCO) of organic waste in agro-food and industrial streams. Top. Catal. 2005, 33, 207–224. [Google Scholar] [CrossRef]
  13. Rokhina, E.V.; Virkutyte, J. Environmental Application of Catalytic Processes: Heterogeneous Liquid Phase Oxidation of Phenol With Hydrogen Peroxide. Crit. Rev. Environ. Sci. Technol. 2010, 41, 125–167. [Google Scholar] [CrossRef]
  14. Debellefontaine, H.; Chakchouk, M.; Foussard, J.N.; Tissot, D.; Striolo, P. Treatment of organic aqueous wastes: Wet air oxidation and wet peroxide oxidation®. Environ. Pollut. 1996, 92, 155–164. [Google Scholar] [CrossRef]
  15. Domingues, F.S.; Freitas, T.K.F.D.S.; de Almeida, C.A.; de Souza, R.P.; Ambrosio, E.; Palácio, S.M.; Garcia, J.C. Hydrogen peroxide-assisted photocatalytic degradation of textile wastewater using titanium dioxide and zinc oxide. Environ. Technol. 2017, 40, 1223–1232. [Google Scholar] [CrossRef] [PubMed]
  16. Gomes, H.T.; Miranda, S.M.; Sampaio, M.J.; Silva, A.M.T.; Faria, J.L. Activated carbons treated with sulphuric acid: Catalysts for catalytic wet peroxide oxidation. Catal. Today 2010, 151, 153–158. [Google Scholar] [CrossRef]
  17. Ribeiro, R.S.; Silva, A.M.T.; Figueiredo, J.L.; Faria, J.L.; Gomes, H.T. Removal of 2-nitrophenol by catalytic wet peroxide oxidation using carbon materials with different morphological and chemical properties. Appl. Catal. B Environ. 2013, 140, 356–362. [Google Scholar] [CrossRef]
  18. Ribeiro, R.S.; Silva, A.M.T.; Figueiredo, J.L.; Faria, J.L.; Gomes, H.T. Catalytic wet peroxide oxidation: A route towards the application of hybrid magnetic carbon nanocomposites for the degradation of organic pollutants. A review. Appl. Catal. B Environ. 2016, 187, 428–460. [Google Scholar] [CrossRef]
  19. Ribeiro, R.S.; Silva, A.M.T.; Pastrana-Martínez, L.M.; Figueiredo, J.L.; Faria, J.L.; Gomes, H.T. Graphene-based materials for the catalytic wet peroxide oxidation of highly concentrated 4-nitrophenol solutions. Catal. Today 2015, 249, 204–212. [Google Scholar] [CrossRef]
  20. Cui, Y.; Huang, J.; Fu, X.; Wang, X. Metal-free photocatalytic degradation of 4-chlorophenol in water by mesoporous carbon nitride semiconductors. Catal. Sci. Technol. 2012, 2, 1396–1402. [Google Scholar] [CrossRef]
  21. Yao, Y.; Wu, G.; Lu, F.; Wang, S.; Hu, Y.; Zhang, J.; Huang, W.; Wei, F. Enhanced photo-Fenton-like process over Z-scheme CoFe2O4/g-C3N4 Heterostructures under natural indoor light. Environ. Sci. Pollut. Res. 2016, 23, 21833–21845. [Google Scholar] [CrossRef]
  22. Kumar, A.; Kumar, A.; Sharma, G.; Naushad, M.; Veses, R.C.; Ghfar, A.A.; Stadler, F.J.; Khan, M.R. Solar-driven photodegradation of 17-β-estradiol and ciprofloxacin from waste water and CO2 conversion using sustainable coal-char/polymeric-g-C3N4/rGO metal-free nano-hybrids. New. J. Chem. 2017, 41, 10208–10224. [Google Scholar] [CrossRef]
  23. Zhou, C.; Lai, C.; Huang, D.; Zeng, G.; Zhang, C.; Cheng, M.; Hu, L.; Wan, J.; Xiong, W.; Wen, M.; et al. Highly porous carbon nitride by supramolecular preassembly of monomers for photocatalytic removal of sulfamethazine under visible light driven. Appl. Catal. B Environ. 2018, 220, 202–210. [Google Scholar] [CrossRef]
  24. Su, M.; He, C.; Sharma, V.K.; Asi, M.A.; Xia, D.; Li, X.Z.; Deng, H.; Xiong, Y. Mesoporous zinc ferrite: Synthesis, characterization, and photocatalytic activity with H2O2/visible light. J. Hazard. Mater. 2012, 211, 95–103. [Google Scholar] [CrossRef] [PubMed]
  25. Jia, Y.; Lee, B.W.; Liu, C. Magnetic ZnFe2O4 Nanocubes: Synthesis and Photocatalytic Activity with Visible Light/H2O2. IEEE Trans. Magn. 2017, 53, 1–5. [Google Scholar] [CrossRef]
  26. Xu, J.; Wang, H.; Gu, H.; Zeng, C.; Yang, Y. Facile synthesis of Cu2O nanocubes and their enhanced photocatalytic property assisted by H2O2. Asian J. Chem. 2013, 25, 1733–1736. [Google Scholar]
  27. Kalam, A.; Al-Sehemi, A.G.; Assiri, M.; Du, G.; Ahmad, T.; Ahmad, I.; Pannipara, M. Modified solvothermal synthesis of cobalt ferrite (CoFe2O4) magnetic nanoparticles photocatalysts for degradation of methylene blue with H2O2/visible light. Results Phys. 2018, 8, 1046–1053. [Google Scholar] [CrossRef]
  28. Deng, X.; Wang, C.; Shao, M.; Xu, X.; Huang, J. Low-temperature solution synthesis of CuO/Cu2O nanostructures for enhanced photocatalytic activity with added H2O2: Synergistic effect and mechanism insight. RSC Adv. 2017, 7, 4329–4338. [Google Scholar] [CrossRef]
  29. Pesakhov, S.; Benisty, R.; Sikron, N.; Cohen, Z.; Gomelsky, P.; Khozin-Goldberg, I.; Dagan, R.; Porat, N. Effect of hydrogen peroxide production and the Fenton reaction on membrane composition of Streptococcus pneumoniae. BBA Biomembr. 2007, 1768, 590–597. [Google Scholar] [CrossRef]
  30. Pham, A.N.; Xing, G.; Miller, C.J.; Waite, T.D. Fenton-like copper redox chemistry revisited: Hydrogen peroxide and superoxide mediation of copper-catalyzed oxidant production. J. Catal. 2013, 301, 54–64. [Google Scholar] [CrossRef]
  31. Fujihira, M.; Satoh, Y.; Osa, T. Heterogeneous photocatalytic oxidation of aromatic compounds on TiO2. Nature 1981, 293, 206. [Google Scholar] [CrossRef]
  32. Chiou, C.-H.; Wu, C.-Y.; Juang, R.-S. Influence of operating parameters on photocatalytic degradation of phenol in UV/TiO2 process. Chem. Eng. J. 2008, 139, 322–329. [Google Scholar] [CrossRef]
  33. Nishimi, T.; Kamachi, T.; Kato, K.; Kato, T.; Yoshizawa, K. Mechanistic Study on the Production of Hydrogen Peroxide in the Anthraquinone Process. Eur. J. Org. Chem. 2011, 2011, 4113–4120. [Google Scholar] [CrossRef]
  34. Hans-Joachim, P.G.R. Production of Hydrogen Peroxide. US Patent 2158525, 10 October 1935. [Google Scholar]
  35. Thenard, L.J. Observations sur des nouvelles combinaisons entre l’oxigène et divers acides. Ann. Chim. Phys. 1818, 8, 306–312. [Google Scholar]
  36. Kato, S.; Jung, J.; Suenobu, T.; Fukuzumi, S. Production of hydrogen peroxide as a sustainable solar fuel from water and dioxygen. Energy Environ. Sci. 2013, 6, 3756–3764. [Google Scholar] [CrossRef] [Green Version]
  37. Sandelin, F.; Oinas, P.; Salmi, T.; Paloniemi, J.; Haario, H. Kinetics of the Recovery of Active Anthraquinones. Ind. Eng. Chem. Res. 2006, 45, 986–992. [Google Scholar] [CrossRef]
