Progressive MXene-based photocatalytic and electrocatalytic sustainable reduction of CO2 to chemicals: comprehensive review and future directions

Latiful Kabir a, Karna Wijaya b and Won-Chun Oh *a
aDepartment of Advanced Materials Science & Engineering, Hanseo University, Chungnam 356-706, South Korea. E-mail: wc_oh@hanseo.ac.kr; Fax: +82-41-688-3352; Tel: +82-41-660-1337
bDepartment of Chemistry, Faculty of Mathematics and Natural Sciences, Universitas Gadjah Mada, Yogyakarta, Indonesia

Received 23rd March 2024 , Accepted 1st May 2024

First published on 1st May 2024


Abstract

To reverse the impact of CO2, it is necessary not only to curb the dependence on fossil fuels but also develop effective strategies to capture and utilize CO2 in the atmosphere. However, the solar-to-chemical efficiency for the reduction of CO2 using photocatalysts is not satisfactory in practical applications due to their poor adsorption and activation of CO2, rapid recombination of photogenic charge carriers, and deactivation of their surface sites. Thus, herein, we aim to concisely summarize the state-of-the-art development of 2D-MXene-based photocatalysts for CO2 reduction. First, we briefly introduce the mechanism for the photocatalytic reduction of CO2. Second, the synthesis and properties of MXene composites are summarized together with their role in electrochemical photocatalytic CO2 reduction. Third, various types of MXene-based photocatalysts such as MXene/metal oxide, MXene/nitride, MXene/perovskite, MXene/LDH, and MXene-based photocatalytic CO2 reduction photocatalysts are classified and shown. Finally, we present the related challenges and perspectives on MXene-based photocatalysts.


1. Introduction

Presently, industry relies primarily on fossil fuels and related chemical products. To date, most fossil fuels are used as raw materials and fuels to drive conversion and separation processes. Consequently, these industries are the main sources of CO2 and other greenhouse gases. Moreover, the significant increase in the concentration levels of greenhouse gases in the atmosphere, such as CO2, has raised serious concerns about the fossil fuel-based energy supply. In the late 19th century (Fig. 1) the concentration of CO2 in the atmosphere increased from 280 to 400 ppm.1–7
image file: d4se00405a-f1.tif
Fig. 1 Number of publications related to (a) photocatalytic CO2 reduction published since 2000. (b) Photocatalytic CO2 reduction and published papers related to MXenes since 2017.5 The data was collected from the Web of Science.

As an example of an essential process in maintaining the carbon/oxygen cycle, photosynthesis in green plants is a natural phenomenon necessary for the maintenance of life on Earth. It consists of two sequential steps, which are known as the light and dark reactions (Fig. 2a). In the light reaction, chlorophyll reacts with sunlight to form adenosine triphosphate (ATP), which oxidizes water to O2. Alternatively, in the dark reaction, CO2 is gradually reduced to form carbohydrates using the energy stored in ATP. Consequently, carbon dioxide essentially provides the energy needed for most life on Earth via natural photosynthesis and is the basis for human survival.8–16 In this case, many researchers have consistently attempted to convert atmospheric CO2 and water into chemical fuels using the sun as the energy source. As schematically illustrated in Fig. 2b, the first path may generate a sufficient photo voltage using a photocell (PV), and then an electrochemical reaction occurs at the anode for CO2 reduction and the cathode for water oxidation. In this case, the study of a new electrode system is required to accelerate the electrocatalytic reaction and improve the reaction selectivity. Different components can be optimized individually and combined in the electrode system to enable the best overall performance. The second route involves the use of interfacial charge transfer effects to induce light harvesting, charge separation, and reaction within particles by dispersing light-absorbing semiconductor particles (photocatalyst) decorated with a suitable electrical cocatalyst (photocatalyst) in aqueous solutions (Fig. 2c). The advantage of this path is that it facilitates a much simpler and smaller device design.17–20 Future success is expected to heavily rely on the development of high-performance carbon dioxide reduction electrocatalysts or photocatalysts.


image file: d4se00405a-f2.tif
Fig. 2 Comparison of (a) natural photosynthesis, (b) electrochemical synthesis on electrocatalysts in a photovoltaic cell, and (c) photochemical synthesis on powered photocatalysts.20 CO2 is one of the most thermodynamically stable molecules due to its strong C–O double bond with a binding energy of 750 kJ mol−1, which is significantly greater than the binding energy of C–C (336 kJ mol−1), C–O (327 kJ mol−1), and C–H (411 kJ mol−1) bonds. Thermodynamically, the reduction of CO2 through an electrochemical catalyst or photocatalytic approach is an endothermic reaction, which requires significant energy input to break the CO bond. Carbon dioxide reduction can proceed through various reaction paths with the movement of 2, 4, 6, 8, 12 or more electrons and produce various reduction products such as carbon monoxide (CO), formic acid (HCOOH), methanol (CH3OH), methane (CH4), and ethylene (C2H4).21

According to the studies reported thus far, various technologies such as electrochemical,22 thermochemical,23 biological and photoelectrochemical methods24,25 have been employ to realize CO2 reduction. Among them, the reduction of CO2 using solar photocatalysts has attracted significant attention due to their potential and synergistic effects to simultaneously solve the greenhouse effect and energy shortage (Fig. 1a).26 CO2 reduction by photocatalysts has several advantages, as follows: (i) the light energy source used in the CO2 reduction process by photocatalysts is the clean and abundant solar power, (ii) CO2 reduction by photocatalysts can be performed under mild conditions (e.g., atmospheric pressure and room temperature) and (iii) renewable hydrocarbon fuels such as CH4, CH3OH, and C2H6 may be converted directly from CO2, thereby realizing carbon fixation and recycling.27 Therefore, the reduction of CO2 to hydrocarbon fuel by photocatalysts can realize the effect of killing two birds with one stone in terms of energy supply and environmental protection.

However, the reduction of CO2 by photocatalysts limited by the poor adsorption of CO2 on the catalyst surface, rapid charge recombination, and deactivation of the surface area. Due to these shortcomings, the photocatalytic effect for CO2 reduction by photocatalysts is not significant, and the related research is in its infancy and requires continuous research. Therefore, there is a big difference between research results and practical application. It is also urgent to classify published studies on MXene-functionalized catalysts for photocatalytic CO2 reduction and to discover the relationship between MXenes and photocatalytic activity. Many reviews partially introduced MXene-based materials for the reduction of CO2 when summarizing their applications in other photocatalytic applications and energy conversion.28 Thus, a logical overview is also needed to highlight the latest developments in MXene-based materials for photocatalytic CO2 reduction. Therefore, herein, we aim to describe the latest research on MXene-functionalized photocatalysts applied in CO2 reduction and their development. Firstly, we present a brief overview of the photocatalytic reduction mechanism of CO2. This is followed by a comprehensive summary of the synthesis and properties of MXenes. Secondly, the pivotal roles played by MXenes in photocatalytic CO2 reduction, such as driving the adsorption of CO2, strengthening the separation of photoinduced charge carriers, acting as a strong support, and photothermal effects, are summarized. Thirdly, MXene-based complex photocatalysts will be classified in various categories and described. Considering the rapid development of the Ti3C2Tx MXene as a catalyst and its favorable availability, a comprehensive review of photocatalysts and electrocatalysts based on Ti3C2Tx MXene is necessary. Thus, we also present a comprehensive review of the latest developments on its related studies in this review, highlighting the essential role of Ti3C2Tx MXene in the reduction of CO2. Finally, we present a summary with future perspectives related to MXene-combined photocatalysts.

2. Reduction method

The three main approaches for reducing the amount of CO2 in the atmosphere are as follows: (1) direct reduction of CO2 emissions by reducing the use of fossil fuels, (2) methods for collecting and storing CO2 by new technologies (CCS), and (3) efficient utilization and conversion of CO2. Typically, various technologies, including thermodynamic, biological, photoelectric catalysts, electrocatalysts, and photocatalytic reduction reactions, have been used to convert carbon dioxide into hydrocarbon fuels. Among them, the use of electrocatalysts and photocatalytic methods are widely studied for CO2 reduction, which both have unique advantages.29 Semiconductor photocatalysts that mimic photochemical reduction and plant photosynthesis processes by combining H2O with solar light can realize products of satisfactory purity through the potential scalable production of CH3OH, C2H5OH, and hydrocarbon compounds called solar fuels (Fig. 3). To date, several methods have been used to reduce CO2, among which photocatalyst and electrochemical reduction methods have been widely used in research investigating CO2 conversion.30 The photocatalytic method has the advantage of existing low energy technology and it can efficiently convert and store solar energy into chemical energy, while allowing carbon to be recycled. The entire process of CO2 reduction by photocatalysts is carried out by light absorption, charge separation, CO2 adsorption to the photocatalytic surface, surface redox reaction, and product desorption.
image file: d4se00405a-f3.tif
Fig. 3 (a) Schematic energy change diagram for photocatalytic CO2 reduction and H2O oxidation on a photocatalyst semiconductor. (b) Schematic of the photocatalytic reaction procedure, illustrating the factors that may impact the photocatalytic activity and performance.21

