Home / Publications / Heterostructures based on g-C3N4 for the photocatalytic CO2 reduction

Heterostructures based on g-C3N4 for the photocatalytic CO2 reduction

Share
Cite this
GOST
Cite
GOST copy
Roman F. Alekseev et al. Heterostructures based on g-C3N4 for the photocatalytic CO2 reduction // Russian Chemical Reviews. 2024. Vol. 93. No. 5. RCR5124
GOST all authors (up to 50) copy
Roman F. Alekseev, Andrey A. Saraev, Anna Yu. Kurenkova, Ekaterina A. Kozlova Heterostructures based on g-C3N4 for the photocatalytic CO2 reduction // Russian Chemical Reviews. 2024. Vol. 93. No. 5. RCR5124
 | 
RIS
Cite
RIS copy
TY - GENERIC
DO - 10.59761/RCR5124
UR - https://rcr.colab.ws/publications/10.59761/RCR5124
TI - Heterostructures based on g-C3N4 for the photocatalytic CO2 reduction
T2 - Russian Chemical Reviews
PB - Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii
AU - Alekseev, Roman F.
AU - Saraev, Andrey A.
AU - Kurenkova, Anna Yu.
AU - Kozlova, Ekaterina A.
PY - 2024
SP - RCR5124
IS - 5
VL - 93
ER -
 | 
BibTeX
Cite
BibTeX copy
@misc{2024_Alekseev,
author = {Roman F. Alekseev and Andrey A. Saraev and Anna Yu. Kurenkova and Ekaterina A. Kozlova},
title = {Heterostructures based on g-C3N4 for the photocatalytic CO2 reduction},
month = {jun},
year = {2024}
}
 | 
MLA
Cite
MLA copy
Alekseev, Roman F., et al. “Heterostructures based on g-C3N4 for the photocatalytic CO2 reduction.” Russian Chemical Reviews, vol. 93, no. 5, Jun. 2024, p. RCR5124. https://rcr.colab.ws/publications/10.59761/RCR5124.
Publication views
102

Keywords

photocatalysis
carbon dioxide reduction
charge transfer mechanism
graphitic carbon nitride
heterogeneous catalysis
heterostructures
photocatalyst
solar energy

Abstract

The interest of the global scientific community in the problems of CO2 utilization and returning to the carbon cycle has markedly increased in recent years. Among various CO2 transformation processes, photocatalytic reduction is one of the most promising. Currently, much attention is paid to photocatalysts based on graphitic carbon nitride, since the use of g-C3N4 makes it possible to perform CO2 reduction under visible or solar light irradiation. To increase the reduction efficiency, g-C3N4 is subjected to various modifications with the most popular and promising approach being the synthesis of composite photocatalysts based on g-C3N4 with other semiconductors to form heterostructures. Depending on the type of semiconductor, transfer of photogenerated charge carriers in these systems can occur by various mechanisms, which largely determine the course of the process and the rates of formation of reaction products. This review addresses studies on the synthesis of composite photocatalysts based on g-C3N4, with emphasis being placed on the mechanisms of charge carrier transfer and the distribution of products of CO2 reduction.

The bibliography includes 235 references.

1. Introduction

A primary task of rational management and protection of the environment is to reduce the concentration and emissions of greenhouse gases. The development of methods for reducing concentrations of greenhouse gases is the subject of many studies[1-3], with much attention being paid to utilization of carbon dioxide, the major component of greenhouse gases[4-8]. Today, there are several industrial processes for converting СО2 to valuable products such as urea, salicylic acid, ethylene carbonate and methanol. Due to the high thermodynamic stability of the CO2 molecule, traditional processes of CO2 conversion to products are carried out at high temperature and high pressure, e.g., synthesis of urea from СО2 and NH3 takes plates at a temperature of 185 °С and a pressure of 150 bar[9-11], which substantially increases the energy expenditure and decreases the economic attractiveness of these processes, despite the readily available feedstock.

For this reason, of particular interest are the ways to decrease the energy consumption of CO2 conversion processes. One of the approaches to address this task is the photocatalytic reduction of CO2, which can occur under ambient conditions[7, 12]. This process is based on the use of renewable resources, solar light and water, which makes photocatalytic reduction a promising method for CO2 utilization, but its industrial implementation is primarily limited by the lack of efficient photocatalysts. Nevertheless, photocatalytic reduction of CO2 complies with the principles of sustainable development and, in the future, it may become the most facile and inexpensive way for decreasing the CO2 concentration in the atmosphere[12, 13]. Moreover, this process gives organic compounds such as CH4, CH3OH and HCOOH [equations (1),(2),(3),(4), (5)], standard electrode potentials vs. NHE at pH = 7 are given], which can be used as synthetic fuel and in chemical industy[12, 14-17].

\[ \begin{equation} \mathrm{CO}_2+2 \mathrm{H}^{+}+2 \mathrm{e}^{-} \longrightarrow \mathrm{HCOOH}, E^0=-0.66 \mathrm{~V} \end{equation} \]
(1)

\[ \begin{equation} \mathrm{CO}_2+2 \mathrm{H}^{+}+2 \mathrm{e}^{-} \longrightarrow \mathrm{CO}+\mathrm{H}_2 \mathrm{O}, E^0=-0.52 \mathrm{~V} \end{equation} \]
(2)

\[ \begin{equation} \mathrm{CO}_2+4 \mathrm{H}^{+}+4 \mathrm{e}^{-} \longrightarrow \mathrm{CH}_2 \mathrm{O}+\mathrm{H}_2 \mathrm{O}, E^0=-0.49 \mathrm{~V} \end{equation} \]
(3)

\[ \begin{equation} \mathrm{CO}_2+6 \mathrm{H}^{+}+6 \mathrm{e}^{-} \longrightarrow \mathrm{CH}_3 \mathrm{OH}+\mathrm{H}_2 \mathrm{O}, E^0=-0.40 \mathrm{~V} \end{equation} \]
(4)

\[ \begin{equation} \mathrm{CO}_2+8 \mathrm{H}^{+}+8 \mathrm{e}^{-} \longrightarrow \mathrm{CH}_4+2 \mathrm{H}_2 \mathrm{O}, E^0=-0.25 \mathrm{~V} \end{equation} \]
(5)

In recent years, the attention of researchers has been attracted to a new polymer semiconductor, graphitic carbon nitride g-C3N4 (Figure 1), materials based on which can be used in various photocatalytic reactions[18, 19]. It is found that g-C3N4 has a planar structure based on heptazine or triazine units[20], in which carbon and nitrogen atoms are sp2-hybridized. This promising semiconductor has high chemical and thermal stability; in addition, g-C3N4 can be obtained from readily available precursors by simple methods[21, 22]. As compared with many semiconductors traditionally used in photocatalysis, for example TiO2, g-C3N4 has a rather narrow band gap (2.7 eV); in combination with the highly negative position of the conduction band (–1.3 V vs. NHE), this results in light absorption over a broad range of wavelengths and a high reduction potential of photogenerated electrons[23, 24]. In addition, g-C3N4 has a number of properties inherent in two-dimensional materials such as high mobility of charge carriers, high surface area to volume ratio and the presence of quantum size effect[25, 26]. The drawbacks of g-C3N4 include fast recombination of photoinduced electron – hole pairs[27] and low adsorption ability, which hampers any heterogeneous catalytic process[28, 29]. A popular and efficient approach to overcome these difficulties is to form heterostructures with other semiconductors or materials, which would enhance the efficiency of charge carrier separation and the adsorption properties.

Figure 1
Number of publications dealing with \( \text{g-C}_{3}\text{N}_{4} \)-based photocatalysts in recent years according to the Scopus database.

It is known that photocatalysts are activated upon absorption of light quanta with energy equal to or exceeding the size of the band gap energy of the semiconductor. As this takes place, the electron migrates from the valence band to the conduction band, and an electron vacancy (hole) appears in the conduction band. Then the photogenerated electrons and holes can either migrate to the photocatalyst surface and be involved in redox reactions with adsorbed reagents or recombine in the photocatalyst bulk to release heat. The recombination of electron – hole pairs sharply decreases the efficiency of the photocatalytic reaction; therefore, an important task of photocatalysis is separation of the photogenerated charge carriers to increase their lifetime.

On contact of two semiconductors, the difference between the Fermi levels induces the electron transfer, giving rise to a built-in electric field, which, in turn, promotes the transfer of photoinduced charge carriers[30-32]. The charge carrier separation in composite structures can proceed by different mechanisms. The classical schemes of charge carrier transfer between two semiconductors include type I and II heterojunctions. Type I heterojunction appears when the conduction band (CB) of one semiconductor (SC 2 in Figure 2а) is located lower, while its valence band (VB) is located higher in energy than these bands of another semiconductor (SC 1 in Figure 2а). In such a structure, electrons and holes would migrate from the wide-band-gap semiconductor to the narrow-band-gap one and be accumulated in one semiconductor and subsequently recombine. In type II heterojunction, both the valence and conduction bands of one semiconductor are higher in energy than those of another semiconductor, and the transfer of electrons from a semiconductor with higher conduction band level (SC 1 in Figure 2b) to a semiconductor with lower conduction band level (SC 2 in Figure 2b) is formed, while holes migrate from lower valence band in SC 2 to higher valence band in SC 1 (Figure 2b). In this case, electrons and holes are accumulated in different semiconductors, and their lifetime is thus increased.

Figure 2
Schematic image of the electron transfer for heterojunctions of type I (a) and type II (b), direct S-scheme (c), direct Z-scheme (b) and indirect Z-scheme (c).

As a rule, charge carrier separation mechanisms in composite photocatalysts composed of g-C3N4 and another semiconductor are based on type II heterojunction or direct Z-scheme. Later, S-scheme heterojunction was proposed[33]. The direct Z-scheme transfer mechanism occurs between two semiconductors the energy structure of which is similar to that described above for type II heterojunction. However, in this case, an electron with a low reduction potential (in SC 2) recombines with a hole that has a low oxidation potential (in SC 1); in comparison with type II heterojunction, this gives rise to charge carriers with a higher redox potential. The separation of charge carriers between two semiconductors according to the S-scheme is similar to the direct Z-scheme; however, the S-scheme takes into account the semiconductor band bending caused by alignment of their Fermi levels and generation of the built-in electric field in the photocatalyst[34]. As a rule, Z-scheme is implemented by two n-type and p-type semiconductors, while S-scheme is implemented by n-type semiconductors, considered as an oxidation photocatalyst and reduction photocatalyst (see Figure 2)[3, 35, 36]. Both Z-scheme and S-scheme can be either direct (Figure 2d) or indirect (Figure 2e); in the latter case, compounds with high electrical conductivity, e.g., metals or some carbon materials such as graphene, serve as mediators[37]. It is important that implementation of either Z- or S-scheme in a composite photocatalyst not only increases the efficiency of separation of the electron – hole pairs, but also maximizes the redox capacity of charge carriers.

One more popular method for increasing the lifetime of photogenerated charge carriers is deposition of metals on the semiconductor surface. The Schottky junction is formed at the metal-semiconductor interface, and electrons from the conduction band of the semiconductor move to Fermi level of metal, which leads to alignment of the metal and semiconductor Fermi levels and bending of semiconductor bands near the interface.

Figure 3
Band structure of some reduction and oxidation photocatalysts[38].

Apart from the lifetime of photogenerated charge carriers, another important characteristic of semiconductor photocatalysts is the size of the optical band gap (below referred to as band gap), which determines the minimum energy of light quantum necessary for generation of electron – hole pairs (Figure 3)[38]. As a rule, the band gap energy is derived from the diffuse reflectance spectroscopy by the method proposed by Jan Tauc[39]. First, the absorption coefficient \( F(R) \) is found using the Kubelka – Munk equation:

\[ \begin{equation} F(R)=\frac{(1-R)^2}{2 R} \end{equation} \]
(6)

where \( R \) is the reflection factor of the sample determined from the diffuse reflectance spectroscopy data.

Using the Tauc equation, the dependence of the absorption coefficient \( F(R) \) on the photon energy is found:

\[ \begin{equation} (F(R) h \nu)^{(1/ \gamma)} = B(h \nu - E_g) \end{equation} \]
(7)

where \( h \) is the Planck constant, \( \nu \) is the photon frequency, \( E_g \) is the band gap energy, \( B \) is a constant, \( \gamma \) is the factor equal to 1/2 for direct junctions and to 2 for indirect junctions.

The intersection of the tangent to the Tauc plot with the abscissa gives the band gap energy of the semiconductor. It is noteworthy that this method is suitable for determining the band gap energy of unmodified semiconductors. Determining the band gap energy for composite materials, especially in the case of different types of electron transfer (direct/indirect) in semiconductors gives, most often, incorrect results. Nevertheless, in many publications, analysis of band gap changes based on diffuse reflectance spectroscopy is often performed for composite materials. In this case, the spectrum of the photocatalyst is usually a linear combination of the spectra of independet components, except the cases where quantum size effects or localized surface plasmon resonance influence the optical properties of materials. In some cases, the band gap of modified semiconductors can be estimated by drawing the tangent to the region called Urbach tail in the diffuse reflectance spectra. Then the intersection of this tangent with the tangent to the Tauc plot gives a correct band gap energy for a single component of a photocatalyst[40].