  38. Lim, M.; Son, Y.; Khim, J. The effects of hydrogen peroxide on the sonochemical degradation of phenol and bisphenol A. Ultrason. Sonochem. 2014, 21, 1976–1981. [Google Scholar] [CrossRef] [PubMed]
  39. Pétrier, C. 31—The use of power ultrasound for water treatment. In Power Ultrasonics; Gallego-Juárez, J.A., Graff, K.F., Eds.; Woodhead Publishing: Oxford, UK, 2015; pp. 939–972. [Google Scholar]
  40. Shende, T.; Andaluri, G.; Suri, R.P.S. Kinetic model for sonolytic degradation of non-volatile surfactants: Perfluoroalkyl substances. Ultrason. Sonochem. 2019, 51, 359–368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Ziembowicz, S.; Kida, M.; Koszelnik, P. The impact of selected parameters on the formation of hydrogen peroxide by sonochemical process. Sep. Purif. Technol. 2018, 204, 149–153. [Google Scholar] [CrossRef]
  42. Kiss, A.A.; Geertman, R.; Wierschem, M.; Skiborowski, M.; Gielen, B.; Jordens, J.; John, J.J.; Van Gerven, T. Ultrasound-assisted emerging technologies for chemical processes. J. Chem. Technol. Biotechnol. 2018, 93, 1219–1227. [Google Scholar] [CrossRef]
  43. Ashokkumar, M. Advantages, Disadvantages and Challenges of Ultrasonic Technology. In Ultrasonic Synthesis of Functional Materials; Springer International Publishing: Cham, Switzerland, 2016; pp. 41–42. [Google Scholar]
  44. Chinta, S.; Lunsford, J.H. A mechanistic study of H2O2 and H2O formation from H2 and O2 catalyzed by palladium in an aqueous medium. J. Catal. 2004, 225, 249–255. [Google Scholar] [CrossRef]
  45. Dissanayake, D.P.; Lunsford, J.H. Evidence for the Role of Colloidal Palladium in the Catalytic Formation of H2O2 from H2 and O2. J. Catal. 2002, 206, 173–176. [Google Scholar] [CrossRef]
  46. Dissanayake, D.P.; Lunsford, J.H. The direct formation of H2O2 from H2 and O2 over colloidal palladium. J. Catal. 2003, 214, 113–120. [Google Scholar] [CrossRef]
  47. Liu, Q.; Lunsford, J.H. The roles of chloride ions in the direct formation of H2O2 from H2 and O2 over a Pd/SiO2 catalyst in a H2SO4/ethanol system. J. Catal. 2006, 239, 237–243. [Google Scholar] [CrossRef]
  48. Lunsford, J.H. The direct formation of H2O2 from H2 and O2 over palladium catalysts. J. Catal. 2003, 216, 455–460. [Google Scholar] [CrossRef]
  49. Choudhary, V.R.; Gaikwad, A.G.; Sansare, S.D. Activation of Supported Pd Metal Catalysts for Selective Oxidation of Hydrogen to Hydrogen Peroxide. Catal. Lett. 2002, 83, 235–239. [Google Scholar] [CrossRef]
  50. Choudhary, V.R.; Samanta, C. Hydrogen peroxide formation in the interaction of oxygen with boron-containing Pd catalysts prereduced by hydrazine in aqueous acidic medium containing bromide anions. Catal. Lett. 2005, 99, 79–81. [Google Scholar] [CrossRef]
  51. Choudhary, V.R.; Samanta, C.; Gaikwad, A.G. Drastic increase of selectivity for H2O2 formation in direct oxidation of H2 to H2O2 over supported Pd catalysts due to their bromination. Chem. Commun. 2004, 2054–2055. [Google Scholar] [CrossRef]
  52. Choudhary, V.R.; Samanta, C.; Jana, P. A novel route for in-situ H2O2 generation from selective reduction of O2 by hydrazine using heterogeneous Pd catalyst in an aqueous medium. Chem. Commun. 2005, 5399–5401. [Google Scholar] [CrossRef]
  53. Choudhary, V.R.; Sansare, S.D.; Gaikwad, A.G. Direct Oxidation of H2 to H2O2 and Decomposition of H2O2 Over Oxidized and Reduced Pd-Containing Zeolite Catalysts in Acidic Medium. Catal. Lett. 2002, 84, 81–87. [Google Scholar] [CrossRef]
  54. Edwards, J.K.; Hutchings, G.J. Palladium and Gold–Palladium Catalysts for the Direct Synthesis of Hydrogen Peroxide. Angew. Chem. Int. Ed. 2008, 47, 9192–9198. [Google Scholar] [CrossRef]
  55. Edwards, J.K.; Solsona, B.; Landon, P.; Carley, A.F.; Herzing, A.; Watanabe, M.; Kiely, C.J.; Hutchings, G.J. Direct synthesis of hydrogen peroxide from H2 and O2 using Au-Pd/Fe2O3 catalysts. J. Mater. Chem. 2005, 15, 4595–4600. [Google Scholar] [CrossRef]
  56. Edwards, J.K.; Solsona, B.E.; Landon, P.; Carley, A.F.; Herzing, A.; Kiely, C.J.; Hutchings, G.J. Direct synthesis of hydrogen peroxide from H2 and O2 using TiO2-supported Au-Pd catalysts. J. Catal. 2005, 236, 69–79. [Google Scholar] [CrossRef]
  57. Chen, J.; Eberlein, L.; Langford, C.H. Pathways of phenol and benzene photooxidation using TiO2 supported on a zeolite. J. Photochem. Photobiol. A 2002, 148, 183–189. [Google Scholar] [CrossRef]
  58. Ahmed, S.N.; Haider, W. Heterogeneous photocatalysis and its potential applications in water and wastewater treatment: A review. Nanotechnology 2018, 29, 342001. [Google Scholar] [CrossRef] [Green Version]
  59. Radhika, N.P.; Selvin, R.; Kakkar, R.; Umar, A. Recent advances in nano-photocatalysts for organic synthesis. Arab. J. Chem. 2016. [Google Scholar] [CrossRef] [Green Version]
  60. Kaynan, N.; Berke, B.A.; Hazut, O.; Yerushalmi, R. Sustainable photocatalytic production of hydrogen peroxide from water and molecular oxygen. J. Mater. Chem. A 2014, 2, 13822–13826. [Google Scholar] [CrossRef]
  61. Carraway, E.R.; Hoffman, A.J.; Hoffmann, M.R. Photocatalytic Oxidation of Organic Acids on Quantum-Sized Semiconductor Colloids. Environ. Sci. Technol. 1994, 28, 786–793. [Google Scholar] [CrossRef]
  62. Kormann, C.; Bahnemann, D.W.; Hoffmann, M.R. Photocatalytic production of hydrogen peroxides and organic peroxides in aqueous suspensions of titanium dioxide, zinc oxide, and desert sand. Environ. Sci. Technol. 1988, 22, 798–806. [Google Scholar] [CrossRef]
  63. Teranishi, M.; Naya, S.I.; Tada, H. In situ liquid phase synthesis of hydrogen peroxide from molecular oxygen using gold nanoparticle-loaded titanium(IV) dioxide photocatalyst. J. Am. Chem. Soc. 2010, 132, 7850–7851. [Google Scholar] [CrossRef]
  64. Zhuang, H.; Yang, L.; Xu, J.; Li, F.; Zhang, Z.; Lin, H.; Long, J.; Wang, X. Robust Photocatalytic H2O2 Production by Octahedral Cd3(C3N3S3)2 Coordination Polymer under Visible Light. Sci. Rep. 2015, 5, 16947. [Google Scholar] [CrossRef] [Green Version]
  65. Shao, D.; Zhang, L.; Sun, S.; Wang, W. Oxygen Reduction Reaction for Generating H2O2 through a Piezo-Catalytic Process over Bismuth Oxychloride. ChemSusChem 2018, 11, 527–531. [Google Scholar] [CrossRef] [Green Version]