The electrochemical reduction potentials for various products of CO2 reduction at pH 7 are presented in Table 1. The CO2 reduction reaction by a single electron requires a negative potential of −1.9 eV, which creates a reduction process by one electron. Alternatively, the CO2 reduction reaction by proton-assisted multiple electrons requires a relatively low redox potential (Table 1). Photocatalysts may facilitate the reduction process by forming a potential. For this purpose, ideal photocatalysts generally require two properties. Firstly, the redox potential of the photoexcited VB hole must be sufficiently positive to enable the hole to serve as an electron receptor. Secondly, the redox potential of photoexcited CB electrons should be more negative than that of the oxidation–reduction couple, which is a CO2/reduction product.29 As shown in Fig. 3, the CO2 photo-reduction process can be largely divided into three stages. CO2 and a reducing agent (taking H2O) are physically reacted on the surface of the photocatalyst. The dissociation energy by the combination of C and O is 750 kJ mol−1.31 Thus, it is very important in the reduction reaction to adsorb CO2 and have activity. Light plays an important role in the formation of the bandgap energy in photocatalysts as an energy source, and the electrons in the valence band, VB, induce an excited state to the conduction band, CB, to create holes in VB. The interrelationship between these two bands is that the electrons and holes generated by light exhibit activity and move to the surface of the photocatalyst. Unfortunately, most electron–hole pairs combine before reaching the surface, significantly lowering the reduction efficiency of carbon dioxide due to the effect of photocatalysts. The adsorbed CO2 and H2O strongly react with the holes and electrons on the surface of the photocatalyst to produce CO, CH4, CH3OH, HCOOH, and O2. As shown in Table 1, some products are generated via different paths, and therefore require different potentials.32

Table 1 Reduction potentials for CO2 reduction.32
Reaction E 0 (V) vs. NHE at pH = 7
Reduction potentials of CO 2
2H+ + 2e → H2 −0.41
CO2 + e− → CO2 −1.9
CO2 + 2H+ + 2e → HCO2H −0.61
CO2 + 2H+ + 2e → CO + H2O −0.53
CO2 + 4H+ + 4e → C + 2H2O −0.2
CO2 + 4H+ + 4e → HCHO + H2O −0.48
CO2 + 6H+ + 6e → CH3OH + H2O −0.38
CO2 + 8H+ + 8e → CH4 + 2H2O −0.24
2CO2 + 8H2O + 12e → C2H4 + 12OH −0.34
2CO2 + 9H2O + 12e → C2H5OH + 12OH −0.33
3CO2 + 13H2O + 18e → C3H7OH + 18OH −0.32
[thin space (1/6-em)]
Reduction potentials of H 2 CO 3
2H2CO3 + 2H+ + 2e → H2C2O4 + 2H2O −0.8
H2CO3 + 2H+ + 2e → HCOOH + H2O −0.576
H2CO3 + 4H+ + 4e → HCHO + 2H2O −0.46
H2CO3 + 6H+ + 6e → CH3OH + 2H2O −0.366
H2CO3 + 4H+ + 4e → C + 3H2O −0.182
[thin space (1/6-em)]
Reduction potentials of CO 3 2−
2CO32− + 4H+ + 2e → C2O42− + 2H2O 0.07
CO32− + 3H+ + 2e → HCOO + H2O −0.099
CO32− + 6H+ + 4e → HCHO + 2H2O −0.213
CO32− + 8H+ + 6e → CH3OH + 2H2O −0.201
CO32− + 6H+ + 4e → C + 3H2O 0.065


Accordingly, to improve the CO2 reduction activity of photocatalysts, it is necessary to expand their light absorption range by co-catalyst doping, surface modification, and morphology control to generate more optical carriers. In addition, it is necessary to improve the separation and transmission efficiency of the optical carriers and extend the life of the charge carriers. Finally, more surface-active sites must be created for photocatalytic redox reactions.

Generally, the efficiency of CO2 reduction by a photocatalyst is measured by the yield of the product. Here, the general unit for reduction amount is mol h−1 g−1 of the catalyst, and the molar unit (μmol) or concentration unit (ppm) of the product. In photocatalytic-based measurements, the efficiency of photocatalysts usually depends on the amount of photocatalyst, the intensity of light, and the area exposed to light. As a result, the amount of products produced per gram of photocatalyst used within a certain period under light irradiation can be calculated as the apparent quantum yield. It is calculated using the reported equations33,34 based on the amount of product and the number of incident photons. When the photocatalytic reduction reaction creates a complex product, the number of electrons reacted in the equation means the sum of the electrons reacted, resulting in each product.35,36 Thus, in photocatalytic-based measurements, the quantum yield of the photocatalytic reduction of CO2 to different products can be calculated using the following equations:

R = n (product)/time × m (catalyst)1
 
image file: d4se00405a-t1.tif(2)
 
image file: d4se00405a-t2.tif(3)
 
image file: d4se00405a-t3.tif(4)
 
image file: d4se00405a-t4.tif(5)
 
image file: d4se00405a-t5.tif(6)
 
image file: d4se00405a-t6.tif(7)
 
image file: d4se00405a-t7.tif(8)

Moreover, the photoelectrochemical CO2 reduction method has several operating parameters. Thus, additional research on scale-up, cell design, appropriate photocatalyst design, membrane research, light sources and reaction conditions for system integration is needed. In addition, eco-friendly renewable energy can be considered an electric energy source. The overall plan is summarized in Fig. 4, where an electrochemical reactor, such as an electrolyzer, is powered by electricity generated from a renewable source and drives an energy-boosting process that converts CO2 and H2O into electric fuels.37 The fuel generated by photoelectrochemical reduction reactions can be easily used to power transportation and other human activities through storage, distribution and existing infrastructure. The CO2 and H2O generated during the consumption of the generated fuel need to be captured and fed back to the reactor.


image file: d4se00405a-f4.tif
Fig. 4 Overall schematic of photoelectrochemical CO2 reduction method with several operating parameters.

The key step in converting CO2 into a carbonized water compound by the electrochemical method is to chemically convert CO2 molecules into reduced carbon species, which is considered a difficult process due to the poor dynamics of CO2 electrical reduction. It is possible to present a method to solve this problem by electrochemical reduction using a photocatalytic electrode. Electrochemical CO2 reduction is evaluated as a more attractive method because it has several characteristic advantages over other approaches.38–41 For example, by controlling the external element (i.e., overpotential), which is an electrical element, it is possible to perform a major reduction reaction under ambient conditions, where the reaction rate is easily controlled. These products are produced by different electrodes and can be separated naturally using individual reaction units, which is one of the main factors to minimize the cost associated with the separation process after the reduction reaction. As illustrated in Fig. 5, in a typical CO2 photoelectrolytic cell, a cathode and an anode are placed in two chambers separated by ion conductivity membranes. At the anode, water is oxidized to molecular oxygen, while at the anode, CO2 is reduced to a reduced piece of carbon, initiating the primary reduction reaction (Fig. 4 and 5). Fig. 4 describes the ideal commercial process for CO2 reduction by the photo-catalyst method. It presents the future vision together with the problems currently faced in this process. In the case of the ideal reduction process, it is intended to scale it up in a small laboratory-scale electrolyzer, derive the optimal CO2 reduction conditions in the related processes, and present a sketch to create the optimal field scale-up data based on them.


image file: d4se00405a-f5.tif
Fig. 5 Schematic diagram of an ion-conducting membrane-functionalized CO2 electrolyte system.41

3. Synthesis of different photocatalysts

3.1 Traditional photocatalysts

The proposed electrochemical conditions for the photocatalytic and electrochemical reduction of CO2, various catalyst materials, and various strategies to improve the activity of these catalysts have been studied. One of the important factors for enhancing the photocatalytic activity in the reduction of CO2 is the selection of an appropriate photocatalyst. In this case, understanding both the practical application of photocatalysts and their mechanisms is an important topic. Photocatalysts can be classified into two basic groups, i.e., homogeneous and heterogeneous photocatalysts, depending on their chemical structure. Fig. 6 shows the results of a study on the CO2 reduction by photocatalysts to obtain good efficiency and selectivity for specific products. The approach of using photocatalysts still has many factors to consider for its practical implementation. Consequently, there are several factors limiting the alleviation of environmental pollution and large-scale application of photocatalysts in the energy sector. Two of the main things to consider in the use of heterogeneous photocatalysts are the low photocatalytic efficiency and lack of suitable visible light response photocatalysts.42,43 Also, low bandgap energy and cost efficiency must be considered to solve these problems. Another problem to be solved is the decrease in photocatalytic efficiency due to the rapid recombination of photo-generated electrons and holes.
image file: d4se00405a-f6.tif
Fig. 6 Schematic representation of band gap energy locations and band gap energies of various semiconductors and relative redox potentials of the compounds involved in CO2 reduction at pH 7.42