It is known that graphitic carbon nitride can behave as either p-type or n-type semiconductor depending on the external effects such as the applied voltage and redox potential of a reagent[41-45]. For example, g-C3N4 behaves as a p-type semiconductor when a negative voltage is applied and as an n-type semiconductor when the voltage is positive; therefore, it can be considered as an amphoteric semiconductor[42, 46, 47]. The amphoteric semiconductor properties also depend on the g-C3N4 bulk structure and surface functionalization[48]. Hence, the charge carrier transfer mechanism in composite photocatalysts composed of g-C3N4 and a semiconductor with a high potential of photogenerated holes can be considered as a Z-scheme or S-scheme, as described below.

Traditionally, g-C3N4 is obtained by heat treatment of various nitrogen-containing precursors such as melamine, dicyandiamide, and urea[49-51], but, most often, this method does not provide a material with a high specific surface area. In recent years, numerous new approaches to the synthesis of g-C3N4 have been proposed providing the possibility of varying the particle size, pore volume and specific surface area of the resulting material[52]. The methods of synthesis of g-C3N4 are comprehensively addressed in the literature[22, 53-57]; therefore, they are not discussed here.

This review is focused on g-C3N4 heterostructures with other materials used for photocatalytic CO2 reduction. It is noteworthy that there are a few published reviews dealing with modification of g-C3N4 and fabrication of heterostructures for the photocatalytic reduction of CO2[58-64] considering photocatalysts in the light of charge carrier transfer mechanisms or the effect of methods of synthesis on the photocatalytic properties. There are also reviews on using g-C3N4-based systems for other photocatalytic reactions (hydrogen evolution, decomposition of dyes, water treatment, etc.)[54, 65, 66]. A widely used photocatalyst is g-C3N4 with platinum deposited on its surface, which is obtained by photoreduction of H2PtCl6 in a solution of an electron donor, most often, triethanolamine or by reduction with a solution of NaBH4[67, 68]. Modification of semiconductors with transition metals to enhance the photocatalytic activity has also been studied in detail and presented in the literature[69-71].

This review is devoted to the most recent advances in the field of g-C3N4-based heterostructures for photocatalytic CO2 reduction. The attention is focused on the materials used and the most likely mechanism of charge carrier transfer between the semiconductors. In addition, a distinctive feature of this review is comparison of the photocatalyst activity in terms of the overall rate of consumption of photogenerated electrons, which makes it possible to compare the activities of photocatalysts with different selectivities to CO2 reduction products. The review addresses the most widely used and promising systems containing metal oxides and sulfides, two-dimensional metal carbides (MXenes) and carbon materials. Since in some cases the literature provides information about different possible heterojunctions for the same type of heterostructure, the role of the type of charge carrier separation mechanism is also analyzed in each Section.

2. g-C3N4-based heterostructures with metal oxides

2.1. Heterostructures with titanium dioxide

Titanium dioxide is one of the most popular photocatalysts used both for CO2 reduction and for other photocatalytic processes, which is due to the low cost, stability and low toxicity of TiO2[72-74]. However, the use of TiO2-based photocatalysts is limited by the large band gap (~3.2 eV), since only UV photons have sufficient energy to induce the photoexcitation in TiO2[12, 75]. As a consequence, TiO2 has low activity under the sunlight, in which the fraction of ultraviolet radiation is much smaller than that of the visible radiation. To increase the response of TiO2-based photocatalysts to visible light, TiO2 is modified with narrow-band-gap semiconductors, for example, g-C3N4[76].

Titanium dioxide is known to have three stable crystalline phases — anatase, rutile, and brookite (Figure 4)[21, 77-79]. As a rule, the first two of them are used in photocatalytic studies; anatase is considered to be more active than rutile, because of the higher reduction potential of electrons and the ability to form hole traps[80, 81], which is attributable to different predominant orientations of the crystallite surfaces in rutile and anatase. In the case of anatase, the crystallite surface is enriched with {101} and {001} faces, while in the case of rutile, {110}, {100} and {101} faces predominate[82]. It is known that the {001} face is more reactive in photocatalytic reactions[83, 84]. Meanwhile, rutile has a narrower band gap than anatase and, hence, it can be used for reactions under irradiation at longer wavelengths[85]. The conduction band minimum in TiO2 is approximately –0.2 V (vs. NHE at pH = 7)[21], which is insufficient for CO2 reduction, since the formation potential of most products is more negative [see Eqns (1) – (5)]. However, the conduction band minimum in g-C3N4 is approximately –1.20 V (vs NHE at pH = 7), which is sufficient for the formation of various organic compounds from CO2[21].

Figure 4
Crystal structure of \( \text{TiO}_2 \) phases: (a) anatase, (b) rutile, (c) brookite[79]. Published with permission from the Royal Society of Chemistry.

In the formation of g-C3N4-based composite photocatalysts, the particles of other components are usually distributed over the g-C3N4 surface, because it has a layered structure where the longitudinal particle size substantially exceeds the transverse size. These composite structures are commonly designated by X/g-C3N4 (X is a component of a g-C3N4-based photocatalyst), e.g., TiO2/g-C3N4. This notation is used in this review, although authors of original publications may use other designations for heterostructures.

A fairly widespread method for the synthesis of TiO2/g-C3N4 composites is physical mixing of components, according to which the components (TiO2 and g-C3N4) are synthesized separately[19, 86]. TiO2 is obtained, most often, by hydrolysis, hydrothermal or solvothermal treatment of titanium-containing precursors such as titanium alkoxides[19, 73].

Mehregan et al.[87] proposed a method for the fabrication of TiO2/g-C3N4 composite by hydrothermal treatment [hereinafter, slash (/) is used in the designations of heterostructures, while a hyphen (–) designates semiconductors doped with metal or non-metal atoms]. Titanium dioxide was prepared by the sol – gel method using titanium tetrabutoxide as the precursor, while g-C3N4 was obtained by melamine calcination followed by exfoliation to give a layered structure (Figure 5). A suspension consisting of weighed portions of TiO2 and g-C3N4 obtained in this way (in 2 : 1 w/w ratio of TiO2 to g-C3N4) and aqueous ethanol was sonicated and then placed in an autoclave and kept at 120 °C for 3 h. The CO2 reduction was carried out in the gas phase in the presence of water vapour on exposure to visible light. The authors studied the effect of light intensity on the reaction rate and showed a pronounced increase in the photocatalyst activity with as the light power density increased from 20 to 80 mW cm–2. Thus, the light intensity of 80 mW cm–2 provided the highest product formation rates, 33 and 1.4 mmol g–1 h–1 for CH4 and CH3OH, respectively. In some other publications, it is also shown that increase in the light power density increases the rates of electron and hole generation, which results in a higher photocatalyst activity[88, 89].

Figure 5
Schematic diagram of the approach to the fabrication of \( \text{TiO}_{2}/ \text{g-C}_{3}\text{N}_{4} \) heterostructures.

It should be emphasized that Evonik P25, a commercial TiO2 powder, is used most often for the synthesis of various TiO2-based photocatalysts[12]. This TiO2 powder is composed of 80% anatase and 20% rutile with an average particle size of 25 nm and has a specific surface area of 40 – 60 m2 g–1[90]. Wang et al.[91] synthesized TiO2/g-C3N4 composite photocatalysts with various component ratios by ball milling of TiO2 (anatase) with g-C3N4 pretreated with a solution of HNO3 to form a layered structure (Figure 6). Then the mixture was calcined at 400 °С for 1 h to form the composite material with TiO2 particles deposited on the g-C3N4 surface. The authors demonstrated a change in the positions of the valence and conduction bands depending on the component ratio and increase in the band gap with increasing TiO2 content. This study describes the liquid-phase СО2 reduction in which the suspension with a photocatalyst in aqueous NaOH and triethanolamine solution was placed in a reactor, and then purged with СО2 flow. The highest activity after 4 h of UV irradiation (8 W lamp) was found for the photocatalyst with TiO2 to g-C3N4 ratio of 1 : 2. The rates of СО and СН4 formation were 14 and 18 mmol g–1 h–1, respectively, which was twice as high as that for pristine g-C3N4.

Figure 6
Schematic diagram of the synthesis of the \( \text{TiO}_{2}/ \text{g-C}_{3}\text{N}_{4} \) photocatalyst[91].

Truc et al.[92] synthesized composite photocatalysts based on g-C3N4 and niobium-doped TiO2 in which the direct Z-scheme heterojunction was formed (Figure 7)[92]. The introduction of Nb into TiO2 structure leads to the formation of Ti3+ ions and the appearance of an additional energy level near the TiO2 conduction band, thus reducing the band gap size from 3.2 to 2.9 eV. The change in the band gap by introducing metal and non-metal ions into a parent substance is a common method used to shift the absorption edge of photocatalysts[93-97]. The Nb – TiO2/g-C3N4 composite photocatalysts were obtained by calcination of a mixture of Nb – TiO2 with melamine. In a series of experiments on the gas-phase photocatalytic CO2 reduction on exposure to light from two 30-W white lamps, the highest activity was detected for the photocatalyst with 1 : 1 ratio of (Nb – TiO2) to g-C3N4. Apart from CO and CH4, the reaction products were found to contain formic acid and oxygen resulting from water oxidation. The formation rates of CO, CH4, HCOOH and O2 were 420, 560, 700 and 1700 mmol g–1 h–1, respectively. The results obtained in this work demonstrate the high potential of the integrated approach to the synthesis of photocatalysts, including both doping and the formation of heterojunctions in composite materials.

Figure 7
Band structure of single materials \( \text{TiO}_2 \) (a) and \( \text{g-C}_{3}\text{N}_{4} \) (b) and schematic image of the processes that take place in doped \( \text{Nb-TiO}_{2} \) (c) and composite \( \text{Nb-TiO}_{2}/ \text{g-C}_{3}\text{N}_{4} \) (d) photocatalysts[92].

To compare the activity of photocatalysts with different selectivities to reaction products, it was proposed to use the total number of electrons absorbed in the photocatalytic process per unit time and per unit mass of the photocatalyst, We, which is calculated by the equation[98]:

\[ \begin{equation} W_e = 2W_{CO} + 8W_{CH_4} + 6W_{CH_{3}OH} + 2W_{HCOOH} \end{equation} \]
(8)

where WCO, WCH4, WCH3OH , and WHCOOH are the rates of formation of the products per unit mass of the photocatalyst (as a rule, mmol g–1 h–1); coefficients are determined by the number of electrons needed for the formation of this product. Methane makes the greatest contribution to We, because its formation requires more electrons than the formation of other compounds [see Eqns (1) – (5)]. Equation (8) or similar relations are convenient for calculating the apparent quantum efficiency (AQE) [Equation (9)] and are widely used in the studies of the photocatalytic CO2 reduction, although the terminology may differ[7, 98-101].

\[ \begin{equation} AQE = \frac {W^*_{e}}{N_{ph}} \times 100 \% \end{equation} \]
(9)

where We* is the product formation rate with allowance for the electron balance (usually μmol h–1), Nph is the photon flux from the light source calculated from the data on the irradiation power and spectrum (mmol h–1).

Unfortunately, AQE values for the photocatalytic CO2 reduction (usually not exceeding 1%) are rarely reported in the literature; therefore, comparison of the photocatalyst activity will be based on calculation of the rate We[12]. However, it is noteworthy that AQE is an important characteristic of photocatalytic systems, as it reflects the photon utilization efficiency, while the rate We depends, among other factors, on the light irradiation power. The only outstanding AQE value was repotrted by Zhang et al.[102] A composite photocatalyst based on g-C3N4 and Ag-doped TiO2 has an AQE of 2.4% in the liquid-phase CO2 reduction. It is shown that the addition of Ag promotes implementation of S-scheme heterojunctions and electron transfer from TiO2 to g-C3N4. In addition, high efficiency of CO2 reduction is attained owing to high specific surface area and defects in the g-C3N4 structure.

The data on TiO2/g-C3N4 composite systems that were reported in recent years are summarized in Table 1, which also indicates the reaction conditions and reagents added to the reactor. To carry out the photocatalytic reaction, a flow of ultrapure СО2 is usually purged through the reactor; however, the СО2 generation directly in the reactor, e.g., by the reaction of NaHCO3 and H2SO4, can also be used. Most often, the source of protons in photocatalytic CO2 reduction is water, but CH4 can also serve for this purpose. Generally, TiO2 is one of the most popular semiconductors for the synthesis of photocatalysts; therefore, there are numerous heterostructures based on TiO2. A combination of TiO2 with g-C3N4 in the TiO2/g-C3N4 composite photocatalysts gives We values reaching 270 mmol g–1 h–1 under visible light irradiation[87], while modification of the TiO2/g-C3N4 composite leads to even a more pronounced increase in the activity. As can be seen from Table 1, high We value was attained for composite system comprising Z-scheme heterojunction and synthesized from TiO2 doped with metal ions; therefore, this method appears to be most efficient.

Table 1
\[ \]
Review of some publications on CO2 reduction in the presence of photocatalysts based on TiO2/g-C3N4 (GP means that СО2 reduction is carried out in the gas phase, while LP refers to liquid-phase reduction). Refs. [87, 91, 92, 102-112]
(1)

2.2. Heterostructures with zinc oxide

Zinc oxide ZnO is widely used in various fields owing to its mechanical, electrical, optical and photocatalytic properties[113, 114]. The most thermodynamically stable ZnO phase under ambient conditions is wurtzite (Figure 8)[115, 116]. ZnO has a band gap of approximately 3.4 eV and n-type conduction, which makes it similar to TiO2 for photocatalytic applications[117]. Hence, combination of ZnO with narrow-band-gap semiconductors such as g-C3N4 is also a promising approach for increasing the photocatalytic activity similarly to TiO2/g-C3N4 composites.