  66. Su, Y.; Zhang, L.; Wang, W.; Shao, D. Internal Electric Field Assisted Photocatalytic Generation of Hydrogen Peroxide over BiOCl with HCOOH. ACS Sustain. Chem. Eng. 2018, 6, 8704–8710. [Google Scholar] [CrossRef]
  67. Isaka, Y.; Kato, S.; Hong, D.; Suenobu, T.; Yamada, Y.; Fukuzumi, S. Bottom-up and top-down methods to improve catalytic reactivity for photocatalytic production of hydrogen peroxide using a Ru-complex and water oxidation catalysts. J. Mater. Chem. A 2015, 3, 12404–12412. [Google Scholar] [CrossRef] [Green Version]
  68. Isaka, Y.; Oyama, K.; Yamada, Y.; Suenobu, T.; Fukuzumi, S. Photocatalytic production of hydrogen peroxide from water and dioxygen using cyano-bridged polynuclear transition metal complexes as water oxidation catalysts. Catal. Sci. Technol. 2016, 6, 681–684. [Google Scholar] [CrossRef] [Green Version]
  69. Isaka, Y.; Yamada, Y.; Suenobu, T.; Nakagawa, T.; Fukuzumi, S. Production of hydrogen peroxide by combination of semiconductor-photocatalysed oxidation of water and photocatalytic two-electron reduction of dioxygen. RSC Adv. 2016, 6, 42041–42044. [Google Scholar] [CrossRef] [Green Version]
  70. Mase, K.; Yoneda, M.; Yamada, Y.; Fukuzumi, S. Efficient Photocatalytic Production of Hydrogen Peroxide from Water and Dioxygen with Bismuth Vanadate and a Cobalt(II) Chlorin Complex. ACS Energy Lett. 2016, 1, 913–919. [Google Scholar] [CrossRef]
  71. Thakur, S.; Kshetri, T.; Kim, N.H.; Lee, J.H. Sunlight-driven sustainable production of hydrogen peroxide using a CdS–graphene hybrid photocatalyst. J. Catal. 2017, 345, 78–86. [Google Scholar] [CrossRef]
  72. Charanpahari, A.; Gupta, N.; Devthade, V.; Ghugal, S.; Bhatt, J. Ecofriendly Nanomaterials for Sustainable Photocatalytic Decontamination of Organics and Bacteria. In Handbook of Ecomaterials; Martínez, L.M.T., Kharissova, O.V., Kharisov, B.I., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 1–29. [Google Scholar]
  73. Chiranjeevi, T.; Pragya, R.; Gupta, S.; Gokak, D.T.; Bhargava, S. Minimization of Waste Spent Catalyst in Refineries. Procedia Environ. Sci. 2016, 35, 610–617. [Google Scholar] [CrossRef]
  74. Monai, M.; Melchionna, M.; Fornasiero, P. Chapter One—From metal to metal-free catalysts: Routes to sustainable chemistry. In Advances in Catalysis; Song, C., Ed.; Academic Press: Cambridge, MA, USA, 2018; Volume 63, pp. 1–73. [Google Scholar]
  75. Gautam, R.K.; Chattopadhyaya, M.C. Chapter 9—Magnetic Nanophotocatalysts for Wastewater Remediation. In Nanomaterials for Wastewater Remediation; Gautam, R.K., Chattopadhyaya, M.C., Eds.; Butterworth-Heinemann: Boston, MA, USA, 2016; pp. 189–238. [Google Scholar]
  76. Shiraishi, Y.; Kanazawa, S.; Sugano, Y.; Tsukamoto, D.; Sakamoto, H.; Ichikawa, S.; Hirai, T. Highly selective production of hydrogen peroxide on graphitic carbon nitride (g-C3N4) photocatalyst activated by visible light. ACS Catal. 2014, 4, 774–780. [Google Scholar] [CrossRef]
  77. Kofuji, Y.; Ohkita, S.; Shiraishi, Y.; Sakamoto, H.; Tanaka, S.; Ichikawa, S.; Hirai, T. Graphitic Carbon Nitride Doped with Biphenyl Diimide: Efficient Photocatalyst for Hydrogen Peroxide Production from Water and Molecular Oxygen by Sunlight. ACS Catal. 2016, 6, 7021–7029. [Google Scholar] [CrossRef]
  78. Kofuji, Y.; Ohkita, S.; Shiraishi, Y.; Sakamoto, H.; Ichikawa, S.; Tanaka, S.; Hirai, T. Mellitic Triimide-Doped Carbon Nitride as Sunlight-Driven Photocatalysts for Hydrogen Peroxide Production. ACS Sustain. Chem. Eng. 2017, 5, 6478–6485. [Google Scholar] [CrossRef]
  79. Ong, W.-J.; Tan, L.-L.; Ng, Y.H.; Yong, S.-T.; Chai, S.-P. Graphitic Carbon Nitride (g-C3N4)-Based Photocatalysts for Artificial Photosynthesis and Environmental Remediation: Are We a Step Closer To Achieving Sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef] [PubMed]
  80. Patnaik, S.; Martha, S.; Acharya, S.; Parida, K.M. An overview of the modification of g-C3N4 with high carbon containing materials for photocatalytic applications. Inorg. Chem. Front. 2016, 3, 336–347. [Google Scholar] [CrossRef]
  81. Zhao, Z.; Sun, Y.; Dong, F. Graphitic carbon nitride based nanocomposites: A review. Nanoscale 2015, 7, 15–37. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, G.; Lan, Z.A.; Wang, X. Surface engineering of graphitic carbon nitride polymers with cocatalysts for photocatalytic overall water splitting. Chem. Sci. 2017, 8, 5261–5274. [Google Scholar] [CrossRef] [Green Version]
  83. Mamba, G.; Mishra, A.K. Graphitic carbon nitride (g-C3N4) nanocomposites: A new and exciting generation of visible light driven photocatalysts for environmental pollution remediation. Appl. Catal. B Environ. 2016, 198, 347–377. [Google Scholar] [CrossRef]
  84. Lima, M.J.; Silva, A.M.T.; Silva, C.G.; Faria, J.L. Graphitic carbon nitride modified by thermal, chemical and mechanical processes as metal-free photocatalyst for the selective synthesis of benzaldehyde from benzyl alcohol. J. Catal. 2017, 353, 44–53. [Google Scholar] [CrossRef]
  85. Wang, Y.; Wang, X.; Antonietti, M. Polymeric Graphitic Carbon Nitride as a Heterogeneous Organocatalyst: From Photochemistry to Multipurpose Catalysis to Sustainable Chemistry. Angew. Chem. Int. Ed. 2012, 51, 68–89. [Google Scholar] [CrossRef]
  86. Domcke, W.; Ehrmaier, J.; Sobolewski, A.L. Solar Energy Harvesting with Carbon Nitrides and N-Heterocyclic Frameworks: Do We Understand the Mechanism? ChemPhotoChem 2019, 3, 10–23. [Google Scholar] [CrossRef]
  87. Corp, K.L.; Schlenker, C.W. Ultrafast Spectroscopy Reveals Electron-Transfer Cascade That Improves Hydrogen Evolution with Carbon Nitride Photocatalysts. J. Am. Chem. Soc. 2017, 139, 7904–7912. [Google Scholar] [CrossRef]
  88. Haider, Z.; Cho, H.-I.; Moon, G.-H.; Kim, H.-I. Minireview: Selective production of hydrogen peroxide as a clean oxidant over structurally tailored carbon nitride photocatalysts. Catal. Today 2018, 335, 55–64. [Google Scholar] [CrossRef]
  89. Goto, H.; Hanada, Y.; Ohno, T.; Matsumura, M. Quantitative analysis of superoxide ion and hydrogen peroxide produced from molecular oxygen on photoirradiated TiO2 particles. J. Catal. 2004, 225, 223–229. [Google Scholar] [CrossRef]
  90. Maurino, V.; Minero, C.; Mariella, G.; Pelizzetti, E. Sustained production of H2O2 on irradiated TiO2—Fluoride systems. Chem. Commun. 2005, 2627–2629. [Google Scholar] [CrossRef] [PubMed]