3.2 Synthesis of MXenes

Generally, MXenes are synthesized through the top-down method by etching the “A” layer from the metal interlayer compound MAX.44 Most of the MAX phases are layered ternary carbides and nitrides.142 The general formula for MAX is represented by Mn+1AXn, where M is an early transition metal, A is a group 11–16 element (e.g., Si, Al and Ga), X is nitrogen or carbon, and n = 1–4.45 To date, more than 150 types of basic MAX phases have been successfully synthesized and studied.46 This number is expected to increase in the future due to the diverse composition of various elements and numerous possibilities for precursor selection for MXenes.46 When the MAX phase is placed in a fluorine-containing aqueous solution, the relatively weak M–A bond (or M–X bond) is broken by the etchant, leading to the successful elimination of the A atomic layer.47 It is noteworthy that M, A, X and n produce different MAX phases with varying bond strengths. To synthesize MAX and successfully transform it into MXene, a high concentration of acid is required, and to obtain high-quality MXene, the deterioration of the 2D flakes can be minimized with the use of acid. In this case, changing the etching conditions (acid concentration, etching temperature, and etching time) according to the different types of MAX is a very important factor.48 After the A atomic layer is etched, a small number and a large number of MXene layers are generated. Given that most synthesis processes are performed in HF or F ion-containing aqueous solutions, functional groups such as –OH, –O and –F are created on the transition metal surface of MXenes.49

The chemical formula of MXenes is Mn+1XnTx (n = 1–4), where Tx represents edge functional groups such as H and halogen elements. Over the last decade, there has been a series of breakthroughs in the preparation of MXenes. Ti3C2, the first MXene reported in 2011, was synthesized by immersing Ti3AlC2 MAX in 50% HF at room temperature.50 Subsequently, Ti2AlC, Ta4AlC3, (Ti0.5Nb0.5)2AlC, (V0.5Cr0.5)3AlC2, and Ti3AlCN were successfully etched in HF solution in 2012.50 After dimethyl sulfoxide (DMSO) was inserted into Ti3C2, and then sonicated, single-layer Ti3C2 was synthesized.51 In 2014, the etchant was replaced by NH4HF2 (ref. 52) and LiF/HF53 solutions with about 5% formation of HF, reducing the risk associated with the use of high-concentration HF. In 2015, the large-scale delamination of multilayer MXene was achieved using an organic-based solution instead of DMSO.54 In 2016, Yury's group produced a larger single Ti3C2Tx flake with fewer defects using the minimum concentration layer stripping (MILD) method.55 In addition, ultrasonic treatment was not required, and multilayer Ti3C2Tx could be separated by hand shaking. In 2017, nitride MXene (Mo2N and V2N) was obtained by ammonium treatment of Mo2CTX and V2CTX MXene at 600 °C. In 2018, it was reported that the C atom of MXene was replaced by an N atom.56

Subsequently, in 2020, Huang's group synthesized various MXenes using non-traditional MAX precursors with Ga, Si, and Zn elements via Lewis acid etching pathways.57 Recently, many new synthetic protocols have emerged in addition to the above-mentioned methods for the Ti3AlC2 MAX using LiF in aqueous solution or with ultrasonication. For example, a thermal reduction strategy was employed to reduce the Ti2SC MAX phase to Ti2C MXene.58 Ti2C MXene was obtained by the electrochemical etching method from Ti2AlC in HCl solution.59 Another unique new method is the alkali metal-based hydrothermal and HF-free pathway, thereby obtaining high purity (92 wt%) Ti3C2Tx.60 Feng et al. synthesized Ti3C2 MXene from Ti3AlC2 by an iodine-based method in anhydrous acetonitrile.61Fig. 7 presents a simple method for the synthesis of MXene with the analysis results.


image file: d4se00405a-f7.tif
Fig. 7 Simple MXene preparation procedure with some analysis data (XRD, SEM (a–f) Raman, adsorption–desorption isotherms and pore size distributions).46