Figure 8
Crystal structure of wurtzite phase of ZnO[115].

A variety of methods for the preparation of ZnO/g-C3N4 composites such as hydrothermal and solvothermal synthesis, layer-by-layer deposition and other have been reported in the literature[19]. For example, Chen et al.[118] synthesized the ZnO/g-C3N4 composite photocatalyst by evaporation of a suspension consisting of a methanol solution, g-C3N4 and zinc acetate prepared in advance. The reduction involved CO2 and water vapour, which were formed inside the reactor upon the reaction between NaHCO3 and H2SO4. The We value found for the composite photocatalyst was 9.4 times higher than that for pristine g-C3N4. The use of We for comparison of the photocatalyst activities is especially important here, because the photocatalytic СО2 reduction in the presence of g-C3N4 is dominated by the formation of CO, whereas in the case of ZnO/g-C3N4 composites, the reaction almost exclusively gives СН4. The authors also studied the stability of the most active ZnO/g-C3N4 photocatalyst and demonstrated that the activity decreased by about 7% by the third cycle, which indicated a high stability of the synthesized photocatalyst.

Guo et al.[119] synthesized the ZnO@g-C3N4 composite photocatalyst with the core@shell structure by depositing g-C3N4 on porous ZnO nanosheets using two-step calcination and studied the effect of reaction temperature on the photocatalyst activity towards the CO2 reduction. A temperature rise from 150 to 200 °C resulted in a 3-fold and 2-fold increase in the rate of CH4 and CO formation, respectively. However, as the temperature was further increased to 250 °C, the formation rate of CO2 reduction products increased by only 33% compared to the rate at 200 °C. This effect may be caused by diffusion processes and adsorption – desorption equilibrium, which are affected by temperature. Since impossibility of initiation of photocatalytic processes by thermal energy has been proved in the literature both experimentally and theoretically, the change in the activity can be due only to the change in the above mentioned dark stages of the reaction[12, 120-122].

Zhu et al.[123] deposited copper nanoparticles on the ZnO/g-C3N4 materials with different contents of ZnO to be used in the photocatalytic CO2 reduction in aqueous solution under mercury lamp irradiation. The highest rates of CO, CH4 and CH3OH formation equal to 64, 41 and 93 mmol g–1 h–1, respectively, were attained by using the 3% Cu/(30% ZnO/g-C3N4) photocatalyst (The contents of the components are given in mass percent.). These values are not only markedly higher than those for pristine g-C3N4, but they are also higher than those for the 30% ZnO/g-C3N4 composite photocatalyst. The authors proposed a possible microstructure of the 3% Cu/(30% ZnO/g-C3N4) photocatalyst and a mechanism of separation of the photogenerated charge carriers (Figure 9). Presumably, the charge carrier separation mechanism implemented in the photocatalysts is similar to type II heterojunction, but occurs via copper nanoparticles, which act as not only electron traps, but also as sources of electrons, causing a significant increase in the photocatalytic reaction rate.

Figure 9
Possible mechanism of electron transfer in the \( \text{Cu/ZnO/g-C}_{3}\text{N}_{4} \) heterostructures[123].

Data on some ZnO/g-C3N4 systems are summarized in Table 2. Currently, ZnO has been less explored for photocatalytic applications than TiO2; hence, relatively few data on heterostructures based on ZnO composites with g-C3N4 have been reported in the literature. It is important that the activity of ZnO/g-C3N4 photocatalysts in the CO2 reduction is at the level of TiO2/g-C3N4 composite activity. Moreover, the We value of 1010 mmol g–1 h–1 for Cu/ZnO/g-C3N4 is especially notable, because three СО2 reduction products, CO, CH4 and CH3OH, are formed at high rates in this case. Most types of heterojunctions present in the ZnO/g-C3N4 photocatalysts reported in the literature correspond to type II. However, the highest We value was attained for a three-component composite photocatalyst with a complex of heterojunctions.

Table 2
\[ \]
Review of some publications on CO2 reduction in the presence of ZnO/g-C3N4-based photocatalysts. Refs. [118, 119, 123-127]
(2)

2.3. Heterostructures with cerium dioxide

Cerium dioxide CeO2 is an n-type semiconductor with a wide band gap (2.8 – 3.1 eV) and the fluorite structure[128, 129]. An important feature of this compound is the high proneness of Ce4+ cations to be reduced to Ce3+[130]. This change in the oxidation state results in a change in the stoichiometry and formation of oxygen vacancies, which are known to enhance the absorption of visible light and also act as photoinduced charge carrier traps or as adsorption sites (Figure 10)[131, 132].

Figure 10
Crystal structure of the \( \text{CeO}_2 \) fluorite phase[132].

Liang et al.[133] investigated hollow g-C3N4@CeO2 photocatalysts for CO2 reduction under irradiation with a xenon lamp with a 420 nm cut-off filter (l > 420 nm). The heat treatment of the composite photocatalyst in a hydrogen atmosphere resulted in the formation of oxygen vacancies and partial Ce4+ reduction to Ce3+. The highest rates of CH4, CH3OH and CO formation were attained in the presence of the g-C3N4@49.7%CeO2 photocatalyst and amounted to 1.2, 1.7 and 5.6 mmol g–1 h–1, respectively, which exceeds these values for single g-C3N4 and CeO2 catalysts. The synergistic effect is caused by the formation of type II heterojunction between the two semiconductors and by the large number of oxygen vacancies in CeO2. The Ce4+ cations can trap photoinduced electrons, whereas the Ce3+ cations apparently provide the formation of the CO2 – radical anion, which is an intermediate of CO2 reduction[134]. Moreover, the hollow structure of the photoсatalyst increases the light utilization efficiency due to multiple reflections[135].

Wang et al.[136] synthesized a series of CeO2/g-C3N4 composite photocatalysts with a built-in electric field. The electric field formation was confirmed by density functional theory (DFT) calculations. The calculation results indicate that electrons are accumulated on the g-C3N4 surface, while holes are concentrated on CeO2. This gives rise to a built-in electric field directed from CeO2 to g-C3N4, which promotes the transfer of photogenerated charge carriers according to the S-scheme. The highest product formation rates attained on the 1.75% CeO2/g-C3N4 photocatalyst were 0.56 and 15 mmol g–1 h–1 for CO and CH4, respectively, which is almost 20 times higher than those attained with single g-C3N4 or CeO2.

Li et al.[137] reported a multistep hydrothermal synthesis of the CeO2/g-C3N4 composite photocatalyst; then partially reduced graphene oxide (rGO) was deposited on the composite surface. Partially reduced graphene oxide is a graphene-like material with specific structural defects and oxidized groups on the surface[138]. The combination of these characteristics gives rise to a two-dimensional material with a large specific surface area, high electron mobility and chemical stability, which makes it a promising component for the synthesis of composite photocatalysts[139]. The photocatalytic CO2 reduction was carried out in a suspension consisting of an alkaline solution of TEOA and a photocatalyst under xenon lamp irradiation. The CO and CH4 formation rates were 63 and 33 mmol g–1 h–1 for rGO/CeO2/g-C3N4 vs 15 and 5.2 mmol g–1 h–1 for g-C3N4. The results of DFT calculations suggest the formation of a built-in electric field in the composite photocatalyst caused by changes in the Fermi levels, similarly to what was described by Li et al.[137]. In the resulting multicomponent heterostructure, the formation of S-scheme heterojunction for electrons was suggested.

Data on the CeO2/g-C3N4 systems are summarized in Table 3. The use of CeO2 in photocatalytic studies is primarily due to the variable stoichiometry. The highest We values for CeO2-based composite photocatalysts were obtained for heterostructures in which the charge carrier separation mechanism corresponds to the S-scheme and which are used for liquid-phase CO2 reduction. As can be seen from Table 3, the formation of S-scheme hetero-junction provides the highest activity of CeO2/g-C3N4-based photocatalysts. However, the activity of photocatalysts of this type is usually lower than that of TiO2/g-C3N4 or ZnO/g-C3N4. Apparently, CeO2 may be a promising material for the synthesis of photocatalysts for СО2 reduction; however, additional studies along this line are needed.

Table 3
\[ \]
Review of some studies on CO2 reduction in the presence of CeO2/g-C3N4-based photocatalysts. Refs. [133, 136, 137, 140-145]
(3)

2.4. Heterostructures with iron oxide α-Fe2O3

The iron oxide α-Fe2O3 (hematite) is a readily available, thermodynamically stable and environmentally benign narrow-band-gap n-type semiconductor[146-149]. However, the narrow band gap (2.2 eV) not only enhances the visible light absorption, but also markedly decreases the lifetime of photogenerated electron – hole pairs[150]. Moreover, the energy level of the bottom of the conduction band hampers the use of this material in reduction reactions due to the low potential of photogenerated electrons[19, 151]. Thus, α-Fe2O3 can be used in the photocatalytic CO2 reduction only upon the formation of heterostructures with other semiconductors. In this respect, g-C3N4 can act as a semiconductor with an appropriate level of the conduction band minimum and with wider band gap compared to that of α-Fe2O3.

Guo et al.[152] used the hydrothermal method to prepare the α-Fe2O3/g-C3N4 composite for the photocatalytic CO2 reduction to CH3OH in water under xenon lamp irradiation with a 420 nm cut-off filter. The highest CH3OH formation rate is 5.6 mmol g–1 h–1, which is almost three times higher than that for pristine g-C3N4 (1.9 mmol g–1 h–1). This increase is attributable to the formation of direct Z-scheme, which prevents recombination of photogenerated charge carriers and promotes the generation of electrons with a high reduction potential.

Duan and Mei[153] synthesized the α-Fe2O3/g-C3N4 photocatalyst by hydrothermal method from a colloidal solution containing both components (Figure 11). The CO2 reduction reaction was conducted in an aqueous solution of DMF and TEOA under irradiation with a 60 W white light emitting diode (LED). The major reaction product was CH3OH formed at a high rate of 74 mmol g–1 h–1. The solvent effect on the reaction rate was investigated by replacing DMF with acetonitrile, but the photocatalyst activity significantly decreased, indicating that the use of DMF in the photocatalytic CO2 reduction may be promising for α-Fe2O3/g-C3N4 photocatalysts (but not necessarily for other ones[154]). It is worth noting that under the light irradiation, TEOA can be converted in the reaction medium to give carbon-containing products, which can affect the photocatalytic reaction rate[155-158]. This effect was ruled out by conducting an experiment with 13CO2; this confirmed that CO2 was the only source of CH3OH.

Figure 11
Schematic image of the synthesis of the \( \alpha \text{-Fe}_{2}\text{O}_{3} /\text{g-C}_{3}\text{N}_{4} \) photocatalyst[153].

Padervand et al.[159] synthesized a complex composite photocatalyst, K4Nb6O17/Fe3N/α-Fe2O3/C3N4, by one-step thermal pyrolysis of the precursor. The photocatalyst combines the properties of three semiconductors with different positions of bands and magnetic properties of Fe3N, which produces a highly efficient system for photocatalytic reactions. The highest product formation rates in the gas-phase CO2 reduction under visible light irradiation were 7.0 and 1.3 mmol g–1 h–1 for CO and CH4, respectively. The authors also proposed an electron transfer mechanism: since K4Nb6O17 has a wider band gap with a very low top of valence band, type I heterojunction with α-Fe2O3 can be formed simultaneously with S-scheme heterojunction between α-Fe2O3 and g-C3N4. Moreover, Fe3N acts as an electron trap, similarly to metals. These features explain the high CO2 conversion upon photocatalyst irradiation with visible light.

Data on α-Fe2O3/g-C3N4 systems reported recently are summarized in Table 4. The highest We value of 440 mmol g–1 h–1 was found for the reaction in an aqueous solution of DMF + TEOA, whereas standard conditions of CO2 reduction in aqueous solutions or in the presence of water vapour resulted in a lower rate, even when complex heterostructures were used as photocatalysts. Thus, the formation of composite photocatalysts consisting of g-C3N4 and narrow-band-gap semiconductors such as α-Fe2O3 is not a promising trend. In this case, high rates of product formation are not attained, probably because of the low lifetime of photogenerated charge carriers caused by the narrow band gaps of both components.

Table 4
\[ \]
Review of some studies on CO2 reduction in the presence of α-Fe2O3/g-C3N4-based photocatalysts. Refs. [152, 153, 159-161]
(4)

3. Heterostructures with metal sulfides

3.1. Heterostructures with cadmium sulfide

Cadmium sulfide CdS is a semiconductor material widely used for photocatalytic CO2 reduction and many other photocatalytic reactions, because it has a narrow band gap (~2.4 eV) and appropriate arrangement of the valence and conduction bands[162-165]. Meanwhile, the photocatalytic application of CdS is limited not only by recombination of electron – hole pairs, but also by photocorrosion caused by the oxidation of sulfide ions by photogenerated holes[166-169]. An effective method to overcome these drawbacks is to fabricate composite photocatalysts based on CdS and other semiconductors.