  91. Hirakawa, T.; Nosaka, Y. Selective Production of Superoxide Ions and Hydrogen Peroxide over Nitrogen- and Sulfur-Doped TiO2 Photocatalysts with Visible Light in Aqueous Suspension Systems. J. Phys. Chem. C 2008, 112, 15818–15823. [Google Scholar] [CrossRef]
  92. Tsukamoto, D.; Shiro, A.; Shiraishi, Y.; Sugano, Y.; Ichikawa, S.; Tanaka, S.; Hirai, T. Photocatalytic H2O2 Production from Ethanol/O2 System Using TiO2 Loaded with Au–Ag Bimetallic Alloy Nanoparticles. ACS Catal. 2012, 2, 599–603. [Google Scholar] [CrossRef]
  93. Cai, R.; Kubota, Y.; Fujishima, A. Effect of copper ions on the formation of hydrogen peroxide from photocatalytic titanium dioxide particles. J. Catal. 2003, 219, 214–218. [Google Scholar] [CrossRef]
  94. Shiraishi, Y.; Kofuji, Y.; Sakamoto, H.; Tanaka, S.; Ichikawa, S.; Hirai, T. Effects of Surface Defects on Photocatalytic H2O2 Production by Mesoporous Graphitic Carbon Nitride under Visible Light Irradiation. ACS Catal. 2015, 5, 3058–3066. [Google Scholar] [CrossRef]
  95. Zhang, H.; Guo, L.-H.; Zhao, L.; Wan, B.; Yang, Y. Switching Oxygen Reduction Pathway by Exfoliating Graphitic Carbon Nitride for Enhanced Photocatalytic Phenol Degradation. J. Phys. Chem. Lett. 2015, 6, 958–963. [Google Scholar] [CrossRef]
  96. Yang, L.; Dong, G.; Jacobs, D.L.; Wang, Y.; Zang, L.; Wang, C. Two-channel photocatalytic production of H2O2 over g-C3N4 nanosheets modified with perylene imides. J. Catal. 2017, 352, 274–281. [Google Scholar] [CrossRef]
  97. Perry, S.C.; Pangotra, D.; Vieira, L.; Csepei, L.-I.; Sieber, V.; Wang, L.; Ponce de León, C.; Walsh, F.C. Electrochemical synthesis of hydrogen peroxide from water and oxygen. Nat. Rev. Chem. 2019, 3, 442–458. [Google Scholar] [CrossRef]
  98. Cui, Y.; Ding, Z.; Liu, P.; Antonietti, M.; Fu, X.; Wang, X. Metal-free activation of H2O2 by g-C3N4 under visible light irradiation for the degradation of organic pollutants. Phys. Chem. Chem. Phys. 2012, 14, 1455–1462. [Google Scholar] [CrossRef] [Green Version]
  99. Chen, J.; Wagner, P.; Tong, L.; Wallace, G.G.; Officer, D.L.; Swiegers, G.F. A Porphyrin-Doped Polymer Catalyzes Selective, Light-Assisted Water Oxidation in Seawater. Angew. Chem. Int. Ed. 2012, 51, 1907–1910. [Google Scholar] [CrossRef] [PubMed]
  100. Koppenol, W.H. Oxygen Activation by Cytochrome P450:  A Thermodynamic Analysis. J. Am. Chem. Soc. 2007, 129, 9686–9690. [Google Scholar] [CrossRef] [PubMed]
  101. David, A.A.; Robert, E.H.; Willem, H.K.; Sergei, V.L.; Merényi, G.; Neta, P.; Ruscic, B.; David, M.S.; Steenken, S.; Wardman, P. Standard electrode potentials involving radicals in aqueous solution: Inorganic radicals (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1139. [Google Scholar]
  102. Ghaly, M.Y.; Härtel, G.; Mayer, R.; Haseneder, R. Photochemical oxidation of p-chlorophenol by UV/H2O2 and photo-Fenton process. A comparative study. Waste Manag. 2001, 21, 41–47. [Google Scholar] [CrossRef]
  103. Macías-Sánchez, J.; Hinojosa-Reyes, L.; Guzmán-Mar, J.L.; Peralta-Hernández, J.M.; Hernández-Ramírez, A. Performance of the photo-Fenton process in the degradation of a model azo dye mixture. Photochem. Photobiol. Sci. 2011, 10, 332–337. [Google Scholar] [CrossRef]
  104. Pérez-Moya, M.; Graells, M.; Castells, G.; Amigó, J.; Ortega, E.; Buhigas, G.; Pérez, L.M.; Mansilla, H.D. Characterization of the degradation performance of the sulfamethazine antibiotic by photo-Fenton process. Water Res. 2010, 44, 2533–2540. [Google Scholar] [CrossRef]
  105. Li, S.; Dong, G.; Hailili, R.; Yang, L.; Li, Y.; Wang, F.; Zeng, Y.; Wang, C. Effective photocatalytic H2O2 production under visible light irradiation at g-C3N4 modulated by carbon vacancies. Appl. Catal. B Environ. 2016, 190, 26–35. [Google Scholar] [CrossRef] [Green Version]
  106. Zhao, S.; Guo, T.; Li, X.; Xu, T.; Yang, B.; Zhao, X. Carbon nanotubes covalent combined with graphitic carbon nitride for photocatalytic hydrogen peroxide production under visible light. Appl. Catal. B Environ. 2018, 224, 725–732. [Google Scholar] [CrossRef]
  107. Shiraishi, Y.; Kanazawa, S.; Kofuji, Y.; Sakamoto, H.; Ichikawa, S.; Tanaka, S.; Hirai, T. Sunlight-driven hydrogen peroxide production from water and molecular oxygen by metal-free photocatalysts. Angew. Chem. Int. Ed. 2014, 53, 13454–13459. [Google Scholar] [CrossRef]
  108. Kim, H.-I.; Choi, Y.; Hu, S.; Choi, W.; Kim, J.-H. Photocatalytic hydrogen peroxide production by anthraquinone-augmented polymeric carbon nitride. Appl. Catal. B Environ. 2018, 229, 121–129. [Google Scholar] [CrossRef]
  109. Shi, L.; Yang, L.; Zhou, W.; Liu, Y.; Yin, L.; Hai, X.; Song, H.; Ye, J. Photoassisted Construction of Holey Defective g-C3N4 Photocatalysts for Efficient Visible-Light-Driven H2O2 Production. Small 2018, 14, 1703142. [Google Scholar] [CrossRef] [PubMed]
  110. Zhu, Z.; Pan, H.; Murugananthan, M.; Gong, J.; Zhang, Y. Visible light-driven photocatalytically active g-C3N4 material for enhanced generation of H2O2. Appl. Catal. B Environ. 2018, 232, 19–25. [Google Scholar] [CrossRef]
  111. Qu, X.; Hu, S.; Li, P.; Li, Z.; Wang, H.; Ma, H.; Li, W. The effect of embedding N vacancies into g-C3N4 on the photocatalytic H2O2 production ability via H2 plasma treatment. Diam. Relat. Mater. 2018, 86, 159–166. [Google Scholar] [CrossRef]
  112. Wang, R.; Zhang, X.; Li, F.; Cao, D.; Pu, M.; Han, D.; Yang, J.; Xiang, X. Energy-level dependent H2O2 production on metal-free, carbon-content tunable carbon nitride photocatalysts. J. Energy Chem. 2018, 27, 343–350. [Google Scholar] [CrossRef] [Green Version]
  113. Moon, G.-H.; Kim, W.; Bokare, A.D.; Sung, N.-E.; Choi, W. Solar production of H2O2 on reduced graphene oxide–TiO2 hybrid photocatalysts consisting of earth-abundant elements only. Energy Environ. Sci. 2014, 7, 4023–4028. [Google Scholar] [CrossRef]
  114. Kim, H.; Gim, S.; Jeon, T.H.; Kim, H.; Choi, W. Distorted Carbon Nitride Structure with Substituted Benzene Moieties for Enhanced Visible Light Photocatalytic Activities. ACS Appl. Mater. Interfaces 2017, 9, 40360–40368. [Google Scholar] [CrossRef]
  115. Dong, S.; Liu, C.; Chen, Y. Boosting exciton dissociation and molecular oxygen activation by in-plane grafting nitrogen-doped carbon nanosheets to graphitic carbon nitride for enhanced photocatalytic performance. J. Colloid Interface Sci. 2019, 553, 59–70. [Google Scholar] [CrossRef]