3.3 Synthesis and heterojunction of MXene-based semiconductors

3.3.1 Typical hydrothermal method. The hydrothermal method refers to the synthesis of a material through chemical reactions in a sealed and heated solution above ambient pressure and temperature. Hydrothermal methods have many advantages, such as relatively mild operating temperature conditions, one-step synthesis, environmentally friendly processes, and good diffusion in solution. In addition, hydrothermal preparation is more cost effective in terms of energy, instruments and material precursors compared to other solution preparation techniques. Many hydrothermal processes are performed in autoclaves maintained at high temperatures and pressures, making it possible to manufacture heterocomposites with high crystallinity and small particle size. As a result, this method enables the synthesis of MXene-semiconductor-based photocatalysts without affecting the performance of the synthesis process. An SrTiO3/Ti3C2 hybrid heterojunction composite was prepared through a hydrothermal process, as shown in Fig. 7. In general, 1 mL of titanium isopropoxide (TIP, C12H28O4Ti) was dissolved in 50 mL of acetonitrile and ethanol solution (2[thin space (1/6-em)]:[thin space (1/6-em)]3) and strongly stirred with NH3. H2O (0.38 g) and H2O (0.91 g) for 6 h. Subsequently, the resulting solution was centrifuged to obtain the derived TiO2, which was washed many times with deionized water and ethyl alcohol, and heated at 60 °C for 24 h. After self-stirring for 15 min, the obtained product was placed in a 50 mL Teflon-lined stainless steel autoclave and heated at 140 °C for 4 h. The resulting samples were washed and heated at 60 °C.62,63 Gao and coworkers effectively manufactured Ti3C2/TiO2 nanohybrids with high photocatalytic performance through a hydrothermal process for the decomposition of methyl orange dye under ultraviolet lighting.64 Xiao Lu and coworkers synthesized Ti3C2Tx/TiO2/PANI heterostructured compounds through hydrothermal methods together with in situ polymerization technology.65 Zhou et al. reported the effective synthesis of a Ti3C2/CeO2 nanohybrid through one-pot hydrothermal technology with well-distributed CeO2 nanorods on Ti3C2 2D nanosheets.66 Tian et al. logically reported the synthesis of a new type of UiO-66-NH2/Ti3C2/TiO2 ternary composite through Ti3C2Tx MXenes induced by the introduction of stable Zr-MOF precursors in water via a hydrothermal process.67 In addition, Fang et al. reported the synthesis of a three-way CdS/Ti3C2–OH/ln2S3 ternary composite photocatalyst with effective visible light-driving photocatalytic performance via easy hydrothermal technology.68 Fang et al. also described that the Ti3C2OH/Bi2WO6:Yb3+, Tm3+ photocatalyst was effectively synthesized via an easy hydrothermal method.69 Li et al. studied the design motif of the in situ growth grafting of Ti3C2 MXene and MoS2 2D nanosheets on the (101) plane of TiO2 nanosheets, where most of the highly active (001) planes were exposed by hydrothermal technology. Also, Li et al. reported the synthesis of a BiOBr/Ti3C2 photocatalyst via an easy hydrothermal technique.70
3.3.2 Organo-solvothermal technique. The solvothermal method is similar to hydrothermal technology, except that organosolvents are used instead of water in the synthesis procedure. In addition, when alcohol and glycerol are used as the reaction intermediates, the reaction is called alcoholthermal and glycothermal, respectively. This synthesis approach is useful for the preparation of nanocomposites with excellent crystalline properties. The solvothermal process is one of the most typical and capable synthesis processes for constructing MXene-based semiconductor-combined photocatalysts with multiple morphological structures. Of course, even in this process, an autoclave is filled with organosolvents to react under ambient reaction conditions. Hao et al. effectively synthesized MXene/TiO2/NiFeCo ternary hybrid composites through easy solvothermal technology for oxygen generation reactions.71–73 Tie et al. synthesized ZnS nanoparticle-decorated Ti3C2 MXene sheets via ultrasonication and solvothermal process toward improved photocatalytic hydrogen generation,74 which was about four times larger than that of bare ZnS (124.6 mmol g−1 h−1). This procedure improved the overall clean energy system as a possible candidate for H2 production using an MXene/ZnS photocatalyst and provides new possibilities for the further expansion of the water-phase applications of MXene-based samples. The particle size, phase composition, and particle shape of ceramic powder can be confirmed during the calcination procedure given that the product is a precursor material designed in most solution synthesis processes. Under temperature control during the calcination process and a controlled atmosphere, a Ti3C2/TiO2 nanohybrid could be synthesized via simple calcination technology. The synthesized Ti3C2 nanosheet was heated at 80 °C for 12 h. Thereafter, the heated Ti3C2 was heated at a heating rate of 10 °C min−1 to 350 °C, 450 °C, 550 °C, and 650 °C to obtain TT350, TT450, TT550 and TT650, respectively.46 In another study, Li et al. produced a Ti3C2/TiO2 photocatalyst via the primary calcination method, where 0.3 g of F–Ti3C2 was heated for 4 h under an air atmosphere at a temperature of 550 °C.75 Ding et al. proposed the synthesis of Ti3C2/TiO2/g-C3N4 heterostructures through ultrasonic auxiliary calcination technology for the decomposition of organic pollutants.76 In addition, Ti3C2/TiO2/g-C3N4 heterogeneous structures were synthesized by combining Ti3C2 and TiO2 with O/OH functional groups, which can serve as a suitable support by transferring abundant electrons. As a result, it was suggested that the high photocatalytic activity for the removal of aniline and rhodamine B was improved by 5 times and 1.33 times compared to g-C3N4, respectively, under visible light irradiation.
3.3.3 Simple sol–gel procedure. The production of silica by the sol–gel method was first proposed by French chemist M. Evelmen in 1846. He recognized that silica esters are gradually hydrolyzed in the presence of water vapor to produce hydrated silica. Subsequently, the actual application of the sol–gel process began in the middle of 20th century when silica coatings were manufactured on glass by Schott Glaswerke. Nevertheless, many studies on the sol–gel method were presented much later, and the first international workshop was proposed in 1981 for its application to glass and ceramics using gel.77 The sol–gel method has become one of the common solid synthesis procedures for preparing new metal oxide samples and mixed oxide nanomaterials, playing an important role in determining the texture and surface characteristics of nanocomposites. In addition, the sol–gel method has been used to synthesize MXene-based semiconductor photocatalysts with high purity and homogeneity. Initially, the Ti3C2 MXene material was synthesized using distilled water at a molar ratio of 0.5 mg mL−1, and then sonicated for 10 min. Subsequently, Co-doped BFO NPs were added to the solution at a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio of CH3COOH and ethylene glycol (C2H6O2) and a molar ratio of 0.01 M. The Bi1−xLaxFe1−yMnyO3 sample was sonicated at 60 °C for 1 h. Thereafter, the Bi1−xLaxFe1−yMnyO3 sample was individually dissolved in a Ti3C2 solution for all the hybrids, and then the Ti3C2 MXene/Bi1−xLaxFe1−yMnyO3 solution was stirred at 80 °C for 2 h for the co-precipitation process.78
3.3.4 Ion exchange procedure. Ion exchange was used in the study of inorganic substances that can be ‘base’ exchanged, and in 1858, C. H. Eichorn demonstrated that natural zeolite minerals could reversibly exchange cations.79 The importance of these properties in water softening was recognized at the beginning of the century by H. Gans, who patented a series of synthetic amorphous aluminosilicates for this purpose. These materials have not only been used for nuclear waste treatment, but also widely used until recently to soften industrial and household water supplies. Furthermore, ion exchange pathways involve replacing ions in ion particles, while the framework residue caused by exposing the parent ion particles to innovative ions remains intact.79 Ion exchange occurring in nanoparticles compared to bulk samples can be completed quickly at a specific temperature. In addition, when the ion exchange reaction proceeds, the reaction that changes the reaction environment to produce subsequent products occurs quickly. Y. Li et al. attempted to combine a TiO2/Ti3C2 binary hybrid from layered MXene by simultaneously performing oxidation and alkalization after ion exchange and calcination processes. To manufacture the TiO2/Ti3C2 binary hybrid, 120 mg of Ti3C2 MXene sheets was dissolved in solutions with 1 M NaOH (180 mL) and 30% H2O2 (3.6 mL). Subsequently, the obtained final solution was transferred to a Teflon-lined autoclave and heated at the appropriate temperature for 12 h. The produced Ti3C2/Na2Ti3O7 photocatalyst was rinsed several times with deionized water and dried in a vacuum oven at 60 °C for 12 h. Then, a Ti3C2/H2Ti3O7 photocatalyst was prepared by adding the Ti3C2/Na2Ti3O7 photocatalyst to an HCl solution for 24 h. Finally, the TiO2/Ti3C2 hybrid was pyrolyzed for 3 h in a muffle furnace (TiO2/Ti3C2-300, TiO2/Ti3C2-400 and TiO2/Ti3C2 500) at various temperatures (300 °C, 400 °C and 500 °C), respectively.80
3.3.5 Novel self-assembly method (SAM). SAM of nanoscale component molecules exists in the natural world given that soft materials are assembled to produce cell membranes and biomolecule fibers. SAM is a procedure in which nanoscale materials voluntarily arrange predefined components into an ordered superstructure, which can be utilized in different applications. SAM is carried out in a procedure in which molecules are arranged in a controlled structure of separated samples to reduce the free energy of the entire reaction system. Electrostatic SAM is an incredibly well-known simple approach for the synthesis of heterojunction composites. Due to its smooth process environment, this self-assembly method not only reinforces the management of the morphology of the product system, but also promises a narrow size distribution. For example, Bi(NO3)·5H2O was typically added to CH3COOH (solution A). Thereafter, solution A was stirred vigorously in a specific quantity of Ti3C2 solution for 2 h. Thereafter, NaBr was added to distilled water (solution B). The resultant was washed three times with ethanol and distilled water, respectively, and then the obtained product was heated at 60 °C for 12 h.81 In another report on the production of MXenes, Zhang et al. synthesized 2D/2D nanocomposites in the form of a Ti3C2 MXene/α-Fe2O3 hybrids via ultrasonic-assisted self-assembly technology and suggested that these structures form an ideal structure for the photolysis of contaminants by forming uniform 2D α-Fe2O3 sheets and Ti3C2 MXene layers.82 Cai et al. synthesized a Ti3C2/Ag3PO4 Schottky photocatalyst and suggested that Ti3C2 could greatly improve the photocatalytic activity and strength of Ag3PO4.83 Yang et al. reported the fabrication of g-C3N4/Ti3C2 nanocomposites via easy electrostatic SAM for photocatalytic H2O2 evolution under visible light irradiation.84 Liu et al. synthesized Ti3C2/g-C3N4 hybrid nanocomposites through new evaporation-induced self-assembly technology, which were used for the photolysis of ciprofloxacin.85 Zhang et al. described the synthesis of polymorphic nanocomposites such as Ti3C2 MXene/α-Fe2O3/ZnFe2O4via ultrasonic-assisted self-assembly strategies to uniformly disperse magnetic α-Fe2O3/ZnFe2O4 heterostructures on the surface of Ti3C2 MXene.86 Zhuang et al. suggested that the new 2D/1D photocatalyst of Ti3C2/TiO2 was effectively synthesized through electrostatic SAM.87 In addition, Huang et al.88 reported that the Ti3C2 MXene/Bi2WO6 nanocomposite effectively decomposes acetone and formaldehyde, which was also synthesized via electrostatic self-assembly technology. Fang and co-authors effectively produced Ti3C2/Ag2WO4 photocatalysts via this method and suggested that the presence of conductive Ti3C2 greatly improved the photocatalytic results and corrosion resistance of Ag2WO4.89 Lin et al. reported that after O-functionalized g-C3N4 nanoparticles were prepared on the surface of a 2D structure by annealing, the 2D-2D O-functionalized g-C3N4/Ti3C2 MXene Schottky-junction was produced by the in situ electrostatic magnetic assembly of the positively charged O-functionalized g-C3N4 nanoparticles and negatively charged Ti3C2 MXene.90