Vu et al.[170] synthesized the CdS/g-C3N4 composite and studied it in the photocatalytic CO2 reduction under irradiation with a solar simulator (100 mW cm–2); the photocatalyst was suspended in a solution containing acetonitrile, TEOA, water and [Co(bpy)3]Cl2. An aqueous solution of acetonitrile was used as a solvent, TEOA served as an electron donor and the organometallic complex acted as a co-catalyst. It is noteworthy that no CO2 reduction products were detected in the absence of TEOA and the Co complex. Carbon monoxide was formed as the major product of CO2 reduction, CH4 was detected in a small amount, and H2 was formed as a by-product. The rate of СО evolution was 240 mmol g–1 h–1, which was almost four times higher than that with pristine g-C3N4. The selectivity to СО was 73%. The photocorrosion of CdS was inhibited via migration of photogenerated holes to g-C3N4: study of the photocatalyst stability indicated no loss of activity after four reaction cycles. The synergistic effect of CdS/g-C3N4 heterostructure is attributable to implementation of the direct Z-scheme with a bridging C – S – Cd bond at the interface, which results in increasing rate of transfer and separation of photogenerated charge carriers (Figure 12).

Figure 12
Mechanism of the photoinduced electron transfer in the \( \text{CdS/g-C}_{3}\text{N}_{4} \) photocatalysts in a solution with \( [\text{Co(bpy)}_{3}]\text{Cl}_{2} \)[170]. Published with permission from the American Chemical Society.

Guo et al.[171] synthesized a photocatalyst based on g-C3N4 and Zn0.2Cd0.8S nanoparticles using a combination of ultrasonic treatment and hydrothermal method (Figure 13). The ZnxCd1–xS solid solutions are known as photocatalysts that are activated under visible light and have a tunable band structure, which can be controlled by varying the Zn : Cd ratio, as in other solid solutions[172-175]. The photocatalytic activity in the CO2 reduction was measured in an aqueous suspension of the photocatalyst at 80 °C under irradiation with a xenon lamp with a 420 nm cut-off filter. Methanol was the only detected product, which formed at a rate of 12 mmol g–1 h–1; this was higher than the methanol formation rate in the reaction performed using single components of the photocatalyst. Stability testing showed a decrease in the activity by only 7% after 28 h of the reaction, which indicated successful inhibition of photocorrosion. The authors suggested that separation of the photogenerated charge carriers in this photocatalyst corresponds to type II heterojunction, which promotes increase in both stability and activity.

Figure 13
Schematic image of the synthesis of the \( \text{Zn}_{0.2}\text{Cd}_{0.8}\text{S/g-C}_{3}\text{N}_{4} \) photocatalyst[171].

Data on some CdS/g-C3N4-based systems are summarized in Table 5. Since CdS is also a narrow-band-gap semiconductor, its combination with g-C3N4 usually does not provide high rates of CO2 reduction. To attain high activity, additional modification or more complex reaction system is required. In particular, the two highest We values were found for systems using m-CdS/g-C3N4 photocatalyst dispersed in a solution, whereas We observed for the gas-phase СО2 reduction in the presence of (Au/ZnxCd1–xS)@g-C3N4 was only 7.9 mmol g–1 h–1.

Table 5
\[ \]
Review of some studies on CO2 reduction in the presence of CdS/g-C3N4-based photocatalysts. Refs. [170, 171, 176-178]
(5)

3.2. Heterostructures with tin sulfide

Tin disulfide SnS2 is a non-toxic semiconductor with a narrow band gap (~2.2 eV); therefore, it may be of interest for photocatalytic studies as a combination with wider-band-gap semiconductors[179-184].

Wang et al.[185] synthesized SnS2/g-C3N4 composite photocatalysts with type II heterojunction by using self-assembly based on the electrostatic interactions between the components in an ethanol solution. The prepared photocatalysts with different SnS2 contents were tested in the photocatalytic CO2 reduction under xenon lamp irradiation. The highest rate of formation of the major product (CO) was 0.64 mmol g–1 h–1 in the presence of 60%SnS2/g-C3N4, which is much higher than the rates attained using single g-C3N4 or SnS2. The stability testing carried out for the most active sample showed that the activity remained almost unchanged after four cycles of the photocatalytic reaction. Since the results of X-ray diffraction and X-ray photoelectron spectroscopy did not show any significant changes either, it can be assumed that the synthesized photocatalyst has a high photocatalytic stability. Study of the charge carrier transfer mechanism provided the conclusion that type II heterojunction is formed between g-C3N4 and SnS2. This increases the charge carrier separation rate, but, simultaneously, it also leads to a decrease in the redox potentials of photogenerated electrons and holes.

Yin et al.[186] obtained the SnS2/g-C3N4 composite photocatalyst by a two-step method comprising sonication and hydrothermal synthesis followed by photodeposition of Au nanoparticles. The resulting photocatalyst was studied in CO2 reduction in an aqueous solution of TEOA under irradiation with a xenon lamp. The major reaction products were CO and CH4, with their formation rates being 94 and 75 mmol g–1 h–1, respectively, while g-C3N4 and SnS2 taken separately had lower activity. The synergistic effect is due to two factors. First, both the hydrophilicity and CO2 adsorption increase in the series g-C3N4 < SnS2 < SnS2/g-C3N4 < SnS2/Au/g-C3N4, which leads to increasing adsorption of the reactants on the photocatalyst surface. Second, it is assumed that gold nanoparticles act as electron mediators for the Z-scheme implemented in the composite photocatalyst, which facilitates the electron transfer and increases the rate of charge carrier separation.

SnS2 is not the only tin sulfide that is used as a photocatalyst. For example, Huo et al.[187] reported hydrothermal synthesis of a composite photocatalyst consisting of porous g-C3N4 and Sn2S3 modified with diethylenetriamine (DETA) (Figure 14). The CO2 reduction experiment was carried out in the gas phase in which CO2 and water were formed upon the reaction of NaHCO3 with HCl. A xenon lamp with a >420 nm cut-off filter was used as the light source. The major reaction products were CH4 and CH3OH, with the rates of their formation being 4.9 and 1.5 mmol g–1 h–1, respectively; this is higher than the activities of the initial g-C3N4 and Sn2S3-DETA. The replacement of H2SO4 by HCl did not result in any noticeable change in the reaction rate, which implies that the acid is not involved in the photocatalytic reaction. It is assumed that the synergistic effect is due to implementation of direct Z-scheme in the synthesized composite photocatalyst with a built-in electric field.

Figure 14
Schematic image of the synthesis of the \( \text{Sn}_{2}\text{S}_{3}\text{-DETA/g-C}_{3}\text{N}_{4} \) photocatalyst[187].

Table 6
\[ \]
Review of some studies on CO2 reduction in the presence of photocatalysts based on tin sulfides and g-C3N4. Refs. [185-191]
(6)

The data on the systems based on tin sulfides and g-C3N4 are summarized in Table 6. Despite the fact that the tin sulfide band gap is narrower than that of g-C3N4, tin sulfides (II, III, IV) are also used to form heterostructures. Apparently, SnS is potentially the most appropriate photocatalyst in this series for the photocatalytic CO2 reduction, because the highest We value of 980 mmol g–1 h–1 was observed for SnS/g-C3N4 system in the gas-phase reduction of CO2 under irradiation with a solar simulator. For comparison, the SnS2/g-C3N4 system modified by gold nanoparticles, which was tested in the liquid-phase reduction of СО2 under irradiation with full-spectrum xenon lamp, provided We of 790 mmol g–1 h–1. Note that tin sulfides are not widely used for the synthesis of heterostructures with g-C3N4, and there are few publications describing the use of tin sulfides as photocatalytic materials for CO2 reduction, although SnS-based photocatalysts can potentially exhibit a fairly high activity in other reactions, for example, decomposition of dyes[192].

4. Heterostructures with 2D materials

4.1. Heterostructures with MXenes

MXenes are a new class of 2D metal carbides, nitrides or carbonitrides discovered in 2011[193]. As a rule, MXenes are obtained from MAX phases by removing layers of A element (usually Al) with a potent etching reagent containing F anions (Figure 15)[193-198]. Different methods of synthesis result in the formation of different surface functional groups (e.g., –O, –F, –OH and –Cl). The general formula of MXenes can be written as Mn+1XnTx, where M is transition metal, X is C or N, and T is a functional group. MXenes have a number of remarkable properties such as high electrical conductivity, hydrophilicity and ordered layered structure, also, the composition of functional groups on their surface can be controlled[197, 198]. MXenes have already proved to possess a high potential for many applications including photocatalysis and are considered as promising materials for photocatalytic CO2 reduction, especially in combination with other semiconductors forming heterostructures. The electronic structure of MXenes is characterized, most often, by the absence of band gap and high electron work function; therefore, they can be used as co-catalysts to form Schottky junction at the semiconductor – MXene interface, similarly to heterostructures based on transition metals and semiconductors.[199]

Figure 15
Periodic Table of Elements with highlighted elements present in other MAX phases (a) and basic diagram of MXene synthesis (b)[198].

Li et al.[200] developed mesoporous Ti3C2Tx/g-C3N4 photocatalysts for CO2 reduction under irradiation with a xenon lamp (Figure 16). The MXene phase was prepared by a widely used method involving etching of Ti3AlC2 with hydrofluoric acid to remove Al layers. The reduction of CO2 in the presence of Ti3C2Tx/g-C3N4 gave CO and CH4 as the major products, with the formation rates being 4.0 mmol g–1 h–1 for CO and 2.2 mmol g–1 h–1 for CH4; these values attained with mesoporous g-C3N4 were 3.1 and 0.88 mmol g–1 h–1, respectively. This effect is attributable to higher specific surface area, larger number of defects, higher electron transfer rate and, consequently, lower recombination rate of photogenerated charge carriers.

Figure 16
Schematic image of processes that take place in the \( \text{Ti}_{3}\text{C}_{2}\text{T}_{x}/\text{g-C}_{3}\text{N}_{4} \) photocatalyst[200].

М.Tahir and B.Tahir[201] synthesized a composite material consisting of g-C3N4 and layered bentonite clay (Bt), on which Ti3C2 particles were then deposited using the ultrasonic self-assembly technique. Bentonite mainly consists of smectite minerals, particularly montmorillonite (usually Ca montmorillonite and Na montmorillonite)[202], has a columnar multilayer structure and can significantly enhance CO2 adsorption and charge carrier separation rate owing to bentinite surface properties and the presence of metal cations, which increase the photocatalytic activity (Figure 17)[202, 203]. Indeed, the photocatalytic experiments on CO2 reduction demonstrate that the Ti3C2/g-C3N4/Bt ternary composite provides a much higher rates of formation of the major products (CO and CH4), especially CH4. It is presumed that the increase in the selectivity to CH4 is caused by the efficient heterotransfer of charge carriers between the phases, which promotes the eight-electron reduction of CO2 to СН4. The effect of a sacrificial agent is also investigated, and it is found that the addition of acetic acid increases the rate of CH4 formation by a factor of 4.2. CH3COOH acts as an electron donor, decreases the charge carrier recombination rate and serves as a source of H2, which is a more thermodynamically favourable reagent for CO2 reduction than water[203-207].

Figure 17
(а) Structure of bentonite; (b) schematic image of processes taking place in the Ti3C2/g-C3N4/Bt photocatalyst[202, 203].

Studies dealing with composite photocatalysts for СО2 reduction based on other MXenes, apart from Ti3C2, have also been reported. For example, Madi et al.[208] synthesized the V2C/g-C3N4 photocatalyst by physical mixing of the components and sonication. The activity of the resulting photocatalyst in the gas-phase reduction of CO2 under irradiation with a 35 W xenon lamp was considerably higher than the activity of g-C3N4 and somewhat higher than the activity of V2AlC/g-C3N4. The authors attributed this synergistic effect to high electronic conductivity, which is one of the main benefits of V2C and to the formation of Schottky junction between g-C3N4 and V2C, which promotes the separation of photogenerated charge carriers[209-213]. A drawback of the composite is low stability: the photocatalyst activity decreases with every irradiation cycle, with the decrease reaching 40% by the end of the third cycle.

Data on the MXene/g-C3N4-based systems are summarized in Table 7. Analysis shows that, unfortunately, modification of g-C3N4 with Ti3C2 does not lead to a significant increase in the activity. It is noteworthy that data on the gas-phase CO2 reduction giving CO and CH4 as the major products are mainly presented in the literature. An outstanding We value of 8400 mmol g–1 h–1 was attained for the system containing bentonite clay under irradiation with a xenon lamp and with addition of acetic acid to the reaction system. A relatively simple V2C/g-C3N4 heterostructure had We of 490 mmol g–1 h–1, which far exceeded the activity of most Ti3C2/g-C3N4 systems. It can be concluded that MXenes have a great potential as co-catalysts for photocatalytic CO2 reduction and that intensive research along this line is needed.

Table 7
\[ \]
Review of some studies on CO2 reduction in the presence of photocatalysts based on MXene/g-C3N4. Refs. [93, 108, 126, 200, 201, 208, 214-218]
(7)

4.2. Heterostructures with partially reduced graphene oxide

As mentioned above, partially reduced graphene oxide (rGO) is a promising material for photocatalysis due to its excellent conductivity and mechanical and optical properties combined with ready availability and the ease of synthesis from graphene oxide (GO)[138]. Moreover, a change in the reduction conditions makes it possible to obtain rGO samples that markedly differ in properties[219]. A traditional method for rGO production is the modified Hummers method, in which graphite is chemically oxidized and then exfoliated to produce graphene oxide[220]. In this stage, oxygenated functional groups appear in the graphene layers (Figure 18)[221]. During the subsequent reduction, some of these groups are removed, and the degree of reduction varies depending on the conditions[222]. It is noteworthy that the production of graphene oxide is accompanied by not only the formation of numerous oxygenated groups on the surface, but also by disruption of the conjugated structure of graphene, which provides for high mobility of charge carriers. The reduction of graphene oxide leads to the partial removal of oxygenated groups and restores the conjugated structure[223].