  116. Wang, H.; Guan, Y.; Hu, S.; Pei, Y.; Ma, W.; Fan, Z. Hydrothermal Synthesis of Band Gap-Tunable Oxygen-Doped g-C3N4 with Outstanding “two-Channel” Photocatalytic H2O2 Production Ability Assisted by Dissolution-Precipitation Process. Nano 2019, 14, 1950023. [Google Scholar] [CrossRef]
  117. Bai, J.; Sun, Y.; Li, M.; Yang, L.; Li, J. The effect of phosphate modification on the photocatalytic H2O2 production ability of g-C3N4 catalyst prepared via acid-hydrothermal post-treatment. Diam. Relat. Mater. 2018, 87, 1–9. [Google Scholar] [CrossRef]
  118. Zang, L. Interfacial Donor–Acceptor Engineering of Nanofiber Materials To Achieve Photoconductivity and Applications. Acc. Chem. Res. 2015, 48, 2705–2714. [Google Scholar] [CrossRef]
  119. Kofuji, Y.; Isobe, Y.; Shiraishi, Y.; Sakamoto, H.; Tanaka, S.; Ichikawa, S.; Hirai, T. Carbon Nitride-Aromatic Diimide-Graphene Nanohybrids: Metal-Free Photocatalysts for Solar-to-Hydrogen Peroxide Energy Conversion with 0.2% Efficiency. J. Am. Chem. Soc. 2016, 138, 10019–10025. [Google Scholar] [CrossRef]
  120. Yang, Y.; Zhang, C.; Huang, D.; Zeng, G.; Huang, J.; Lai, C.; Zhou, C.; Wang, W.; Guo, H.; Xue, W.; et al. Boron nitride quantum dots decorated ultrathin porous g-C3N4: Intensified exciton dissociation and charge transfer for promoting visible-light-driven molecular oxygen activation. Appl. Catal. B Environ. 2019, 245, 87–99. [Google Scholar] [CrossRef]
  121. He, Z.; Kim, C.; Lin, L.; Jeon, T.H.; Lin, S.; Wang, X.; Choi, W. Formation of heterostructures via direct growth CN on h-BN porous nanosheets for metal-free photocatalysis. Nano Energy 2017, 42, 58–68. [Google Scholar] [CrossRef]
  122. Kofuji, Y.; Isobe, Y.; Shiraishi, Y.; Sakamoto, H.; Ichikawa, S.; Tanaka, S.; Hirai, T. Hydrogen Peroxide Production on a Carbon Nitride–Boron Nitride-Reduced Graphene Oxide Hybrid Photocatalyst under Visible Light. ChemCatChem 2018, 10, 2070–2077. [Google Scholar] [CrossRef]
  123. Wang, H.; Zhang, L.; Chen, Z.; Hu, J.; Li, S.; Wang, Z.; Liu, J.; Wang, X. Semiconductor heterojunction photocatalysts: Design, construction, and photocatalytic performances. Chem. Soc. Rev. 2014, 43, 5234–5244. [Google Scholar] [CrossRef] [PubMed]
  124. Zheng, Y.; Yu, Z.; Ou, H.; Asiri, A.M.; Chen, Y.; Wang, X. Black Phosphorus and Polymeric Carbon Nitride Heterostructure for Photoinduced Molecular Oxygen Activation. Adv. Funct. Mater. 2018, 28, 1705407. [Google Scholar] [CrossRef]
  125. Li, R.; Li, C. Chapter One—Photocatalytic Water Splitting on Semiconductor-Based Photocatalysts. In Advances in Catalysis; Song, C., Ed.; Academic Press: Cambridge, MA, USA, 2017; Volume 60, pp. 1–57. [Google Scholar]
  126. Chen, Y.-C.; Smirniotis, P. Enhancement of Photocatalytic Degradation of Phenol and Chlorophenols by Ultrasound. Ind. Eng. Chem. Res. 2002, 41, 5958–5965. [Google Scholar] [CrossRef]
  127. Qin, T.; You, Z.; Wang, H.; Shen, Q.; Zhang, F.; Yang, H. Preparation and photocatalytic behavior of carbon-nanodots/graphitic carbon nitride composite photocatalyst. J. Electrochem. Soc. 2017, 164, H211–H214. [Google Scholar] [CrossRef]
  128. Desipio, M.M.; Thorpe, R.; Saha, D. Photocatalytic Decomposition of Paraquat Under Visible Light by Carbon Nitride and Hydrogen Peroxide. Optik 2018, 172, 1047–1056. [Google Scholar] [CrossRef]
  129. Wang, X.; Li, D.; Nan, Z. Effect of N content in g-C3N4 as metal-free catalyst on H2O2 decomposition for MB degradation. Sep. Purif. Technol. 2019, 224, 152–162. [Google Scholar] [CrossRef]
  130. Dinesh, G.K.; Chakma, S. Mechanistic investigation in degradation mechanism of 5-fluorouracil using graphitic carbon nitride. Ultrason. Sonochem. 2019, 50, 311–321. [Google Scholar] [CrossRef] [PubMed]
  131. Teng, Z.; Yang, N.; Lv, H.; Wang, S.; Hu, M.; Wang, C.; Wang, D.; Wang, G. Edge-Functionalized g-C3N4 Nanosheets as a Highly Efficient Metal-free Photocatalyst for Safe Drinking Water. Chem 2019, 5, 664–680. [Google Scholar] [CrossRef] [Green Version]
  132. Saha, D.; Desipio, M.M.; Hoinkis, T.J.; Smeltz, E.J.; Thorpe, R.; Hensley, D.K.; Fischer-Drowos, S.G.; Chen, J. Influence of hydrogen peroxide in enhancing photocatalytic activity of carbon nitride under visible light: An insight into reaction intermediates. J. Environ. Chem. Eng. 2018, 6, 4927–4936. [Google Scholar] [CrossRef]
  133. Chen, Q.L.; Liu, Y.L.; Tong, L.G. Enhanced visible light photocatalytic activity of g-C3N4 assisted by hydrogen peroxide. Mater. Res. Express 2018, 5, 046203. [Google Scholar] [CrossRef]
  134. Wu, Q.; He, Y.; Zhang, H.; Feng, Z.; Wu, Y.; Wu, T. Photocatalytic selective oxidation of biomass-derived 5-hydroxymethylfurfural to 2,5-diformylfuran on metal-free g-C3N4 under visible light irradiation. Mol. Catal. 2017, 436, 10–18. [Google Scholar] [CrossRef]
  135. Jiang, L.; Yuan, X.; Zeng, G.; Wu, Z.; Liang, J.; Chen, X.; Leng, L.; Wang, H.; Wang, H. Metal-free efficient photocatalyst for stable visible-light photocatalytic degradation of refractory pollutant. Appl. Catal. B Environ. 2018, 221, 715–725. [Google Scholar] [CrossRef]
  136. Wang, H.; Su, Y.; Zhao, H.; Yu, H.; Chen, S.; Zhang, Y.; Quan, X. Photocatalytic oxidation of aqueous ammonia using atomic single layer graphitic-C3N4. Environ. Sci. Technol. 2014, 48, 11984–11990. [Google Scholar] [CrossRef]
  137. Zhang, Q.; Chen, P.; Tan, C.; Chen, T.; Zhuo, M.; Xie, Z.; Wang, F.; Liu, H.; Cai, Z.; Liu, G.; et al. A photocatalytic degradation strategy of PPCPs by a heptazine-based CN organic polymer (OCN) under visible light. Environ. Sci. Nano 2018, 5, 2325–2336. [Google Scholar] [CrossRef]
  138. Wang, H.; Guo, H.; Zhang, N.; Chen, Z.; Hu, B.; Wang, X. Enhanced Photoreduction of U(VI) on C3N4 by Cr(VI) and Bisphenol A: ESR, XPS, and EXAFS Investigation. Environ. Sci. Technol. 2019, 53, 6454–6461. [Google Scholar] [CrossRef]
  139. Wang, Z.; Murugananthan, M.; Zhang, Y. Graphitic carbon nitride based photocatalysis for redox conversion of arsenic(III) and chromium(VI) in acid aqueous solution. Appl. Catal. B Environ. 2019, 248, 349–356. [Google Scholar] [CrossRef]