4. Photocatalytic CO2 reduction

Photocatalytic CO2 reduction is an effective method considering that it does not require additional energy and does not produce negative environmental effects. In this case, the use of natural and abundant sunlight to transform major greenhouse gases into other carbon-containing products is also an ideal approach to solve the issue of high cost. Specifically, the high activation energy for breaking down chemically very stable CO2 molecules is provided by solar energy and photocatalysts.87 To date, many photocatalysts, including oxides and non-oxides, for example, simple oxides and sulfides such as TiO2, ZnO, Fe2O3, ZrO2, SnO2, BiWO3, CdS, and ZnS, and various types of TNTs, GaN Ti-MCM-41, and SiC, have been studied for the photocatalytic reaction of CO2 with H2O. Table 2 summarizes the various photocatalytic systems used for the decomposition of CO2 since 2010.
Table 2 Advances in photocatalytic systems for CO2 reduction with water since 2010
Photocatalyst Radiation source Major products Comments Reference
0.5 wt% Cu/TiO2–SiO2 Xe lamp (2.4 mW cm−2, 250–400 nm) CO and CH4 The synergistic combination of Cu deposition and high surface area of SiO2 support enhanced CO2 photoreduction rates 91
ZnGa2O4 300 W Xe arc lamp CH4 Strong gas adsorption and large specific surface area of the mesoporous ZnGa2O4 photocatalyst contribute to its high photocatalytic activity for converting CO2 into CH4 92
(RuO + Pt)–Zn2GeO4 300 W Xe arc lamp CH4 In the presence of water, ultra-long and ultrathin geometry of the Zn2GeO4 nano-ribbon promotes CO2 photo-reduction, which was significantly enhanced by loading Pt or RuO2 93
Ag/ALa4Ti4O15 (A = Ca, Ba and Sr) 400 W Hg lamp CO, HCOOH, and H2 On the optimized Ag/BaLa4Ti4O15 photocatalyst, CO was the reported as the main product. The molar ratio of O2 production (H2 + CO:O2 = 2[thin space (1/6-em)]:[thin space (1/6-em)]1) demonstrated that water was consumed as a reducing reagent in the photocatalytic process 94
I–TiO2 nanoparticles 450 W Xe lamp CO High photocatalytic activity was observed under visible light and the efficiency of CO2 photoreaction was much greater than undoped TiO2 due to the extension in the absorption spectrum of TiO2 to the visible light region and facilitated charge separation 95
LiNbO3 Natural sunlight or Hg lamp (64.2 mW cm−2) HCOOH MgO-doped LiNbO3 showed an energy conversion efficiency rate of 0.72%, which was lower than that for the gas–solid catalytic reaction of LiNbO3 (2.2%) 96
G-Ti0.91O2 hollow spheres 300 W Xe arc lamp CH4, CO The presence of G nanosheets compactly stacking with Ti0.91O2 nanosheets allows the rapid migration of photo-generated electrons from Ti0.91O2 nanosheets into G and improves the efficiency of the photocatalytic process 97
ZnIn2S4–In2O3 hierarchical tubular 300 W Xe arc lamp CO The optimized ZnIn2S4–In2O3 photocatalyst exhibits outstanding performance for reductive CO2 deoxygenation with considerable CO generation rate (3075 μmol h−1 g−1) and high stability 98
Triazine-based covalent organic framework (SAS/Tr-COF) backbone Visible light CO Remarkably, the as-synthesized Fe SAS/Tr-COF as a representative catalyst achieved an impressive CO generation rate as high as 980.3 μmol g−1 h−1 and selectivity of 96.4% 99


Fig. 8a shows the advantages of H2O oxidation on a metal complex catalyst (H2O oxidation site) with a sacrificial electron acceptor (SA). Fig. 8b shows the advantage of CO2 reduction on a metal complex catalyst (CO2 reduction site) having a sacrificial electron donor (SD). Fig. 8c shows a problem encountered when combining the H2O oxidation site and the CO2 reduction site. The problem that appears in these processes is the reverse oxidation of the products such as organic compounds. In addition, electron movement from the H2O oxidation site to the CO2 reduction site occurs and activation occurs in H2O only when electron storage is performed. In these processes, O2 reduction occurs more easily than CO2. Together with H2O oxidation, there are many problems in constructing homogeneous metal complex systems for CO2 reduction. The inefficient electron transport between the reduction and oxidation catalysts is one of the major difficulties in this process. Another drawback is the short lifetimes of the one-electron-reduced species and the photo-excited state in the presence of O2 generated by H2O oxidation.


image file: d4se00405a-f8.tif
Fig. 8 Advantages and disadvantages of metal complex catalysts for CO2 reduction with H2O oxidation (Adapted from ref. 8). (a) Advantages of H2O oxidation on a metal complex catalyst (H2O oxidation site) with a sacrificial electron acceptor (SA); (b) advantages of CO2 reduction on a metal complex catalyst (CO2 reduction site) with a sacrificial electron donor (SD); and (c) problems encountered when combining H2O oxidation site and CO2 reduction site.

Table 4 presents a summary of the data for catalyst materials used in electrochemical CO2 reduction.

Most of the existing electrocatalysts for CO2 reduction can be classified into three groups, i.e., metal, non-metal, and molecular catalysts. Based on the primary CO2 reduction products, monometallic catalysts can be classified into several subgroups for selective reduction of CO2 (e.g., Au, Ag, and Zn), selective reduction of formic acid (e.g., Sn, In, and Pb), and selective reduction of hydrogen (e.g., Fe, Ni, and Pt). Among the monometallic catalysts, Cu has distinct catalytic properties, producing a wide range of CO2 reduction products, including CO, formic acid, ethanol, and ethylene (Fig. 9).127


image file: d4se00405a-f9.tif
Fig. 9 Outline of major categories of electrocatalytic CO2 reduction.127

5. Photocatalytic results and their discussion

In this section, we review the results and discuss photocatalytic and electrochemical CO2 reduction. Firstly, the adsorption and activation of CO2 on the catalyst surface must occur as the initial reaction. The activation of adsorbed CO2 molecules is an important and very challenging step for CO2 reduction. CO2 adsorption and activation have a significant impact on the inhibition of the subsequent reduction and competitive hydrogen evolution reaction (HER) stages. Adsorption interactions with surface atoms are thought to form partially charged species of CO2δ. It is believed that oxygen coordination, carbon coordination and mixed coordination will be mainly formed in possible adsorption structures of CO2 (Fig. 10). In the example of oxygen coordination, each oxygen atom in CO2 can donate a lone pair of electrons to the center of the Lewis acid (Fig. 10a). In another carbon coordination example, the carbon atoms in CO2 can act as a Lewis acid and obtain electrons from the Lewis base to form a carbonic acid-like structure (Fig. 10b). In addition, in the case of mixed coordination, both oxygen and carbon atoms in CO2 act simultaneously as the electron donor and acceptor, respectively, as shown in Fig. 10c.128
image file: d4se00405a-f10.tif
Fig. 10 (a–c) Three adsorbed CO2δ structures and interactions.128

5.1 Photocatalytic CO2 reduction to methanol (CH3OH)

The formation of methanol from CO2 requires 6e and 6H+. Compared to CO formation, this chemical process is not easy. The typical process of CO2 reduction by a semiconductor photocatalyst proceeds in five sequential stages, i.e., light absorption, charge separation, CO2 adsorption, surface redox reaction, and product desorption. The first step is to generate electron and hole pairs from the absorption of photons. When light reaches the photocatalytic surface by incident light, electrons are excited from the valence band (VB) to the conduction band (CB), leaving the same number of holes in the VB. In this case, for the electrons or holes generated by light to be energetically advantageous in CO2 reduction or water oxidation, the photocatalyst must have an appropriate band gap size. Its CB edges should have a more negative value than the redox potential of CO2 reduction and its VB edge should have a more positive value than the redox potential of water oxidation (pH 7.0 aqueous solution, 0.817 V vs. SHE). In addition, an important performance indicator frequently cited in the literature for photocatalysts is the apparent quantum efficiency (AQE) or external quantum efficiency (EQE). AQE or EQE is defined as the ratio of the number of electrons delivered to a particular product for incident photons at a given wavelength,129 which can be expressed as a product of the efficiencies of light absorption, charge separation, and surface redox reactions. The surface redox reaction is based on the type of reaction product and reaction mechanism during the entire photocatalytic reaction process. There is a wide variety of CO2 reduction products, including methanol (CH3OH),130 carbon monoxide (CO),131 and methane (CH4),132 as presented in Table 3. Among the products, CH3OH has emerged as a new fuel of the 21st century due to its high energy density and high stability and is widely used as a raw material for the synthesis of chemical products.133 The reaction process of CO2–CH3OH is complex and proceeds through several reaction pathways proposed thus far, including formaldehyde production, carbene production, and glyoxal production route.134–137
Table 3 Synthesis method and CO2 reduction methods with different conditions using MXene-based catalysts
No. Catalyst Synthesis method Reduction method Solvent Product Ref.
1 La2TI2O7/Ti3C2 Solvothermal Photocatalytic H2O CO (14.78 μmol g−1), CH4 (11.16 μmol g−1) 100
2 MXene/GO/PDI Impregnation Photocatalytic CH3OH C2H4O2 (771.1 μmol g−1 h−1) 101
3 TiO2/Ti3CN Facile hydrothermal Photo electrocatalytic 0.1 M KHCO3 45.6 μM cm−2 h−1 102
4 Bi2O2SiO3/MXene In situ growth Photocatalytic CO (17.82 μmol g−1 h−1) and CH3OH (2.07 μmol g−1 h−1) 103
5 CdS/Ti3C2 MXene Hydrothermal Electrochemical 0.1 M KHCO3 CO (FE – 94%) 104
6 V2C/p-gC3N4 Physical mixing with ultrasonication Photocatalytic H2O CO (172.9 μmol g−1) and CH4 (266.5 μmol g−1) 105
7 Meso-gC3N4/MXene (Ti3C2Tx) Calcination Photocatalytic H2O CO (3.98 μmol g−1 h−1) and CH4 (2.117 μmol g−1 h−1) 106
8 2D/3D PCN/Ti3C2TA-TiO2 Ultrasonication Photocatalytic H2O/CH3OH (5%) CO (317 μmol g−1 h−1) and CH4 (79 μmol g−1 h−1) 107
9 CeO2/Ti3C2-MXene Hydrothermal Photocatalytic H2O/NaHCO3 CO (40.2 μmol m−2 h−1) 108
10 2D/2D/OD TiO2/C3N4/Ti3C2 Ultrasonication-assisted hydrothermal Photocatalytic NaHCO3/H2SO4 CO (4.39 μmol g−1 h−1) and CH4 (1.20 μmol g−1 h−1) 109
11 Cd0.2Zn0.8S@Ti3C2 (MXenes) Solvothermal Photocatalytic H2O CO (3.31 μmol h−1 g−1) and CH4 ((3.51 μmol h−1 g−1) 110
12 Ru–Ti3CN MXene-TiO2 Hydrothermal Photocatalytic H2O CO (99.58 μmol g−1) and CH4 (8.97 μmol g−1) 111
13 Pd50–Ru50–Ti3C2Tx MXene Microwave irradiation C2H6O2/H2O with 1 M NaBH4 CH3OH (76%) 112
CH4 (0.5%)
CO (0.5%)
14 2D/2D/2D O-gC3N4/Bt/Ti3C2Tx Ultrasonic Photocatalytic H2O/acetic acid CO (365 μmol g-cat−1 h−1) and CH4 (955 μmol g-cat−1 h−1) 113
15 Cu2O/Ti3C2 MXene Hydrothermal Photocatalytic H2O CO (17.55 μmol g−1 h−1) and CH4 (0.96 μmol h−1 g−1) 114
16 Cu0.05Zn2.95In2S6@Ti3C2OH Hydrothermal Photocatalytic H2O/C2H3N with TEOA CO (51.848 μmol g−1 h−1) and CH4 (0.105 μmol g−1 h−1) 115
17 BiVO4/Ti3C2Tx Self-assembly Photocatalytic 0.2 M NaHCO3 CH3OH (20.13 μmol g−1 h−1) 116
18 2D/1D Ti3C2Tx/p-BN Sonication/Stirring Photocatalytic H2O CO (22.53 μmol g−1 h−1) and CH4 (0.85 μmol g−1 h−1) 117