Figure 18
Schematic image of graphite oxidation and subsequent reduction to rGO[221].

Li et al.[224] performed a two-step process consisting of calcination and hydrothermal treatment and thus synthesized a composite photocatalyst by combining g-C3N4 and rGO with pre-treated multi-walled carbon nanotubes (P-MWNT)[224]. The photocatalytic CO2 reduction was carried out under xenon lamp irradiation with a 420 nm cut-off filter in a TEOA and acetonitrile solution. The highest rates of CO and CH4 formation were 180 and 120 mmol g–1 h–1, respectively. The authors proposed a possible structure of the photocatalyst in which carbon nanotubes act as a mediator between g-C3N4 and rGO and increase the rate of electron transfer. Presumably, photogenerated electrons migrate from g-C3N4 to rGO (either directly or via P-MWNT) as a result of generation of a built-in electric field, which considerably increases the efficiency of separation of photogenerated charge carriers[225, 226].

Bafaqeer et al.[227] used the three-component rGO-bridged ZnV2O6/g-C3N4 photocatalyst for the photocatalytic conversion of CO2 to CH3OH in water. For experiments, the authors designed an externally reflected photoreactor to increase the efficiency of photon energy utilization. As a result, the highest CH3OH formation rate on exposure to a xenon lamp was 630 mmol g–1 h–1. Meanwhile, the CH3OH formation rate in a reactor without external reflection was about 520 mmol g–1 h–1. Thus, even without the use of reflected light energy, the obtained photocatalyst activity is quite high, which is attributable to implementation of the S-scheme between two narrow-band-gap semiconductors and to participation of rGO as a mediator for efficient photogenerated charge carrier separation.

Data on the photocatalysts based on rGO/g-C3N4 reported in recent years are summarized in Table 8.

Table 8
\[ \]
Review of some studies on CO2 reduction in the presence of rGO/g-C3N4-based photocatalysts. Refs. [137, 224, 227-232]
(8)

Only ternary systems, mainly containing some type of vanadate or layered double hydroxide (LDH) as the third component, are considered here, because in the case of binary rGO/g-C3N4 photocatalysts, high СО2 reduction rates are not attained. Layered double hydroxides are a group of layered solids formed by different-valence metal ions and hydroxide ions, which possess a lot of unique properties such as possibility of surface functionalization, intercalation of anions and high chemical stability; this makes LDHs promising photocatalytic materials[233-235]. Most experiments on СО2 reduction catalyzed by rGO/g-C3N4-based photocatalysts were carried out in the liquid phase and the highest We values were found for ZnV2O6/rGO/g-C3N4 systems with Z-scheme charge transfer.

5. Conclusion

Development of g-C3N4-based photocatalysts for СО2 reduction has been a significant research subject in recent years, which is confirmed by the steadily increasing number of relevant publications. The high interest of researchers in the g-C3N4-based photocatalysts is caused by unique properties of g-C3N4, mostly, narrow band gap, which allows activation by visible light. However, narrow-band-gap semiconductors are characterized by high recombination rate of photogenerated charge carriers. Other drawbacks of unmodified g-C3N4 are low adsorption capacity towards CO2 and, as a rule, low specific surface area, which result in moderate rate of CO2 reduction. Modification of g-C3N4 mitigates the effect of these drawbacks on the photocatalytic process; the fabrication of heterostructures with other semiconductors represents the most popular modification method, as it allows for the control of the properties of photocatalysts and the subsequent increase in the CO2 reduction rate.

Currently g-C3N4 heterostructures with metal oxides and sulfides, MXenes, and rGO are the most studied composite photocatalysts based on g-C3N4. Among these materials, TiO2 is mostly studied, because it is a traditional material for many photocatalytic applications owing to its ready availability, lack of toxicity and stability. One more promising oxide semiconductor is CeO2 owing to its unique feature, that is, change in the cerium oxidation state, giving rise to electron traps and additional adsorption sites. MXenes are a new class of 2D compounds with promising properties for the formation of heterostructures and photocatalytic applications. Various approaches to MXene synthesis make it possible to vary their surface properties, which opens up the way for development of a broad range of photocatalysts. Partially reduced graphene oxide rGO has also gained popularity in recent years, and a great potential of rGO for the use in photocatalytic CO2 reduction systems has already been shown.

It is worth mentioning that comparison of the current studies on the photocatalytic CO2 reduction is difficult because no unified experimental methodology has yet been developed. Studies of similar photocatalysts under similar reaction conditions by different research groups often result in significant differences not only in the rates of product formation, but also in the set of products, which may probably be caused by incomplete removal of organic impurities from the reaction medium or by low level of experimentation. Nevertheless, even though establishing of a strict correlation is problematic, it is still possible to determine the average productivity of particular systems to identify the most promising photocatalysts. The results considered in the review provide the conclusion that TiO2 is most promising among the traditional materials used in photocatalysis for the formation of heterostructures with g-C3N4; this provides We values at the level of 200 – 600 mmol g–1 h–1, and in some cases, up to 6700 mmol g–1 h–1. In addition, there is a persistent trend towards the use of 2D materials such as MXenes and rGO for the formation of heterostructures. Indeed, in some studies, fairly high CO2 reduction rates have already been attained by using Ti3C2 or rGO to modify g-C3N4. The major products formed upon СО2 reduction catalyzed by heterostructures based on traditional semiconductors and those based on new 2D materials are CO and CH4. However, in some cases, outstanding rates of CH3OH formation have been achieved using rGO.

Thus, data on the activity of photocatalysts reported in the literature indicate that practical implementation of the photocatalytic CO2 reduction requires much more active systems and, probably, the main trend of research in the photocatalytic CO2 reduction for the next decade will be the search for and development of composite photocatalysts based on g-C3N4 combined with various 2D semiconductor materials and with materials (e.g., bentonite) that may significantly improve the textural properties of composites.

6. List of acronyms

AQE — apparent quantum efficiency,
bpy — bipyridine,
CB — conduction band,
DETA — diethylenetriamine,
DMF — dimethylformamide,
g-C3N4 — graphitic carbon nitride,
GO — graphene oxide,
GP — gas phase,
LDH — layered double hydroxide,
LP — liquid phase,
m — mesoporous,
MXene — a class of two-dimensional metal carbides, nitrides or carbonitrides,
NHE — normal hydrogen electrode,
P-MWNT — pre-treated multi-walled carbon nanotubes,
PNS — porous nanosheets,
rGO — partially reduced graphene oxide,
SC — semiconductor,
TEOA — triethanolamine,
VB — valence band,
We
— total number of electrons consumed in a photocatalytic process per unit reaction time and per unit photocatalyst mass.