  140. Wang, H.; Li, Q.; Zhang, S.; Chen, Z.; Wang, W.; Zhao, G.; Zhuang, L.; Hu, B.; Wang, X. Visible-light-driven N2-g-C3N4 as a highly stable and efficient photocatalyst for bisphenol A and Cr(VI) removal in binary systems. Catal. Today 2019, 335, 110–116. [Google Scholar] [CrossRef]
  141. Chen, J.; Xu, X.; Li, T.; Pandiselvi, K.; Wang, J. Toward High Performance 2D/2D Hybrid Photocatalyst by Electrostatic Assembly of Rationally Modified Carbon Nitride on Reduced Graphene Oxide. Sci. Rep. 2016, 6, 37318. [Google Scholar] [CrossRef] [Green Version]
  142. Svoboda, L.; Praus, P.; Lima, M.J.; Sampaio, M.J.; Matýsek, D.; Ritz, M.; Dvorský, R.; Faria, J.L.; Silva, C.G. Graphitic carbon nitride nanosheets as highly efficient photocatalysts for phenol degradation under high-power visible LED irradiation. Mater. Res. Bull. 2018, 100, 322–332. [Google Scholar] [CrossRef]
  143. Yazici, E.; Deveci, H. Factors Affecting Decomposition of Hydrogen Peroxide. In Proceedings of the XIIth International Mineral Processing Symposium, Cappadocia-Nevşehir, Turkey, 6 October 2010; pp. 609–616. [Google Scholar]
  144. Cheng, F.; Wang, H.; Dong, X. The amphoteric properties of g-C3N4 nanosheets and fabrication of their relevant heterostructure photocatalysts by an electrostatic re-assembly route. Chem. Commun. 2015, 51, 7176–7179. [Google Scholar] [CrossRef] [PubMed]
  145. Bicheng, Z.; Xia, P.; Ho, W.; Yu, J. Isoelectric point and adsorption activity of porous g-C3N4. Appl. Surf. Sci. 2015, 344, 188–195. [Google Scholar]
  146. Torres-Pinto, A.; Sampaio, M.J.; Silva, C.G.; Faria, J.L.; Silva, A.M.T. Metal-free carbon nitride photocatalysis with in-situ hydrogen peroxide generation for the degradation of aromatic compounds. Appl. Catal. B Environ. 2019, 252, 128–137. [Google Scholar] [CrossRef]
  147. Hu, J.-Y.; Tian, K.; Jiang, H. Improvement of phenol photodegradation efficiency by a combined g-C3N4 /Fe(III)/persulfate system. Chemosphere 2016, 148, 34–40. [Google Scholar] [CrossRef]
  148. Yuan, X.; Xie, R.; Zhang, Q.; Sun, L.; Long, X.; Xia, D. Oxygen functionalized graphitic carbon nitride as an efficient metal-free ozonation catalyst for atrazine removal: Performance and mechanism. Sep. Purif. Technol. 2019, 211, 823–831. [Google Scholar] [CrossRef]
  149. Li, X.; Pi, Y.; Wu, L.; Xia, Q.; Wu, J.; Li, Z.; Xiao, J. Facilitation of the visible light-induced Fenton-like excitation of H2O2 via heterojunction of g-C3N4 /NH2-Iron terephthalate metal-organic framework for MB degradation. Appl. Catal. B Environ. 2017, 202, 653–663. [Google Scholar] [CrossRef]
  150. Ren, S.; Chen, C.; Zhou, Y.; Dong, Q.; Ding, H. The α-Fe2O3/g-C3N4 composite as an efficient heterogeneous catalyst with combined Fenton and photocatalytic effects. Res. Chem. Intermediat. 2017, 43, 3307–3323. [Google Scholar] [CrossRef]
  151. Desipio, M.M.; Van Bramer, S.E.; Thorpe, R.; Saha, D. Photocatalytic and photo-fenton activity of iron oxide-doped carbon nitride in 3D printed and LED driven photon concentrator. J. Hazard. Mater. 2019, 376, 178–187. [Google Scholar] [CrossRef] [PubMed]
  152. Ding, Q.; Lam, F.L.Y.; Hu, X. Complete degradation of ciprofloxacin over g-C3N4-iron oxide composite via heterogeneous dark Fenton reaction. J. Environ. Manag. 2019, 244, 23–32. [Google Scholar] [CrossRef] [PubMed]
  153. Pedrosa, M.; Sampaio, M.J.; Horvat, T.; Nunes, O.C.; Dražić, G.; Rodrigues, A.E.; Figueiredo, J.L.; Silva, C.G.; Silva, A.M.T.; Faria, J.L. Visible-light-induced self-cleaning functional fabrics using graphene oxide/carbon nitride materials. Appl. Surf. Sci. 2019, 497, 143757. [Google Scholar] [CrossRef]
  154. Moreira, N.F.F.; Sampaio, M.J.; Ribeiro, A.R.; Silva, C.G.; Faria, J.L.; Silva, A.M.T. Metal-free g-C3N4 photocatalysis of organic micropollutants in urban wastewater under visible light. Appl. Catal. B Environ. 2019, 248, 184–192. [Google Scholar] [CrossRef]
  155. Wang, Y.; Li, H.; Yao, J.; Wang, X.; Antonietti, M. Synthesis of boron doped polymeric carbon nitride solids and their use as metal-free catalysts for aliphatic C-H bond oxidation. Chem. Sci. 2011, 2, 446–450. [Google Scholar] [CrossRef]
  156. Min, B.H.; Ansari, M.B.; Mo, Y.H.; Park, S.E. Mesoporous carbon nitride synthesized by nanocasting with urea/formaldehyde and metal-free catalytic oxidation of cyclic olefins. Catal. Today 2013, 204, 156–163. [Google Scholar] [CrossRef]
  157. Lopes, J.C.; Sampaio, M.J.; Fernandes, R.A.; Lima, M.J.; Faria, J.L.; Silva, C.G. Outstanding response of carbon nitride photocatalysts for selective synthesis of aldehydes under UV-LED irradiation. Catal. Today 2019. [Google Scholar] [CrossRef]
  158. Zhang, J.J.; Ge, J.M.; Wang, H.H.; Wei, X.; Li, X.H.; Chen, J.S. Activating Oxygen Molecules over Carbonyl-Modified Graphitic Carbon Nitride: Merging Supramolecular Oxidation with Photocatalysis in a Metal-Free Catalyst for Oxidative Coupling of Amines into Imines. ChemCatChem 2016, 8, 3441–3445. [Google Scholar] [CrossRef]
Figure 1. Photocatalytic activation and formation of reactive oxygen species in the presence of oxygen (O2) and water (H2O). H2O2 = hydrogen peroxide; O2•– = superoxide radical; HO = hydroxyl radical; VB = valence band; CB = conduction band; e = electrons; h+ = photogenerated holes.
Figure 1. Photocatalytic activation and formation of reactive oxygen species in the presence of oxygen (O2) and water (H2O). H2O2 = hydrogen peroxide; O2•– = superoxide radical; HO = hydroxyl radical; VB = valence band; CB = conduction band; e = electrons; h+ = photogenerated holes.
Catalysts 09 00990 g001
Figure 2. Proposed mechanism for selective formation of H2O2 on the photoactivated CN surface. Reprinted with permission from reference [94]. Copyright 2015 American Chemical Society.
Figure 2. Proposed mechanism for selective formation of H2O2 on the photoactivated CN surface. Reprinted with permission from reference [94]. Copyright 2015 American Chemical Society.
Catalysts 09 00990 g002
Figure 3. Formation of carbon vacancy on CN (left) and H2O2 formation pathway (right). Adapted with permission from reference [105]. Copyright 2016 Elsevier B.V.
Figure 3. Formation of carbon vacancy on CN (left) and H2O2 formation pathway (right). Adapted with permission from reference [105]. Copyright 2016 Elsevier B.V.