Table 4 Summary of CO2 reduction electrocatalysts in the recent literature
Electrocatalyst Electrolyte Selectivity and activity Stability Reference
Cu NCs with 44 nm edge length 0.1 M KHCO3 J tot = ≈5.7 mA cm−2, F. E. CO2RR 80%, ethylene 41%, methane 20% @ −1.1 V vs. RHE 118
Cu mesopore electrode (width/depth) 0.1 M KHCO3 J tot = 14.3 mA cm−2, F. E. C2H4 38% (30 nm/40 cm) C2H6 46% (30 nm/70 nm)@−1.7 V vs. NHE; onset potential −0.96 V vs. NHE 119
3D porous hollow fiber Cu electrode 0.3 M KHCO3 J tot = ≈10 mA cm−2, F.E. CO 75% @ −0.4 V vs. RHE 24 h @ −0.4 V vs. RHE 120
Cu NPs 13.1 nm 0.1 M KHCO3 J tot = 20 mA cm−2, H2 0.078, CO 0.016, CH4 0.0018, C2H4 0.0006 (vol. % cm−2) @ −1.1 V vs. RHE 121
Cu NPs 0.1 M NaHCO3 J tot = ≈9 mA cm−2, F. E. CH4 80%, H2 13% @ −1.25 V vs. RHE 1 h @ −1.25 V vs. RHE 122
OD Cu films 0.5 M NaHCO3 J tot = 2.7 mA cm−2, F. E. CO ≈ 40%, HCO2H 33% @ −0.5 V vs. RHE 7 h @ −0.5 V vs. RHE 123
Plasma-activated Cu 0.1 M KHCO3 F. E. C2H4 60% @ −0.9 V vs. RHE; onset E: −0.5 V vs. RHE 124
OD Au NPs 0.5 M NaHCO3 J tot = 6 mA cm−2, F. E. CO 98% @ −0.4 V vs. RHE 8 h @ −0.4 V vs. RHE 125
Au25 cluster 0.1 M KHCO3 J tot = ≈14.3 mA cm−2, F. E. CO 99.6% @ −0.89 V vs. RHE 126


A single atom-based photocatalyst derived from mC3N4, RuSA-mC3N4, was synthesized, as shown in Fig. 11i. Simply, ruthenium salt (RuCl3·7H2O), dicyanamide and P-123 (PEG-PPGPEG) and tetraethyl orthosilicate were used as the photocatalyst precursor, monomers, and template, respectively.138 In the initial stage (Fig. 11i, step 1), dicyanamide was well mixed with SBA-15 and calcined for complete absorption, and further heated and treated with HF to remove the silica mold to obtain the complete mesoporous carbon nitride (mC3N4). The mesoporous carbon nitride synthesized in the above-mentioned process was used as a photoactive support for the dispersion of ruthenium single atoms (Fig. 11i, step 4). Subsequently, an aqueous ruthenium chloride solution was added dropwise to the mesoporous carbon nitride under ultrasonic treatment (30 min), and microwave (MW) heating (P/No MEZ66853207, LG Electronics) was performed 10 to 20 times in a Millipore aqueous solution (Fig. 11i). According to the above-mentioned procedure, a single-atom RuSA-mC3N4 photocatalyst was synthesized, and the presence of porosity and single atom formation were confirmed through HAADF-STEM and X-ray absorption spectroscopy.


image file: d4se00405a-f11.tif
Fig. 11 (i) Schematic depiction of the synthesis of mesoporous C3N4. RuSA–mC3N4: ruthenium single atom; mC3N4: mesoporous carbon nitride; SBA-15: template, HF. (ii) (a) Wide-angle XRD patterns. (b) Ru 3d HR XPS spectra of RuSA–mC3N4. (c) Ru K edge FT-magnitude (solid lines) and imaginary part (dashed lines) EXAFS of RuSA–mC3N4 (black) and RuO2 standard (blue), solid line: magnitude FT, dotted line: imaginary part FT. (d) Ru K edge XANES of RuSA–mC3N4 (black), RuO2 standard (blue), RuCl3 standard (pink) and Ru foil (red). (e) Raman spectra of RuSA–mC3N4. (f). Schematic representation of RuSA–mC3N4 photocatalyst (N, C, or O based on the Raman spectra).138,139

To confirm the structural properties of the photocatalyst, the WAXRD pattern of g-C3N4 from the wide-angle XRD measurement shows a strong high peak reflection at 27.4° due to the interlayer distance of the carbon nitride sheet similar to the (002) reflection of graphite (Fig. 11iia). In the case of the chemical and elemental composition of the catalysts, in the XPS spectrum, the peaks at 530, 398, 288, and 287 eV correspond to O, C, Ru, and N, respectively. Together with ruthenium, the N1s signal bands of all the RuSA-mC3N4 samples were deconvolved with three nitrogen configurations, each containing N. The main N1s peak at 398.9 eV is attributed to sp2-hybridized pyridinic-N bound to carbon atoms (C–N–C). The peak at 400.2 eV is related to the tertiary N atoms of graphitic-N bonded to the C atom in the form of N–(C)3. The peak at 401.4 eV is attributed to nitrogen linked to a hydrogen atom known as pyrrolic-N, each referred to as C–N–H. The structural changes in the C–N scaffold after Ru doping were also confirmed by the Raman analysis (Fig. 11iie). The bands located in the range of 400–600 and 600–700 cm−1 correspond to the Ru–N/C and Ru–O vibrations, respectively, further confirming the presence of ruthenium bound to two different sites. The other bands in the Raman spectrum are the characteristic C–N, D and G bands of graphitic nitride. However, the broad nature of the D and G bands suggests the defective structure of the carbon matrix (Fig. 11iie).