References

1.
Net greenhouse gas balance with cover crops in semi-arid irrigated cropping systems
Acharya P., Ghimire R., Paye W.S., Ganguli A.C., DelGrosso S.J.
Scientific Reports, Springer Nature, 2022
3.
Recent advances in g-C3N4-based heterojunction photocatalysts
Li Y., Zhou M., Cheng B., Shao Y.
Journal of Materials Science and Technology, Springer Nature, 2020
4.
Carbon dioxide and "methanol" economy: advances in the catalytic synthesis of methanol from CO2
Maximov Anton, Beletskaya Irina
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2024
6.
A review on photochemical, biochemical and electrochemical transformation of CO2 into value-added products
Yaashikaa P.R., Senthil Kumar P., Varjani S.J., Saravanan A.
Journal of CO2 Utilization, Elsevier, 2019
7.
Broadening the Action Spectrum of TiO2-Based Photocatalysts to Visible Region by Substituting Platinum with Copper
Saraev A.A., Kurenkova A.Y., Gerasimov E.Y., Kozlova E.A.
Nanomaterials, Multidisciplinary Digital Publishing Institute (MDPI), 2022
8.
Mechanisms of catalytic electrochemical reactions of oxygen reduction (ORR) and carbon dioxide reduction (CO2RR)
Kuzmin A.V., Shainyan B.A.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2023
9.
CO2 utilization: Developments in conversion processes
Alper E., Yuksel Orhan O.
Petroleum, KeAi Communications Co., 2017
10.
A comprehensive review on different approaches for CO2 utilization and conversion pathways
Saravanan A., Senthil kumar P., Vo D.N., Jeevanantham S., Bhuvaneswari V., Anantha Narayanan V., Yaashikaa P.R., Swetha S., Reshma B.
Chemical Engineering Science, Elsevier, 2021
11.
A short review of catalysis for CO2 conversion
Ma J., Sun N., Zhang X., Zhao N., Xiao F., Wei W., Sun Y.
Catalysis Today, Elsevier, 2009
12.
Semiconductor photocatalysts and mechanisms of carbon dioxide reduction and nitrogen fixation under UV and visible light
Kozlova E.A., Lyulyukin M.N., Kozlov D.V., Parmon V.N.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2021
13.
An overview of the reaction conditions for an efficient photoconversion of CO2
Meryem S.S., Nasreen S., Siddique M., Khan R.
Reviews in Chemical Engineering, Walter de Gruyter, 2017
14.
Theoretical Insights into Heterogeneous (Photo)electrochemical CO2 Reduction
Xu S., Carter E.A.
Chemical Reviews, American Chemical Society (ACS), 2018
15.
Thermodynamic Aspects of Electrocatalytic CO2 Reduction in Acetonitrile and with an Ionic Liquid as Solvent or Electrolyte
16.
A critical review on TiO2 based photocatalytic CO2 reduction system: Strategies to improve efficiency
Shehzad N., Tahir M., Johari K., Murugesan T., Hussain M.
Journal of CO2 Utilization, Elsevier, 2018
17.
The green chemistry paradigm in modern organic synthesis
Zlotin S.G., Egorova K.S., Ananikov V.P., Akulov A.A., Varaksin M.V., Chupakhin O.N., Charushin V.N., Bryliakov K.P., Averin A.D., Beletskaya I.P., Dolengovski E.L., Budnikova Y.H., Sinyashin O.G., Gafurov Z.N., Kantyukov A.O., et. al.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2023
18.
Graphitic carbon nitride materials in dual metallo-photocatalysis: A promising concept in organic synthesis
Dam B., Das B., Patel B.K.
Green Chemistry, Royal Society of Chemistry (RSC), 2023
19.
A Comprehensive Review of Graphitic Carbon Nitride (g-C3N4)–Metal Oxide-Based Nanocomposites: Potential for Photocatalysis and Sensing
Alaghmandfard A., Ghandi K.
Nanomaterials, Multidisciplinary Digital Publishing Institute (MDPI), 2022
20.
Functional carbon nitride materials — design strategies for electrochemical devices
Kessler F.K., Zheng Y., Schwarz D., Merschjann C., Schnick W., Wang X., Bojdys M.J.
Nature Reviews Materials, Springer Nature, 2017
23.
Molecular engineering of polymeric carbon nitride: advancing applications from photocatalysis to biosensing and more
Zhou Z., Zhang Y., Shen Y., Liu S., Zhang Y.
Chemical Society Reviews, Royal Society of Chemistry (RSC), 2018
24.
Solvothermal modification of graphitic C3N4 with Ni and Co phthalocyanines: Structural, optoelectronic and surface properties
Lebedev L.A., Chebanenko M.I., Dzhevaga E.V., Martinson K.D., Popkov V.I.
Mendeleev Communications, Elsevier, 2022
25.
Emerging beyond-graphene elemental 2D materials for energy and catalysis applications
Fan F.R., Wang R., Zhang H., Wu W.
Chemical Society Reviews, Royal Society of Chemistry (RSC), 2021
26.
Single noble metal atoms doped 2D materials for catalysis
Liu D., Barbar A., Najam T., Javed M.S., Shen J., Tsiakaras P., Cai X.
Applied Catalysis B: Environmental, Elsevier, 2021
28.
Adsorption of CO2, O2, NO and CO on s-triazine-based g-C3N4 surface
Zhu B., Wageh S., Al-Ghamdi A.A., Yang S., Tian Z., Yu J.
Catalysis Today, Elsevier, 2019
29.
Adsorption investigation of CO2 on g-C3N4 surface by DFT calculation
Zhu B., Zhang L., Xu D., Cheng B., Yu J.
Journal of CO2 Utilization, Elsevier, 2017
30.
Construction of S-scheme g-C3N4/ZrO2 heterostructures for enhancing photocatalytic disposals of pollutants and electrocatalytic hydrogen evolution
31.
In situ fabrication of a novel S-scheme heterojunction photocatalyts Bi2O3/P-C3N4 to enhance levofloxacin removal from water
Zhang X., Zhang Y., Jia X., Zhang N., Xia R., Zhang X., Wang Z., Yu M.
Separation and Purification Technology, Elsevier, 2021
32.
Amine-Modified S-Scheme Porous g-C3N4/CdSe–Diethylenetriamine Composite with Enhanced Photocatalytic CO2 Reduction Activity
Huo Y., Zhang J., Dai K., Liang C.
ACS Applied Energy Materials, American Chemical Society (ACS), 2021
33.
Recent advances, application and prospect in g-C3N4-based S-scheme heterojunction photocatalysts
Hao P., Chen Z., Yan Y., Shi W., Guo F.
Separation and Purification Technology, Elsevier, 2024
34.
Pathways towards a systematic development of Z-scheme photocatalysts for CO2 reduction
Hezam A., Peppel T., Strunk J.
Current Opinion in Green and Sustainable Chemistry, Elsevier, 2023
35.
Review on g-C3N4-based S-scheme heterojunction photocatalysts
Li Y., Xia Z., Yang Q., Wang L., Xing Y.
Journal of Materials Science and Technology, Springer Nature, 2022
36.
Advances in Z‐scheme semiconductor photocatalysts for the photoelectrochemical applications: A review
Li J., Yuan H., Zhang W., Jin B., Feng Q., Huang J., Jiao Z.
Carbon Energy, Wiley, 2022
37.
Graphitic Carbon Nitride-Based Z-Scheme Structure for Photocatalytic CO2 Reduction
Lin J., Tian W., Zhang H., Duan X., Sun H., Wang S.
Energy & Fuels, American Chemical Society (ACS), 2020
38.
Emerging polymeric carbon nitride Z-scheme systems for photocatalysis
Liao G., Li C., Li X., Fang B.
Cell Reports Physical Science, Elsevier, 2021
39.
Optical Properties and Electronic Structure of Amorphous Germanium
Tauc J., Grigorovici R., Vancu A.
Physica Status Solidi (B): Basic Research, Wiley, 1966
40.
How To Correctly Determine the Band Gap Energy of Modified Semiconductor Photocatalysts Based on UV-Vis Spectra.
Makuła P., Pacia M., Macyk W.
Journal of Physical Chemistry Letters, American Chemical Society (ACS), 2018
41.
Biomimetic Donor–Acceptor Motifs in Conjugated Polymers for Promoting Exciton Splitting and Charge Separation
Ou H., Chen X., Lin L., Fang Y., Wang X.
Angewandte Chemie - International Edition, Wiley, 2018
43.
Coating Polymeric Carbon Nitride Photoanodes on Conductive Y:ZnO Nanorod Arrays for Overall Water Splitting
Fang Y., Xu Y., Li X., Ma Y., Wang X.
Angewandte Chemie - International Edition, Wiley, 2018
45.
Comparative Studies of g-C3N4 and C3N3S3 Organic Semiconductors—Synthesis, Properties, and Application in the Catalytic Oxygen Reduction
Wierzyńska E., Pisarek M., Łęcki T., Skompska M.
Molecules, Multidisciplinary Digital Publishing Institute (MDPI), 2023
48.
Switching of semiconducting behavior from n -type to p -type induced high photocatalytic NO removal activity in g-C 3 N 4
49.
A review on g-C3N4 for photocatalytic water splitting and CO2 reduction
Ye S., Wang R., Wu M., Yuan Y.
Applied Surface Science, Elsevier, 2015
50.
Polymeric graphitic carbon nitride (g-C3N4)-based semiconducting nanostructured materials: Synthesis methods, properties and photocatalytic applications
Reddy K.R., Reddy C.V., Nadagouda M.N., Shetti N.P., Jaesool S., Aminabhavi T.M.
Journal of Environmental Management, Elsevier, 2019
51.
Polymeric Photocatalysts Based on Graphitic Carbon Nitride
Cao S., Low J., Yu J., Jaroniec M.
Advanced Materials, Wiley, 2015
53.
Soft and hard templating of graphitic carbon nitride
Yang Z., Zhang Y., Schnepp Z.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2015
55.
A review on synthesis, modification method, and challenges of light-driven H2 evolution using g-C3N4-based photocatalyst
Abu-Sari S.M., Daud W.M., Patah M.F., Ang B.C.
Advances in Colloid and Interface Science, Elsevier, 2022
56.
Graphitic carbon nitrides (g-C3N4) with comparative discussion to carbon materials
57.
Synthesis and modification of ultrathin g-C3N4 for photocatalytic energy and environmental applications
Huang H., Jiang L., Yang J., Zhou S., Yuan X., Liang J., Wang H., Wang H., Bu Y., Li H.
Renewable and Sustainable Energy Reviews, Elsevier, 2023
58.
High-Efficiency g-C3N4 Based Photocatalysts for CO2 Reduction: Modification Methods
Wang Q., Fang Z., Zhang W., Zhang D.
Advanced Fiber Materials, Springer Nature, 2022
59.
Li Y., Zhang M., Zhou L., Yang S., Wu Z., Ma Y.
Wuli Huaxue Xuebao/ Acta Physico - Chimica Sinica, Beijing University Press, 2020
60.
Recent advances in solar‐driven CO 2 reduction over g‐C 3 N 4 ‐based photocatalysts
Xu Q., Xia Z., Zhang J., Wei Z., Guo Q., Jin H., Tang H., Li S., Pan X., Su Z., Wang S.
Carbon Energy, Wiley, 2022
61.
Photocatalytic CO2 reduction over g-C3N4 based heterostructures: Recent progress and prospects
Ghosh U., Majumdar A., Pal A.
Journal of Environmental Chemical Engineering, Elsevier, 2021
62.
Design and application of g-C3N4-based materials for fuels photosynthesis from CO2 or H2O based on reaction pathway insights
63.
Photocatalytic carbon dioxide reduction: Exploring the role of ultrathin 2D graphitic carbon nitride (g-C3N4)
Aggarwal M., Basu S., Shetti N.P., Nadagouda M.N., Kwon E.E., Park Y., Aminabhavi T.M.
Chemical Engineering Journal, Elsevier, 2021
64.
Recent progress on the development of g-C3N4 based composite material and their photocatalytic application of CO2 reductions
Prasad C., Madkhali N., Govinda V., Choi H.Y., Bahadur I., Sangaraju S.
Journal of Environmental Chemical Engineering, Elsevier, 2023
65.
A critical review of g-C3N4-based photocatalytic membrane for water purification
Zhang M., Yang Y., An X., Hou L.
Chemical Engineering Journal, Elsevier, 2021
66.
Recent developments of doped g-C3N4 photocatalysts for the degradation of organic pollutants
Liu X., Ma R., Zhuang L., Hu B., Chen J., Liu X., Wang X.
Critical Reviews in Environmental Science and Technology, Taylor & Francis, 2020
68.
Photocatalysts Pt/TiO2 for CO2 reduction under ultraviolet irradiation
Kurenkova A.Y., Gerasimov E.Y., Saraev A.A., Kozlova E.A.
Russian Chemical Bulletin, Springer Nature, 2023
69.
Noble metal deposited graphitic carbon nitride based heterojunction photocatalysts
Kavitha R., Nithya P.M., Girish Kumar S.
Applied Surface Science, Elsevier, 2020
70.
Selectivity Control of CO2 Reduction over Pt/g-C3N4 Photocatalysts under Visible Light
Saraev A.A., Kurenkova A.Y., Zhurenok A.V., Gerasimov E.Y., Kozlova E.A.
Catalysts, Multidisciplinary Digital Publishing Institute (MDPI), 2023
71.
Comprehensive Review on g-C3N4-Based Photocatalysts for the Photocatalytic Hydrogen Production under Visible Light
Zhurenok A.V., Vasilchenko D.B., Kozlova E.A.
International Journal of Molecular Sciences, Multidisciplinary Digital Publishing Institute (MDPI), 2022
72.
Photocatalysis on TiO2 Surfaces: Principles, Mechanisms, and Selected Results
Linsebigler A.L., Lu G., Yates J.T.
Chemical Reviews, American Chemical Society (ACS), 1995
73.
Titanium dioxide nanotubes: synthesis, structure, properties and applications
Rempel A.A., Valeeva A.A., Vokhmintsev A.S., Weinstein I.A.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2021
74.
Nonstoichiometry, structure and properties of nanocrystalline oxides, carbides and sulfides
Valeeva A.A., Rempel A.A., Rempel S.V., Sadovnikov S.I., Gusev A.I.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2021
75.
CeO2-TiO2 as a visible light active catalyst for the photoreduction of CO2 to methanol
Abdullah H., Khan M.R., Pudukudy M., Yaakob Z., Ismail N.A.
Journal of Rare Earths, Chinese Society of Rare Earths, 2015
76.
Engineering heterogeneous semiconductors for solar water splitting
Li X., Yu J., Low J., Fang Y., Xiao J., Chen X.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2015
79.
New understanding of the difference of photocatalytic activity among anatase, rutile and brookite TiO2
Zhang J., Zhou P., Liu J., Yu J.
Physical Chemistry Chemical Physics, Royal Society of Chemistry (RSC), 2014
80.
Recent Progress in the Abatement of Hazardous Pollutants Using Photocatalytic TiO2-Based Building Materials
Gopalan A., Lee J., Saianand G., Lee K., Sonar P., Dharmarajan R., Hou Y., Ann K., Kannan V., Kim W.