Catalysts 09 00990 g003
Figure 4. Nitrogen defects on CN matrix after thermal treatment. Adapted with permission from reference [110]. Copyright 2018 Elsevier B.V.
Figure 4. Nitrogen defects on CN matrix after thermal treatment. Adapted with permission from reference [110]. Copyright 2018 Elsevier B.V.
Catalysts 09 00990 g004
Figure 5. Energy levels of carbon (C)-doped CN (x denotes de carbon doping load). Reprinted with permission from reference [112]. Copyright 2018 Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences.
Figure 5. Energy levels of carbon (C)-doped CN (x denotes de carbon doping load). Reprinted with permission from reference [112]. Copyright 2018 Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences.
Catalysts 09 00990 g005
Figure 6. Top and side view of CN matrix before (left) and after distortion by benzene doping (right). Red, blue, and yellow represent carbon, nitrogen, and hydrogen atoms, respectively. Adapted with permission from reference [114]. Copyright 2017 American Chemical Society.
Figure 6. Top and side view of CN matrix before (left) and after distortion by benzene doping (right). Red, blue, and yellow represent carbon, nitrogen, and hydrogen atoms, respectively. Adapted with permission from reference [114]. Copyright 2017 American Chemical Society.
Catalysts 09 00990 g006
Figure 7. Electronic structure of different catalysts. Reprinted with permission from reference [77]. Copyright 2016 American Chemical Society. PDI = pyromellitic diimide; BDI = biphenyl diimide.
Figure 7. Electronic structure of different catalysts. Reprinted with permission from reference [77]. Copyright 2016 American Chemical Society. PDI = pyromellitic diimide; BDI = biphenyl diimide.
Catalysts 09 00990 g007
Figure 8. Energy diagram of the CN-PDI-reduced graphene oxide (rGO)- boron nitride (BN) catalyst. Reprinted with permission from reference [122]. Copyright 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany.
Figure 8. Energy diagram of the CN-PDI-reduced graphene oxide (rGO)- boron nitride (BN) catalyst. Reprinted with permission from reference [122]. Copyright 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany.
Catalysts 09 00990 g008
Figure 9. Energy diagram for the Z-scheme structure of the CN-perylene imide (PI) photocatalyst. Adapted with permission from reference [96]. Copyright 2017 Elsevier B.V.
Figure 9. Energy diagram for the Z-scheme structure of the CN-perylene imide (PI) photocatalyst. Adapted with permission from reference [96]. Copyright 2017 Elsevier B.V.
Catalysts 09 00990 g009
Table 1. Metal-free CN-based materials, experimental conditions, and respective photocatalytic results.
Table 1. Metal-free CN-based materials, experimental conditions, and respective photocatalytic results.
Modification on CNPreparation MethodExperimental ConditionsPhotocatalytic ResultsRef.
H2O2 Generated (µmol)Production Rate (µmol gcat−1 h−1)
NoneThermal polymerization of cyanamide under N2 atmosphere9/1 (v/v) propan-2-ol/water (5 mL); 4 g L−1; 2000 W Xe lamp (λ > 420 nm); O260 µmol (24 h)125[76]
NoneThermal polymerization of cyanamide under N2 atmosphere9/1 (v/v) benzyl alcohol/water matrix (5 mL); 4 g L−1; 2000 W Xe lamp (λ > 420 nm); O2109 µmol (24 h)227[76]
NoneThermal polymerization of cyanamide under N2 atmosphere9/1 (v/v) ethanol/water (5 mL); 4 g L−1; 2000 W Xe lamp (λ > 420 nm); O230 µmol (24 h)63[76]
NoneThermal polymerization of cyanamide under N2 atmosphere9/1 (v/v) butan−1-ol/water (5 mL); 4 g L−1; 2000 W Xe lamp (λ > 420 nm); O218 µmol (24 h)38[76]
NoneThermal polymerization of cyanamide under N2 atmosphere9/1 (v/v) propan−1-ol/water (5 mL); 4 g L−1; 2000 W Xe lamp (λ > 420 nm); O26.3 µmol (24 h)13[76]
NoneThermal polymerization of cyanamide under N2 atmosphere9/1 (v/v) propan-2-ol/water (5 mL); 4 g L−1; sunlight; O2120 µmol (9 h)667[76]
NoneThermal polymerization of cyanamide under N2 atmosphere9/1 (v/v) propan-2-ol/water (5 mL); 4 g L−1; sunlight (λ > 420 nm); O270 µmol (9 h)389[76]
NoneThermal polymerization of melamine9/1 (v/v) propan-2-ol/water (30 mL); 1.7 g L−1; 2000 W Xe lamp (λ > 420 nm); O2148 µmol (6 h)493[107]
Adding surface defectsSilica-templated thermal polymerization of cyanamide under N2 atmosphere9/1 (v/v) ethanol/water (5 mL); 4 g L−1; 2000 W Xe lamp (λ > 420 nm); O290 µmol (24 h)188[94]
Adding C vacanciesThermal polymerization of melamine and further treatment under Ar atmospherewater (100 mL); 1.0 g L−1; 300 W Xe lamp (λ > 420 nm); O29 µmol (1 h)90[105]
Adding N vacanciesThermal polymerization of melamine and further treatment under H2 atmospherewater (100 mL); 1.0 g L−1; 300 W Xe lamp (λ > 420 nm); O21.5 µmol (1 h)15[105]
Adding N vacanciesThermal polymerization of dicyandiamide and photo-assisted post-treatment with hydrazine20% (v) propan-2-ol/water (60 mL); 0.83 g L−1; solar simulator (λ > 420 nm); O212.1 µmol (2.5 h)97[109]
Adding N vacanciesThermal polymerization of melamine and further calcinated with sodium borohydridewater (100 mL); 1.0 g L−1; 300 W Xe lamp (λ ≥ 420 nm); O230.0 µmol (1 h)300[110]
Adding N vacanciesThermal polymerization of melamine and further calcinated with sodium borohydridewater (100 mL); 1.0 g L−1; 300 W Xe lamp (λ ≥ 400 nm); air17.0 µmol (1 h)170[110]
Adding N vacanciesThermal polymerization of melamine and further calcinated with sodium borohydridewater (100 mL); 1.0 g L−1; 300 W Xe lamp (λ ≥ 400 nm); N22.5 µmol (1 h)25[110]
Adding N vacanciesThermal polymerization of melamine and H2 plasma treatment50% (v) ethanol/water (200 mL); 1.0 g L−1; 250 W high-pressure sodium lamp (λ > 400 nm); O226000 µmol (12 h)2167[111]
C dopingThermal polymerization of melamine and sonication with glucose5/95 (v/v) propan-2-ol/water; 1.0 g L−1; 300 W Xe lamp; O238.1 µmol (4 h)318[112]
Carbon nanotubes combinationThermal polymerization of dicyandiamide and ammonium chloride and mixed with carbon nanotubes5/95 (v/v) formic acid/water (100 mL); 1.0 g L−1; 300 W Xe lamp (λ ≥ 400 nm); O248.7 µmol (1 h)487[106]
Carbon nanotubes combinationThermal polymerization of dicyandiamide and ammonium chloride and mixed with carbon nanotubes5/95 (v/v) methanol/water (100 mL); 1.0 g L−1; 300 W Xe lamp (λ ≥ 400 nm); O223.1 µmol (1 h)231[106]
Carbon nanotubes combinationThermal polymerization of dicyandiamide and ammonium chloride and mixed with carbon nanotubeswater (100 mL); 1.0 g L−1; 300 W Xe lamp (λ ≥ 400 nm); O21.3 µmol (1 h)13[106]