5.2 Electrochemical CO2 reduction to methanol (CH3OH) on electrocatalyst

To date, various electrocatalytic materials have been employed in the CO2 reduction process. Among them, metal alloys, inorganic and organic–metal compounds, and pyridine-based electrocatalysts are attracting significant attention due to their excellent selectivity. In recent years, metal-free compounds have been developed rapidly because not only MOF electrocatalysts have high selectivity, but also metal utilization can be avoided and the catalytic activity can be improved. The recent development of CO2 to CH3OH reduction by the ECR method and other types of electrocatalysts are discussed in the following section. In addition, the crystal structure, form, and experimental conditions of various electrocatalytic chemical reactions using NHE, RHE, and SHE are summarized (Fig. 12).
image file: d4se00405a-f12.tif
Fig. 12 (i) SEM image of Cu2O(OL-MH)/PPy particles (A). (B) Magnified view of icosahedra and octahedra structures of Cu2O. (C) SEM image of Cu2O(OL-MH)/PPy particles separated from LT paper. Magnified view of two icosahedra is displayed in the inset of (i) (C). (D) TEM image of Cu2O(OL-MH)/PPy particles. (ii) (a) XRD pattern of BP. (b) TEM and (c) HRTEM images of BP. (d) STEM image and EDX elemental mapping images for BP.140,141

5.3 MXene-based photocatalytic reactions

Fig. 13 shows a schematic illustration of the possible chemical reaction on the catalyst surface and the charge transition during the CO2 reduction process. Also, a possible schematic representation of the photocatalysis mechanism over the surface of MXene-based nanomaterials under visible light irradiation is shown in Fig. 13. Considering the rapid growth of the application of MXene-based nanomaterials, several review studies have been reported on their synthesis for various applications.142–144 In particular, MXene-based photocatalysts are synthesized by replacing the precious metal co-catalyst to improve the charge-separation capacity of the photocatalysts. Methods such as mechanical mixing, self-assembly, and in situ decoration/oxidation are the most widely used techniques for manufacturing photocatalytic complexes.145 Among them, mechanical mixing is the simplest technique for manufacturing photocatalytic complexes, involving the mixing of different components in a solution or pulverized powder. Interestingly, negatively charged MXenes are easily combined with positively charged photocatalysts due to electrostatic attraction, resulting in the formation of self-assembled photocatalytic complexes.145 Unlike the mechanical mixing and self-assembly methods, in situ decoration techniques involve synthesizing different components directly onto the MXene surface. As a result, in situ composites and MXenes with strong chemical bonds have significant advantages in some designs.145
image file: d4se00405a-f13.tif
Fig. 13 Schematic depiction of charge separation. (a) Proposed photocatalytic CO2 reduction mechanism of Ti3C2eOH/TiO2.146 (b) Proposed mechanism for photocatalytic H2 production in the CdS/Ti3C2 system.147 (c) Charge-transfer process over (001)TiO2/Ti3C2 and (d) schematic band alignments at {001}TiO2eTi3C2 interfaces.148

Handoko and coworkers149 reported the results of the experimental and theoretical bonding of Ti- and Mo-based MXene catalysts as electrocatalysts for CO2RR. In these reactions, formic acid plays a role in producing the major products of Ti2CTx and Mo2CTxMXene, with a peak faradaic efficiency of 56% or more in Ti2CTx and a partial current density of up to 2.5 mA cm−2 in Mo2CTx. Smaller overpotentials have been shown to occur in the presence of smaller amounts of –F functional groups. This reaction process represents an important step toward experimental proof of the application of MXenes in more complex electrocatalytic reactions in the future, and the results are shown in Fig. 14.


image file: d4se00405a-f14.tif
Fig. 14 Comparison of (a) faradaic efficiency and (b) partial current density for CO2RR to formic acid on Ti2CTx and Mo2CTx MXenes. (c) *HCOOH binding energy plot against *COOH binding energy. (d) Limiting CO2RR potentials. The lines represent the calculated potential where the most negative reaction steps are neutral as a function of *COOH binding energy (PCET-1: *CO2 + H+ + e/*COOH; PCET-2: *COOH + H+ + e/*HCOOH). Calculated free energy diagram at 0 V applied potential for CO2RR to formic acid on (e) Ti2CTx and (f) Mo2CTx MXenes with varying fractions of –F and –O surface-terminating groups.149

Due to the excellent properties of MXenes, an increasing number of MXene-based photocatalysts are being used to reduce CO2. As shown in Table 2, MXene-combined photocatalysts can be divided into five categories including MXene/metal oxides, MXene/nitrides, MXene/LDH, MXene/perovskite, and MXene-derived photocatalysts. Ti3C2 MXene was introduced by Ye et al. in 2018 by decorating the most popular commercial TiO2, P25.150 Initially, Ti3C2 MXene was alkalized with KOH to replace –F with –OH. The CO2 absorption rate of alkalized Ti3C2 MXene (TC-OH) was 7.01 cm3 g−1, which was 38.9 times higher than that of Ti3C2 MXene. DFT calculations also confirmed that the low adsorption energy of CO2 for Ti3C2 MXene –OH (Ead = −0.44 eV) than Ti3C2 MXene (Ead = −0.13 eV). Therefore, the optimal sample (5 Ti3C2 MXene-OH/P25) showed a yield of 11.74 μmol g−1 h−1, which was three times higher for CO2 than that of bare P25, and 16.61 μmol g−1 h−1, which was 277 times higher for CH4. The improved photocatalytic activity was attributed to the excellent electrical conductivity and abundant adsorption sites on Ti3C2 MXene-OH for the adsorption and activation of CO2. A similar strategy was applied to other metal oxide semiconductors, where the alkalized Ti3C2 MXene was used as a co-catalyst to decorate ZnO by Wang et al.152 After modifying 7.5 wt% of alkalized Ti3C2 MXene, Ti3C2 MXene-OH/ZnO showed an improved evolution of CO (30.30 μmol g−1 h−1) and CH4 (20.33 μmol g−1 h−1), which was approximately 7 and 35 times that by pristine ZnO, respectively. The excellent photocatalytic activity was attributed to the improved separation/transfer of photoinduced charge carriers and improved adsorption/activity of CO2 molecules. As another example, MXene was used to enhance the photocatalytic activity in combination with CeO2. Liu et al. synthesized a CeO2/Ti3C2-MXene hybrid with a Schottky junction using a simple hydrothermal synthesis method.153 At the optimal ratio, the CO production rate of CeO2/MXene-5% was 40.2 μmol m−2 h−1, which was 1.5 times higher than that of pure CeO2. The photoelectrons generated by the Schottky junction induced by the built-in electric field moved from the CB of CeO2 to Ti3C2 MXene, facilitating the separation of electrons and holes. In another study, CeO2@Ti3C2TX nanosheets were deposited on Ti3C2TX nanosheets through a hydrothermal synthesis pathway to synthesize CeO2@Ti3C2TX.154 It was confirmed that the light absorption of the composite was extended to the IR region by taking advantage of the narrow bandgap of Ti3C2TX. The yield of alcohol by the 12.5 CT composite under vis-IR irradiation for 4 h was 102.24 and 59.21 μmol gcata.−1 Interestingly, the yield of methanol and ethanol was 81 and 38.12 μmol gcata.−1 only under NIR irradiation, which was higher than that under VL irradiation, respectively (methanol: 76.2 μmol gcata.−1 and ethanol: 37.8 μmol gcata.−1). Lu et al. produced a Cu2O/Ti3C2Tx heterojunction composite via the in situ hydrothermal growth method.114 This improvement in photocatalytic activity is due to the excellent photoelectronic conductivity of MXene. If the size of Ti3C2 MXene is reduced to the size of quantum dots (QD), another effect may be caused. A new co-catalyst effect was reported by Chen et al. through the complexation of Ti3C2 QDs with Cu2O nanowires (NWs).155 According to the photocatalytic effect of Ti3C2 QDs/Cu2O NWs/Cu, methanol production by CO2 reduction was 2.15 times that of Ti3C2 nanosheets/Cu2O NWs/Cu. The experimental results and DFT calculations demonstrated that the introduction of Ti3C2 QDs can improve the stability of Cu2ONNs as well as improve the charge transfer and carrier density. One of the important benchmarks in MXene-based photocatalysts is the 2D/2D heterojunction of ultrathin MXene/Bi2W6 nanosheets.156 In another study, Bi2W6 was in situ grown on the surface of ultrathin Ti3C2 nanosheets. In addition, as the surface area increased from 17 m2 g−1 to 36 m2 g−1, the total yield of CH4 and CH3OH obtained from TB 5, the optimized sample, was 4.6 times higher than that of pristine Bi2W6. In addition, there are other MXene/metal oxides used for photocatalytic CO2 reduction, such as InVO4/Ti3C2Tx,157 Ti3C2/La2Ti2O7,158 Bi2O2SiO3/Ti3C2,103 and meso-TiO2@ZnIn2S4/Ti3C2 (Table 5).160