Nanomaterials, Multidisciplinary Digital Publishing Institute (MDPI), 2020
81.
Sol–gel synthesis, characterisation and photocatalytic activity of pure, W-, Ag- and W/Ag co-doped TiO2 nanopowders
Tobaldi D.M., Pullar R.C., Gualtieri A.F., Seabra M.P., Labrincha J.A.
Chemical Engineering Journal, Elsevier, 2013
82.
Modified TiO2 For Environmental Photocatalytic Applications: A Review
Daghrir R., Drogui P., Robert D.
Industrial & Engineering Chemistry Research, American Chemical Society (ACS), 2013
84.
Reactivity of anatase TiO(2) nanoparticles: the role of the minority (001) surface.
Gong X., Selloni A.
Journal of Physical Chemistry B, American Chemical Society (ACS), 2005
85.
Preparation of anatase/rutile TiO2/SnO2 hollow heterostructures for gas sensor
Jia C., Dong T., Li M., Wang P., Yang P.
Journal of Alloys and Compounds, Elsevier, 2018
86.
The preparation, and applications of g-C3N4/TiO2 heterojunction catalysts—a review
Zhou L., Wang L., Zhang J., Lei J., Liu Y.
Research on Chemical Intermediates, Springer Nature, 2016
87.
Exploring the visible light–assisted conversion of CO2 into methane and methanol, using direct Z-scheme TiO2@g-C3N4 nanosheets: synthesis and photocatalytic performance
Mehregan S., Hayati F., Mehregan M., Isari A.A., Jonidi Jafari A., Giannakis S., Kakavandi B.
Environmental Science and Pollution Research, Springer Nature, 2022
89.
Photocatalytic degradation of carbofuran using semiconductor oxides
Mahalakshmi M., Arabindoo B., Palanichamy M., Murugesan V.
Journal of Hazardous Materials, Elsevier, 2007
90.
What is Degussa (Evonik) P25? Crystalline composition analysis, reconstruction from isolated pure particles and photocatalytic activity test
Ohtani B., Prieto-Mahaney O.O., Li D., Abe R.
Journal of Photochemistry and Photobiology A: Chemistry, Elsevier, 2010
91.
TiO2 modified g-C3N4 with enhanced photocatalytic CO2 reduction performance
Wang H., Li H., Chen Z., Li J., Li X., Huo P., Wang Q.
Solid State Sciences, Elsevier, 2020
92.
The superior photocatalytic activity of Nb doped TiO2/g-C3N4 direct Z-scheme system for efficient conversion of CO2 into valuable fuels
Thanh Truc N.T., Giang Bach L., Thi Hanh N., Pham T., Thi Phuong Le Chi N., Tran D.T., Nguyen M.V., Nguyen V.N.
Journal of Colloid and Interface Science, Elsevier, 2019
94.
Band-Gap Engineering: From Physics and Materials to New Semiconductor Devices
Capasso F.
Science, American Association for the Advancement of Science (AAAS), 1987
96.
One step synthesis of niobium doped titania nanotube arrays to form (N,Nb) co-doped TiO2with high visible light photoelectrochemical activity
Cottineau T., Béalu N., Gross P., Pronkin S.N., Keller N., Savinova E.R., Keller V.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2013
97.
Towards visible-light photocatalysis for environmental applications: band-gap engineering versus photons absorption—a review
98.
Recent Advances in Heterogeneous Photocatalytic CO2 Conversion to Solar Fuels
Li K., Peng B., Peng T.
ACS Catalysis, American Chemical Society (ACS), 2016
99.
Thermally assisted photocatalytic conversion of CO2–H2O to C2H4 over carbon doped In2S3 nanosheets
Wang L., Zhao B., Wang C., Sun M., Yu Y., Zhang B.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2020
100.
Plasmonic Active “Hot Spots”‐Confined Photocatalytic CO 2 Reduction with High Selectivity for CH 4 Production
Jiang X., Huang J., Bi Z., Ni W., Gurzadyan G., Zhu Y., Zhang Z.
Advanced Materials, Wiley, 2022
102.
Enhanced Photocatalytic Reduction of CO2 over pg-C3N4-supported TiO2 Nanoparticles with Ag Modification
Zhang H., Bian H., Wang F., Zhu L., Zhang S., Xia D.
Colloids and Surfaces A: Physicochemical and Engineering Aspects, Elsevier, 2023
103.
Preparation of hierarchical g-C3N4@TiO2 hollow spheres for enhanced visible-light induced catalytic CO2 reduction
104.
Efficient Z-scheme photocatalysts of ultrathin g-C3N4-wrapped Au/TiO2-nanocrystals for enhanced visible-light-driven conversion of CO2 with H2O
105.
Photoreduction of CO2 in the presence of CH4 over g-C3N4 modified with TiO2 nanoparticles at room temperature
Chen M., Wu J., Lu C., Luo X., Huang Y., Jin B., Gao H., Zhang X., Argyle M., Liang Z.
Green Energy and Environment, KeAi Communications Co., 2021
107.
Solar-driven CO2 conversion over Co2+ doped 0D/2D TiO2/g-C3N4 heterostructure: Insights into the role of Co2+ and cocatalyst
Shi H., Du J., Hou J., Ni W., Song C., Li K., Gurzadyan G.G., Guo X.
Journal of CO2 Utilization, Elsevier, 2020
108.
2D/2D/0D TiO2/C3N4/Ti3C2 MXene composite S-scheme photocatalyst with enhanced CO2 reduction activity
He F., Zhu B., Cheng B., Yu J., Ho W., Macyk W.
Applied Catalysis B: Environmental, Elsevier, 2020
109.
Nanocavity-assisted single-crystalline Ti3+ self-doped blue TiO2(B) as efficient cocatalyst for high selective CO2 photoreduction of g-C3N4
Kumar D.P., Rangappa A.P., Shim H.S., Do K.H., Hong Y., Gopannagari M., Reddy K.A., Bhavani P., Reddy D.A., Song J.K., Kim T.K.
Materials Today Chemistry, Elsevier, 2022
111.
Tuning CH4 productivity from visible light driven gas‐phase CO2 photocatalytic reduction on doped g‐C3N4/TiO2 heterojunctions
Hammoud L., Marchal C., Colbeau-Justin C., Toufaily J., Hamieh T., Caps V., Keller V.
Energy Technology, Wiley, 2023
112.
Defective Heterojunctions in CO2 Photoreduction: Enabling Ultrafast Interfacial Charge Transfer and Selective Methanation
Cheng S., Sun Z., Lim K.H., Liu K., Wibowo A.A., Du T., Liu L., Nguyen H.T., Li G.K., Yin Z., Kawi S.
Applied Catalysis B: Environmental, Elsevier, 2024
113.
Zinc Oxide Based Composite Materials for Advanced Supercapacitors
Wang Y., Xiao X., Xue H., Pang H.
ChemistrySelect, Wiley, 2018
114.
Zinc oxide nanoparticles: an excellent biomaterial for bioengineering applications
Ringu T., Ghosh S., Das A., Pramanik N.
Emergent Materials, Springer Nature, 2022
115.
One-Dimensional Zinc Oxide Nanomaterials for Application in High-Performance Advanced Optoelectronic Devices
Ding M., Guo Z., Zhou L., Fang X., Zhang L., Zeng L., Xie L., Zhao H.
Crystals, Multidisciplinary Digital Publishing Institute (MDPI), 2018
117.
Origin of n-type conductivity in ZnO crystal and formation of Zn and ZnO nanoparticles by laser radiation
Kaupužs J., Medvids A., Onufrijevs P., Mimura H.
Optics and Laser Technology, Elsevier, 2019
119.
Improved photocatalytic activity of porous ZnO nanosheets by thermal deposition graphene-like g-C3N4 for CO2 reduction with H2O vapor
120.
Thermodynamic and kinetic analysis of heterogeneous photocatalysis for semiconductor systems
Liu B., Zhao X., Terashima C., Fujishima A., Nakata K.
Physical Chemistry Chemical Physics, Royal Society of Chemistry (RSC), 2014
121.
Temperature Dependence of Solar Light Assisted CO2 Reduction on Ni Based Photocatalyst
Albero J., Garcia H., Corma A.
Topics in Catalysis, Springer Nature, 2016
123.
Hydrocarbon production by addition of Cu‐ZnO on  g‐C 3 N 4  for  CO 2  conversion
Zhu Z., Chen C., Wu R.
Journal of the Chinese Chemical Society, Wiley, 2020
124.
EPR Investigation on Electron Transfer of 2D/3D g‐C 3 N 4 /ZnO S‐Scheme Heterojunction for Enhanced CO 2 Photoreduction
Sayed M., Zhu B., Kuang P., Liu X., Cheng B., Ghamdi A.A., Wageh S., Zhang L., Yu J.
Advanced Sustainable Systems, Wiley, 2021
125.
Efficient visible light activities of Ag modified ZnO/g-C3N4 composite for CO2 conversion
Arif U., Ali F., Bahader A., Ali S., Zada A., Raziq F.
Inorganic Chemistry Communication, Elsevier, 2022
126.
MXene Ti3C2 decorated g-C3N4/ZnO photocatalysts with improved photocatalytic performance for CO2 reduction
Li J., Wang Y., Wang Y., Guo Y., Zhang S., Song H., Li X., Gao Q., Shang W., Hu S., Zheng H., Li X.
Nano Materials Science, Elsevier, 2023
127.
Enhanced photodegradation of tetracycline in wastewater and conversion of CO2 by solar light assisted ZnO/g-C3N4
Pham T.H., Tran M.H., Chu T.T., Myung Y., Jung S.H., Mapari M.G., Taeyoung K.
Environmental Research, Elsevier, 2023
128.
Cerium oxide nanoparticles prepared in self-assembled systems
Bumajdad A., Eastoe J., Mathew A.
Advances in Colloid and Interface Science, Elsevier, 2009
129.
A critical review on relationship of CeO2-based photocatalyst towards mechanistic degradation of organic pollutant
Fauzi A.A., Jalil A.A., Hassan N.S., Aziz F.F., Azami M.S., Hussain I., Saravanan R., Vo D.-.
Chemosphere, Elsevier, 2022
130.
Chemistry and electrochemistry of CeO2-based interlayers: Prolonging the lifetime of solid oxide fuel and electrolysis cells
Erpalov M.V., Tarutin A.P., Danilov N.A., Osinkin D.A., Medvedev D.A.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2023
131.
CeO2 as a photocatalytic material for CO2 conversion: A review
Tran D.P., Pham M., Bui X., Wang Y., You S.
Solar Energy, Elsevier, 2022
133.
Controlled assemble of hollow heterostructured g-C3N4@CeO2 with rich oxygen vacancies for enhanced photocatalytic CO2 reduction
Liang M., Borjigin T., Zhang Y., Liu B., Liu H., Guo H.
Applied Catalysis B: Environmental, Elsevier, 2019
134.
Photocatalytic CO2 reduction
Fang S., Rahaman M., Bharti J., Reisner E., Robert M., Ozin G.A., Hu Y.H.
Nature Reviews Methods Primers, Springer Nature, 2023
135.
Recent advances in the development of sunlight-driven hollow structure photocatalysts and their applications
Nguyen C.C., Vu N.N., Do T.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2015
137.
rGO modified R-CeO2/g-C3N4 multi-interface contact S-scheme photocatalyst for efficient CO2 photoreduction
Li X., Guan J., Jiang H., Song X., Huo P., Wang H.
Applied Surface Science, Elsevier, 2021
138.
Reduced graphene oxide today
Tarcan R., Todor-Boer O., Petrovai I., Leordean C., Astilean S., Botiz I.
Journal of Materials Chemistry C, Royal Society of Chemistry (RSC), 2020
139.
Fabricated rGO-modified Ag2S nanoparticles/g-C3N4 nanosheets photocatalyst for enhancing photocatalytic activity
Li X., Shen D., Liu C., Li J., Zhou Y., Song X., Huo P., Wang H., Yan Y.
Journal of Colloid and Interface Science, Elsevier, 2019
140.
CeO2/3D g-C3N4 heterojunction deposited with Pt cocatalyst for enhanced photocatalytic CO2 reduction
Zhao X., Guan J., Li J., Li X., Wang H., Huo P., Yan Y.
Applied Surface Science, Elsevier, 2021
141.
Fabricated g-C3N4/Ag/m-CeO2 composite photocatalyst for enhanced photoconversion of CO2
Wang H., Guan J., Li J., Li X., Ma C., Huo P., Yan Y.
Applied Surface Science, Elsevier, 2020
142.
Advantageous roles of phosphate decorated octahedral CeO2 {111}/g-C3N4 in boosting photocatalytic CO2 reduction: Charge transfer bridge and Lewis basic site
Li W., Jin L., Gao F., Wan H., Pu Y., Wei X., Chen C., Zou W., Zhu C., Dong L.
Applied Catalysis B: Environmental, Elsevier, 2021
143.
Facile synthesis of a Z-scheme CeO2/C3N4 heterojunction with enhanced charge transfer for CO2 photoreduction
Chen J., Xiao Y., Wang N., Kang X., Wang D., Wang C., Liu J., Jiang Y., Fu H.
Science China Materials, Springer Nature, 2023
144.
g-C3N4 quantum dots-modified mesoporous CeO2 composite photocatalyst for enhanced CO2 photoreduction
Jiang H., Li X., Chen S., Wang H., Huo P.
Journal of Materials Science: Materials in Electronics, Springer Nature, 2020
145.
Enhanced reduction and oxidation capability over the CeO2/g-C3N4 hybrid through surface carboxylation: performance and mechanism
Hu H., Hu J., Wang X., Gan J., Su M., Ye W., Zhang W., Ma X., Wang H.
Catalysis Science and Technology, Royal Society of Chemistry (RSC), 2020
146.
Inorganic nanostructures for photoelectrochemical and photocatalytic water splitting
Osterloh F.E.
Chemical Society Reviews, Royal Society of Chemistry (RSC), 2013
147.
Hierarchical hollow spheres composed of ultrathin Fe2O3 nanosheets for lithium storage and photocatalytic water oxidation
Zhu J., Yin Z., Yang D., Sun T., Yu H., Hoster H.E., Hng H.H., Zhang H., Yan Q.
Energy and Environmental Science, Royal Society of Chemistry (RSC), 2013
148.
Magnetically separable nanocomposites based on ZnO and their applications in photocatalytic processes: A review
Shekofteh-Gohari M., Habibi-Yangjeh A., Abitorabi M., Rouhi A.
Critical Reviews in Environmental Science and Technology, Taylor & Francis, 2018
149.
Activation of Carbon Nitride Solids by Protonation: Morphology Changes, Enhanced Ionic Conductivity, and Photoconduction Experiments
Zhang Y., Thomas A., Antonietti M., Wang X.
Journal of the American Chemical Society, American Chemical Society (ACS), 2008
150.
Semiconductor photocatalysis--mechanistic and synthetic aspects.
Kisch H.
Angewandte Chemie - International Edition, Wiley, 2012
151.
Action and mechanism of semiconductor photocatalysis on degradation of organic pollutants in water treatment: A review
152.
Synthesis of Z-scheme α-Fe2O3/g-C3N4 composite with enhanced visible-light photocatalytic reduction of CO2 to CH3OH
Guo H., Chen M., Zhong Q., Wang Y., Ma W., Ding J.
Journal of CO2 Utilization, Elsevier, 2019
154.
Effect of solvents on photocatalytic reduction of CO2 mediated by cobalt complex
Lin J., Qin B., Zhao G.
Journal of Photochemistry and Photobiology A: Chemistry, Elsevier, 2018
155.
Semiconductor Photocatalysis: “Tell Us the Complete Story!”
Kamat P.V., Jin S.