AQ-COOH couplingThermal polymerization of melamine and sonication with anthraquinone (AQ)-2-carboxylic acid1/9 (v/v) propan-2-ol/water; 0.5 g L−1; 150 W Xe lamp (λ > 400 nm); O2no data361[108]
AQ-COOH couplingThermal polymerization of melamine and sonication with anthraquinone-2-carboxylic acid1/9 (v/v) propan-2-ol/water; 0.5 g L−1; 150 W Xe lamp (λ > 400 nm); airno data270[108]
AQ-NH2 couplingThermal polymerization of melamine and sonication with 2-aminoanthraquinone1/9 (v/v) propan-2-ol/water; 0.5 g L−1; 150 W Xe lamp (λ > 400 nm); O2no data233[108]
AQ-SO3- couplingThermal polymerization of melamine and sonication with sodium anthraquinone-2-sulfonate1/9 (v/v) propan-2-ol/water; 0.5 g L−1; 150 W Xe lamp (λ > 400 nm); O2no data131[108]
AQ-OH couplingThermal polymerization of melamine and sonication with 2-hydroxymethylanthraquinone1/9 (v/v) propan-2-ol/water; 0.5 g L−1; 150 W Xe lamp (λ > 400 nm); O2no data86.9[108]
AQ-COOH couplingThermal polymerization of melamine and sonication with anthraquinone-2-carboxylic acidwater; 0.5 g L−1; 150 W Xe lamp (λ > 400 nm); O2no data24[108]
Benzene substitutionThermal polymerization of urea with trimesic acid1/9 (v/v) ethanol/water (30 mL); 0.5 g L−1; 300 W Xe lamp (λ > 420 nm); O2275 µmol (3 h)300[114]
N dopingThermal polymerization of melamine, sonication with tetracycline hydrochloride and further thermal exfoliation3/7 (v/v) propan-2-ol/water (100 mL); 0.5 g L−1; solar simulator (λ > 420 nm); O214 µmol (1 h)279[115]
O dopingThermal polymerization of dicyandiamide with nitric acid and hydrothermal post-treatmentwater (50 mL); 1.0 g L−1; 250 W high-pressure sodium lamp (λ > 400 nm); O2760 µmol (6 h)633[116]
Phosphate dopingThermal polymerization of melamine and hydrothermal treatment with H3PO42.6 mM EDTA aqueous solution (200 mL); 1.0 g L−1; 250 W high-pressure sodium lamp (λ > 400 nm); O21080 µmol (6 h)900[117]
CN-C60Thermal polymerization of melamine and C601/9 (v/v) propan-2-ol/water; 0.5 g L−1; 150 W Xe lamp (λ > 400 nm); O2no data63.2[108]
CN-GOThermal polymerization of melamine and sonication with GO1/9 (v/v) propan-2-ol/water; 0.5 g L−1; 150 W Xe lamp (λ > 400 nm); O2no data62.3[108]
CN-rGO Thermal polymerization of melamine and sonication with hydrazine-reduced GO1/9 (v/v) propan-2-ol/water; 0.5 g L−1; 150 W Xe lamp (λ > 400 nm); O2no data74.3[108]
CN-PDIThermal polymerization of melamine and pyromellitic dianhydride (PMDA)water (30 mL); 1.7 g L−1; 2000 W Xe lamp (λ > 420 nm); O250.6 µmol (48 h)21[107]
CN-PDIThermal polymerization of melamine and pyromellitic dianhydride (PMDA)9/1 (v/v) propan-2-ol/water (30 mL); 1.7 g L−1; 2000 W Xe lamp (λ > 420 nm); O2210 µmol (6 h)700[107]
CN-BDIThermal polymerization of melamine and biphenyl tetracarboxylic dianhydride (BTCDA)9/1 (v/v) propan-2-ol/water (30 mL); 3.3 g L−1; solar simulator (λ > 400-500 nm); O222.2 µmol (2 h)111[77]
CN-BDIThermal polymerization of melamine and biphenyl tetracarboxylic dianhydride (BTCDA)water (30 mL); 1.7 g L−1; solar simulator (λ > 420 nm); O211.6 µmol (24 h)10[77]
CN-MTIThermal polymerization of melem and mellitic acid trianhydride (MTA)water (30 mL); 1.7 g L−1; Xe lamp (λ > 420 nm); O227.5 µmol (24 h)23[78]
CN-PDI-rGOHydrothermal treatment of melem and GO and thermal polymerization with PMDAwater (30 mL); 1.7 g L−1; 2000 W Xe lamp (λ > 420 nm); O260 µmol (48 h)25[119]
CN-PDI-rGOHydrothermal treatment of melem and GO and thermal polymerization with PMDA9/1 (v/v) propan-2-ol/water (30 mL); 1.7 g L−1; 2000 W Xe lamp (λ > 420 nm); O2550 µmol (9 h)1222[119]
CN-PDI-BNSonication of melem and CN and thermal polymerization with PMDAwater (30 mL); 1.7 g L−1; 200 W Xe lamp (λ > 420 nm); O228 µmol (24 h)23[122]
CN-PDI-BNSonication of melem and CN and thermal polymerization with PMDA9/1 (v/v) propan-2-ol/water (30 mL); 1.7 g L−1; 200 W Xe lamp (λ > 420 nm); O2370 µmol (6 h)1233[122]
CN-PDI-rGO-BNSonication of melem, GO and CN and thermal polymerization with PMDAwater (30 mL); 1.7 g L−1; 200 W Xe lamp (λ > 420 nm); O237 µmol (24 h)31[122]
CN-PDI-rGO-BNSonication of melem, GO and CN and thermal polymerization with PMDA9/1 (v/v) propan-2-ol/water (30 mL); 1.7 g L−1; 200 W Xe lamp (λ > 420 nm); O2550 µmol (6 h)1833[122]
CN-PIThermal polymerization of melamine and cyanuric acid and reflux condensation reaction with perylene tetracarboxylic dianyhdride (PTCDA) and imidazolewater (30 mL); 1.7 g L−1; 300 W Xe lamp (λ > 420 nm); no data120 µmol (2 h)1200[96]
CN-BPThermal polymerization of urea followed by sonication with N-methyl-2-pyrrolidone and BP1/9 (v/v) propan-2-ol/water (30 mL); 1.7 g L−1; 300 W Xe lamp (λ > 420 nm); O2540 µmol (1 h)540[124]
CN-BNHydrothermal treatment and thermal polymerization of thiourea and melamine and sonication with BN dots1/9 (v/v) propan-2-ol/water (50 mL); 1.0 g L−1; 300 W Xe lamp (λ > 420 nm); O272.3 µmol (1 h)72.3[120]
CN-BNThermal polymerization of urea and BN nanosheets1/9 (v/v) methanol/water (40 mL); 0.5 g L−1; 300 W Xe lamp (λ > 305 nm); O2112 µmol (4 h)1400[121]
Table 2. Solar-to-chemical conversion (SCC) and apparent quantum yield (AQY) efficiencies for some studies. Nv = nitrogen vacancies.
Table 2. Solar-to-chemical conversion (SCC) and apparent quantum yield (AQY) efficiencies for some studies. Nv = nitrogen vacancies.
MaterialAQY/%SCC/%Ref.
CN-PDI2.60.10[77]
CN-PI3.2no data[96]
CN-Nv4.30.26[110]
CN-BDI4.60.13[77]
CN-PDI-BN4.80.19[122]
CN-MTI6.00.18[78]
CN-PDI-rGO6.10.20[119]
CN-PDI-BN-rGO7.30.27[122]

Share and Cite

MDPI and ACS Style

Torres-Pinto, A.; Sampaio, M.J.; Silva, C.G.; Faria, J.L.; M. T. Silva, A. Recent Strategies for Hydrogen Peroxide Production by Metal-Free Carbon Nitride Photocatalysts. Catalysts 2019, 9, 990. https://doi.org/10.3390/catal9120990

AMA Style

Torres-Pinto A, Sampaio MJ, Silva CG, Faria JL, M. T. Silva A. Recent Strategies for Hydrogen Peroxide Production by Metal-Free Carbon Nitride Photocatalysts. Catalysts. 2019; 9(12):990. https://doi.org/10.3390/catal9120990

Chicago/Turabian Style

Torres-Pinto, André, Maria J. Sampaio, Cláudia G. Silva, Joaquim L. Faria, and Adrián M. T. Silva. 2019. "Recent Strategies for Hydrogen Peroxide Production by Metal-Free Carbon Nitride Photocatalysts" Catalysts 9, no. 12: 990. https://doi.org/10.3390/catal9120990

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