ACS Energy Letters, American Chemical Society (ACS), 2018
156.
Photocatalytic Hydrogen Generation Using Metal-Decorated TiO2: Sacrificial Donors vs True Water Splitting
Hainer A.S., Hodgins J.S., Sandre V., Vallieres M., Lanterna A.E., Scaiano J.C.
ACS Energy Letters, American Chemical Society (ACS), 2018
157.
Why Seeing Is Not Always Believing: Common Pitfalls in Photocatalysis and Electrocatalysis
Christopher P., Jin S., Sivula K., Kamat P.V.
ACS Energy Letters, American Chemical Society (ACS), 2021
158.
Systematic Assessment of Solvent Selection in Photocatalytic CO2 Reduction
Das R., Chakraborty S., Peter S.C.
ACS Energy Letters, American Chemical Society (ACS), 2021
161.
Artificial Trees for Artificial Photosynthesis: Construction of Dendrite-Structured α-Fe2O3/g-C3N4 Z-Scheme System for Efficient CO2 Reduction into Solar Fuels
Shen Y., Han Q., Hu J., Gao W., Wang L., Yang L., Gao C., Shen Q., Wu C., Wang X., Zhou X., Zhou Y., Zou Z.
ACS Applied Energy Materials, American Chemical Society (ACS), 2020
162.
Self-Templated Synthesis of Nanoporous CdS Nanostructures for Highly Efficient Photocatalytic Hydrogen Production under Visible Light
Bao N., Shen L., Takata T., Domen K.
Chemistry of Materials, American Chemical Society (ACS), 2007
163.
Visible-light-driven hydrogen production with extremely high quantum efficiency on Pt–PdS/CdS photocatalyst
Yan H., Yang J., Ma G., Wu G., Zong X., Lei Z., Shi J., Li C.
Journal of Catalysis, Elsevier, 2009
164.
Photochemical hydrogen production with cadmium sulfide suspensions
Buehler N., Meier K., Reber J.F.
The Journal of Physical Chemistry, American Chemical Society (ACS), 1984
165.
Direct splitting of H2S into H2 and S on CdS-based photocatalyst under visible light irradiation
Ma G., Yan H., Shi J., Zong X., Lei Z., Li C.
Journal of Catalysis, Elsevier, 2008
166.
Photocorrosion inhibition of CdS-based catalysts for photocatalytic overall water splitting
167.
A kinetic study of CdS photocorrosion by intensity modulated photocurrent and photoelectrochemical impedance spectroscopy
168.
Photoinduced self-stability mechanism of CdS photocatalyst: The dependence of photocorrosion and H2-evolution performance
Chen Y., Zhong W., Chen F., Wang P., Fan J., Yu H.
Journal of Materials Science and Technology, Springer Nature, 2022
170.
Synthesis of the g-C3N4/CdS Nanocomposite with a Chemically Bonded Interface for Enhanced Sunlight-Driven CO2 Photoreduction
Vu N., Kaliaguine S., Do T.
ACS Applied Energy Materials, American Chemical Society (ACS), 2020
172.
Fabrication of Heterogeneous-Phase Solid-Solution Promoting Band Structure and Charge Separation for Enhancing Photocatalytic CO2 Reduction: A Case of ZnXCa1–XIn2S4
Zeng C., Huang H., Zhang T., Dong F., Zhang Y., Hu Y.
ACS applied materials & interfaces, American Chemical Society (ACS), 2017
175.
Persian buttercup-like BiOBrxCl1-x solid solution for photocatalytic overall CO2 reduction to CO and O2
Gao M., Yang J., Sun T., Zhang Z., Zhang D., Huang H., Lin H., Fang Y., Wang X.
Applied Catalysis B: Environmental, Elsevier, 2019
176.
Z-scheme g-C3N4 nanosheet photocatalyst decorated with mesoporous CdS for the photoreduction of carbon dioxide
Mkhalid I.A., Mohamed R.M., Ismail A.A., Alhaddad M.
Ceramics International, Elsevier, 2021
178.
Construction of highly efficient Z-scheme ZnxCd1-xS/Au@g-C3N4 ternary heterojunction composite for visible-light-driven photocatalytic reduction of CO2 to solar fuel
Madhusudan P., Shi R., Xiang S., Jin M., Chandrashekar B.N., Wang J., Wang W., Peng O., Amini A., Cheng C.
Applied Catalysis B: Environmental, Elsevier, 2021
179.
Monodisperse SnS2 Nanosheets for High-Performance Photocatalytic Hydrogen Generation
Yu J., Xu C., Ma F., Hu S., Zhang Y., Zhen L.
ACS applied materials & interfaces, American Chemical Society (ACS), 2014
180.
High-Performance Visible-Light-Driven SnS2/SnO2 Nanocomposite Photocatalyst Prepared via In situ Hydrothermal Oxidation of SnS2 Nanoparticles
Zhang Y.C., Du Z.N., Li K.W., Zhang M., Dionysiou D.D.
ACS applied materials & interfaces, American Chemical Society (ACS), 2011
181.
In situ growth of Ag2S quantum dots on SnS2 nanosheets with enhanced charge separation efficiency and CO2 reduction performance
Rangappa A.P., Kumar D.P., Wang J., Do K.H., Kim E., Reddy D.A., Ahn H.S., Kim T.K.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2022
182.
Boosting water decomposition by sulfur vacancies for efficient CO2 photoreduction
Yin S., Zhao X., Jiang E., Yan Y., Zhou P., Huo P.
Energy and Environmental Science, Royal Society of Chemistry (RSC), 2022
184.
Electronic and optical properties of single crystal SnS2: an earth-abundant disulfide photocatalyst
Burton L.A., Whittles T.J., Hesp D., Linhart W.M., Skelton J.M., Hou B., Webster R.F., O'Dowd G., Reece C., Cherns D., Fermin D.J., Veal T.D., Dhanak V.R., Walsh A.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2016
185.
Fabrication of g-C3N4/SnS2 type-II heterojunction for efficient photocatalytic conversion of CO2
Wang H., Liu Z., Wang L., Shou Q., Gao M., Wang H., Nazir A., Huo P.
Journal of Materials Science: Materials in Electronics, Springer Nature, 2023
186.
Enhanced electron–hole separation in SnS2/Au/g-C3N4 embedded structure for efficient CO2 photoreduction
Yin S., Sun L., Zhou Y., Li X., Li J., Song X., Huo P., Wang H., Yan Y.
Chemical Engineering Journal, Elsevier, 2021
189.
Boosted charge transfer and photocatalytic CO2 reduction over sulfur-doped C3N4 porous nanosheets with embedded SnS2-SnO2 nanojunctions
Chen X., Chen Y., Liu X., Wang Q., Li L., Du L., Tian G.
Science China Materials, Springer Nature, 2021
190.
Design of sculptured SnS/g-C3N4 photocatalytic nanostructure for highly efficient and selective CO2 conversion to methane
Omr H.A., Putikam R., Hussien M.K., Sabbah A., Lin T., Chen K., Wu H., Feng S., Lin M., Lee H.
Applied Catalysis B: Environmental, Elsevier, 2023
191.
Synergistic role of Cu-C and Cu-N dual bonding of nanostructured g-C3N4/Cu2SnS3 photocatalysts for efficient CO2 conversion to CO
192.
A review of visible light active SnS photocatalyst for efficient photocatalytic water purification
Talapatadur V., Hegde S.S., Surendra B.S., Murahari P., Ramesh K.
Materials Today: Proceedings, Elsevier, 2023
193.
Two-Dimensional Nanocrystals Produced by Exfoliation of Ti3AlC2
Naguib M., Kurtoglu M., Presser V., Lu J., Niu J., Heon M., Hultman L., Gogotsi Y., Barsoum M.W.
Advanced Materials, Wiley, 2011
194.
Two-Dimensional Transition Metal MXene-Based Photocatalysts for Solar Fuel Generation
Cheng L., Li X., Zhang H., Xiang Q.
Journal of Physical Chemistry Letters, American Chemical Society (ACS), 2019
195.
2D metal carbides and nitrides (MXenes) for energy storage
Anasori B., Lukatskaya M.R., Gogotsi Y.
Nature Reviews Materials, Springer Nature, 2017
196.
Simultaneous Recognition of Dopamine and Uric Acid in the Presence of Ascorbic Acid via an Intercalated MXene/PPy Nanocomposite
You Q., Guo Z., Zhang R., Chang Z., Ge M., Mei Q., Dong W.
Sensors, Multidisciplinary Digital Publishing Institute (MDPI), 2021
197.
Ti3C2 MXene-derived carbon-doped TiO2 coupled with g-C3N4 as the visible-light photocatalysts for photocatalytic H2 generation
Han X., An L., Hu Y., Li Y., Hou C., Wang H., Zhang Q.
Applied Catalysis B: Environmental, Elsevier, 2020
198.
Processing of MAX phases: From synthesis to applications
Gonzalez‐Julian J.
Journal of the American Ceramic Society, Wiley, 2020
199.
Tuning the Work Function of MXene via Surface Functionalization
Koh S.W., Rekhi L., Arramel, Birowosuto M.D., Trinh Q.T., Ge J., Yu W., Wee A.T., Choksi T.S., Li H.
ACS applied materials & interfaces, American Chemical Society (ACS), 2023
200.
Mesoporous g-C3N4/MXene (Ti3C2Tx) heterojunction as a 2D electronic charge transfer for efficient photocatalytic CO2 reduction
202.
Chapter 6 Bentonite Applications
Murray H.H.
Developments in Clay Science, Elsevier, 2006
203.
Photoreduction of CO2 using sol–gel derived titania and titania-supported copper catalysts
Tseng I., Chang W., Wu J.C.
Applied Catalysis B: Environmental, Elsevier, 2002
204.
Role of water and carbonates in photocatalytic transformation of CO2 to CH4 on titania.
Dimitrijevic N.M., Vijayan B.K., Poluektov O.G., Rajh T., Gray K.A., He H., Zapol P.
Journal of the American Chemical Society, American Chemical Society (ACS), 2011
205.
Photocatalytic reduction of CO2 on FeTiO3/TiO2 photocatalyst
Truong Q.D., Liu J., Chung C., Ling Y.
Catalysis Communications, Elsevier, 2012
206.
Photochemical reduction of carbonate to formaldehyde on TiO2 powder
Chandrasekaran K., Thomas J.K.
Chemical Physics Letters, Elsevier, 1983
207.
On the general mechanism of photocatalytic reduction of CO2
Karamian E., Sharifnia S.
Journal of CO2 Utilization, Elsevier, 2016
209.
A hydrofluoric acid-free synthesis of 2D vanadium carbide (V2C) MXene for supercapacitor electrodes
Guan Y., Jiang S., Cong Y., Wang J., Dong Z., Zhang Q., Yuan G., Li Y., Li X.
2D Materials, IOP Publishing, 2020
210.
Production of V2C MXene using a repetitive pattern of V2AlC MAX phase through microwave heating of Al-V2O5-C system
Ghasali E., Orooji Y., Azarniya A., Alizadeh M., Kazem-zad M., TouradjEbadzadeh
Applied Surface Science, Elsevier, 2021
211.
V2C MXene synergistically coupling FeNi LDH nanosheets for boosting oxygen evolution reaction
Chen Y., Yao H., Kong F., Tian H., Meng G., Wang S., Mao X., Cui X., Hou X., Shi J.
Applied Catalysis B: Environmental, Elsevier, 2021
212.
Two-dimensional vanadium carbide (V2C) MXene as electrode for supercapacitors with aqueous electrolytes
Shan Q., Mu X., Alhabeb M., Shuck C.E., Pang D., Zhao X., Chu X., Wei Y., Du F., Chen G., Gogotsi Y., Gao Y., Dall'Agnese Y.
Electrochemistry Communications, Elsevier, 2018
214.
2D/2D Ti3C2 MXene/g-C3N4 nanosheets heterojunction for high efficient CO2 reduction photocatalyst: Dual effects of urea
Yang C., Tan Q., Li Q., Zhou J., Fan J., Li B., Sun J., Lv K.
Applied Catalysis B: Environmental, Elsevier, 2020
215.
Decorating g-C3N4 with alkalinized Ti3C2 MXene for promoted photocatalytic CO2 reduction performance
Tang Q., Sun Z., Deng S., Wang H., Wu Z.
Journal of Colloid and Interface Science, Elsevier, 2020
218.
Construction of Few-Layer Ti3C2 MXene and Boron-Doped g-C3N4 for Enhanced Photocatalytic CO2 Reduction
Wang H., Tang Q., Wu Z.
ACS Sustainable Chemistry and Engineering, American Chemical Society (ACS), 2021
219.
Semiconductor/reduced graphene oxide nanocomposites derived from photocatalytic reactions
Ng Y.H., Iwase A., Bell N.J., Kudo A., Amal R.
Catalysis Today, Elsevier, 2011
220.
Preparation of Graphitic Oxide
Hummers W.S., Offeman R.E.
Journal of the American Chemical Society, American Chemical Society (ACS), 1958
221.
Graphene Oxide (GO) Materials—Applications and Toxicity on Living Organisms and Environment
Ghulam A.N., dos Santos O.A., Hazeem L., Pizzorno Backx B., Bououdina M., Bellucci S.
Journal of Functional Biomaterials, Multidisciplinary Digital Publishing Institute (MDPI), 2022
222.
Boosting Photocatalytic Activity Using Reduced Graphene Oxide (RGO)/Semiconductor Nanocomposites: Issues and Future Scope
Mondal A., Prabhakaran A., Gupta S., Subramanian V.R.
ACS Omega, American Chemical Society (ACS), 2021
223.
The reduction of graphene oxide
Pei S., Cheng H.
Carbon, Elsevier, 2012
224.
2D–1D-2D multi-interfacial-electron transfer scheme enhanced g-C3N4/MWNTs/rGO hybrid composite for accelerating CO2 photoreduction
Li X., Sun B., Wu Q., Fan H., Liu X., Cao J., Yang L., Liu H., Wei M.
Journal of Alloys and Compounds, Elsevier, 2023
227.
Performance analysis of rGO-bridged g-C3N4/ZnV2O6 S-scheme heterojunction for CO2 photoreduction with H2O in an externally reflected photoreactor
Bafaqeer A., Tahir M., Amin N.A., Ummer A.C., Thabit H.A., Dhamodharan D., Ahmed S., Kumar N.
Journal of Alloys and Compounds, Elsevier, 2023
229.
Urchin-like hierarchical CoZnAl-LDH/RGO/g-C3N4 hybrid as a Z-scheme photocatalyst for efficient and selective CO2 reduction
Yang Y., Wu J., Xiao T., Tang Z., Shen J., Li H., Zhou Y., Zou Z.
Applied Catalysis B: Environmental, Elsevier, 2019
230.
Constructed Z-Scheme g-C3N4/Ag3VO4/rGO Photocatalysts with Multi-interfacial Electron-Transfer Paths for High Photoreduction of CO2
Gao M., Sun L., Ma C., Li X., Jiang H., Shen D., Wang H., Huo P.
Inorganic Chemistry, American Chemical Society (ACS), 2021
232.
Indirect Z-Scheme Assembly of 2D ZnV2O6/RGO/g-C3N4 Nanosheets with RGO/pCN as Solid-State Electron Mediators toward Visible-Light-Enhanced CO2 Reduction
Bafaqeer A., Tahir M., Ali Khan A., Saidina Amin N.A.
Industrial & Engineering Chemistry Research, American Chemical Society (ACS), 2019
234.
A review on the recent progress, challenges and perspective of layered double hydroxides as promising photocatalysts
Mohapatra L., Parida K.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2016
235.
2D nanocrystals of metal oxides and hydroxides with nanosheet/nanoflake morphology in biomedicine, energy and chemistry
Tolstoy V.P., Gulina L.B., Meleshko A.A.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2023