Table 5 MXenes/metal oxide photocatalyst systems for CO2 photoreduction
Catalyst Light source Reaction system Reaction condition Performance (μmol g−1 h−1) AQY (%) Reference
MXene/Metal oxides
Ti3C2–OH/P25 300 W Xe lamp Gas–solid CO2 + 3 mL H2O CO: 11.74 CO: 0.32 151
CH4: 16.61 CH4: 1.61 (λ = 380 nm)
Ti3C2–OH/ZnO 300 W Xe lamp Gas–solid 70 kPa CO2 + 3 mL H2O CO: 30.30 NA 152
CH4: 20.33
CeO2/Ti3C2 350 W Xe lamp Gas–solid 0.084 g NaHCO3 + 0.3 mL HCl (2 M) CO: 40.2 μmol m−2 h−1 NA 108
CeO2 @Ti3C2Tx 300 W Xe lamp Gas–liquid CO2 + 30 mL H2O CH3OH: 25.56 NA 154
C2H5OH: 14.80
Cu2O/Ti3C2Tx 300 W Xe lamp Gas–solid CO2 + H2O CO: 17.55 NA 114
CH4: 0.96
Ti3C2 QDs/Cu2O NWs/Cu 300 W Xe lamp Gas–liquid CO2 + 80 mL H2O CH3OH: 25.56 ppm cm−2 h−1 NA 155
Bi2WO6/Ti3C2Tx 3060 W Xe lamp Gas–liquid NaHCO3 + 0.3 mL H2SO4 (2 M) CH4: 1.78 NA 156
CH3OH: 0.44
InVO4/Ti3C2Tx 300 W Xe lamp Gas–solid CO2 + 0.4 mL H2O CO: 13.83 NA 157
CH4: 0.71
La2Ti2O7/Ti3C2 NA Gas–solid CO2 + 1 mL H2O CO: 14.78 NA 102
CH4: 11.16
Bi2O2SiO3/Ti3C2 300 W Xe lamp Gas–liquid CO2 + 50 mL H2O CO: 17.82 NA 159
CH3OH: 2.07
meso-TiO2 @ZnIn2S4/Ti3C2 300 W Xe lamp NA NA CO: 10.17 NA 160
CH4: 11.3


Graphitic carbon nitride (g-C3N4), a type of non-metallic semiconductor, is used as an ideal photocatalyst for CO2 reduction reactions due to its easy production, suitable bandgap and visible light reactivity.161,162 However, its future applications are limited due to its severe recombination of photogenic electron–hole pairs and poor active sites. Therefore, MXene has been used in combination to further improve the photocatalytic performance of g-C3N4. Lv et al. directly calcined a mixture of bulk Ti3C2 and urea to form an ultra-thin 2D/2D Ti3C2/g-C3N4 heterojunction.163 In this system, urea not only acted as the gas template to transform the multilayer Ti3C2 into nanosheets, but also as the precursor of g-C3N4. The optimal sample, 10TC, showed CO and CH4 yields of 5.19 and 0.044 μmol g−1 h−1, and the CO2 conversion rate of 10TC was 8.1 times higher than that of bare g-C3N4. To demonstrate the improved photocatalytic activity, the authors presented the improved CO2 adsorption properties together with effective space separation of photocatalytic charge carriers by the ultrathin 2D/2D heterojunction. In another study reported by Wu et al.,164 the electrostatic effect of B-doped graphitic carbon nitride (B-gCN) was studied, and they suggested the multilayer effect of Ti3C2MXene (FLTC). At this time, the optimized sample (12FLTC/BCN) showed higher CO and CH4 production yields than the raw g-C3N4. It was suggested that the synergistic effect of the boron dopant and the additional effect of FLTC are the main reasons for the further improvement in the photocatalytic performance. A 2D/2D Schottky junction effect consisting of defective g-C3N4 nanosheets with carbon vacancies representing the defects and Ti3C2TX MXene was designed and synthesized by the Jiang group.165 The CO evolution on 20% Ti3C2Tx/Vc-CN showed a much higher efficiency than that of bare CN under visible light irradiation. The enhancement in photocatalytic degradation activity was shown to be promoted the electron transfer effect by exciton dissociation promoted by carbon vacancy and Schottky junction between Ti3C2Tx and Vc-CN. Other MXene/g-C3N4 hybrids such as alkalinized Ti3C2/g-C3N4,166 Ti3C2/porous g-C3N4,151 V2C/g-C3N4 (ref. 167) and Ti3C2/mesoporous g-C3N4 (ref. 106) were also reported (Table 6).

Table 6 MXene/nitride photocatalyst systems for CO2 photoreduction
Catalyst Light source Reaction system Reaction condition Performance (μmol g−1 h−1) AQY (%) Reference
Ti3C2/g-C3N4 300 W Xe lamp (λ > 420 nm) Gas–solid 1.26 g NaHCO3 + 4 mL H2SO4 (2 M) CO: 5.19 NA 163
CH4: 0.044
FLTC/BCN 300 W Xe lamp (λ > 420 nm) Gas–solid CO2 + water vapor CO: 2.88 0.0117% (λ = 420 nm) 164
CH4: 0.16
Ti3C2TX/Vc-CN 300 W Xe lamp (λ > 400 nm) Gas–solid 0.1 g NaHCO3 + 0.1 mL H2SO4 (2 M) CO: 20.54 NA 165
Alkalinized Ti3C2/g-C3N4 300 W Xe lamp (λ > 420 nm) Gas–solid CO2 + water vapor CO: 2.24 0.0099% (λ = 420 nm) 166
CH4: 0.041
Ti3C2/porous g-C3N4 300 W Xe lamp (λ > 420 nm) Gas–solid CO2 + 15 mL H2O CO: trace NA 151
CH4: 0.99
V2C/g-C3N4 35 W HID Xe lamp Gas–solid CO2 + water vapor CO: 37.75 NA 105
CH4: 51.25
Ti3C2/mesoporous g-C3N4 300 W high-pressure Xe lamp Gas–solid 1 mL CO2 + 5 mL H2O CO: 3.98 NA 106
CH4: 2.117
TiO2/C3N4/Ti3C2 350 W Xe lamp Gas–solid 0.084 g NaHCO3 + 0.3 mL H2SO4 (2 M) CO: 4.39 NA 109
CH4: 1.20
g-C3N4/TiO2/C 300 W Xe lamp Gas–solid CO2 + water CO: 8.65 NA 168
CH4: 1.23
Cu–Ti3C2Tx/g-C3N4 300 W Xe lamp (λ > 400 nm) Gas–solid CO2 + water vapor CO: 49.02 NA 169
CH4: 3.60
g-C3N4/Ti3C2/MoSe2 300 W Xe lamp SL3 deuterium Gas–solid CO2 + water CO: 29.87 NA 170
CH4: 17.94
Ti3C2/CN/Cu2O Light source (λ = 354 nm) SL3 deuterium Gas–liquid CO2 + water CH3OH rates: 22.41% NA 171
Ti3C2/CN/ZnO Light source (λ = 354 nm) Gas–liquid CO2 + water CH3OH rates: 20.12% NA 172


The MXene/g-C3N4 hybrid complex has been extended to a system that combines three components. Macyk et al. studied the 2D/2D/0D TiO2/C3N4/Ti3C2 complex heterojunction photocatalyst, where Ti3C2 MXene quantum dots (TCQD) were incorporated in TiO2/C3N4 by electrostatic interaction.109 The dual heterojunction (S-scheme heterojunction at TiO2/C3N4 interface and Schottky heterojunction at C3N4/TCQD interface) played an important role in improving the CO2 reduction activity by the photocatalyst. In these studies, the CO and CH4 yields of TiO2/C3N4/Ti3C2 were eight times higher than that of the original C3N4. Another three-component hybrid, g-C3N4/TiO2/C, was studied by the Xiang group. Initially, g-C3N4 was combined with a single-layer Ti3C2 MXene by electrostatic self-assembly.168 A method of forming TiO2/C by calcining a mixture was suggested, and the single-layer Ti3C2 may prevent the aggregation of g-C3N4 and serve as a cross-linking agent between TiO2 and Ti3C2, which are relatively inert. Therefore, g-C3N4/TiO2/C showed 2.88 times and 5.82 times higher CO and CH4 production than pure g-C3N4, respectively. A Cu–Ti3C2Tx/g-C3N4 photocatalyst was constructed by Su et al.169 In another way, Cu was self-reduced on the surface of Ti3C2Tx and the Cu–Ti3C2Tx cocatalyst was combined with g-C3N4 to synthesize the Cu–Ti3C2Tx/g-C3N4 composite photocatalyst by electrostatic self-assembly. Other MXene/nitride catalysts with three components such as g-C3N4/Ti3C2/MoSe2,170 Ti3C2/CN/ZnO,171 and Ti3C2/Cu2O171 have also been reported.

Another type of complexation is the combination of MXenes and layered double hydroxides (LDH) to promote their photocatalytic activity. Ti3C2 MXene was loaded in the NiAl-LDH system to form a three-dimensional hierarchical NiAl-LDH/Ti3C2 MXene (LDH/TC) nanocomposite.172 This nanocomposite possessed a high specific surface area and designed to enhance the adsorption and utilization of photons due to its three-dimensional hierarchical structure. In addition, due to the structural hydrophilicity of Ti3C2, an intimate contact with NiAl-LDH was formed, and this structural rationality contributed to maximizing the separation of electron–hole pairs by light generation. As a result, the LDH/TC-2 sample showed an enhanced efficiency by 3.2 times and 7.3 times that of the original LDH for CO and CH4 generation, respectively. In addition, the activity of the LDH/TC-2 sample did not significantly decrease after 4 cycles of testing in the cycle measurement of CO yield. A similar NiAl-LDH/Ti3C2 composite was synthesized by Li et al.173 In another study, Zhou et al. synthesized a new type of photocatalyst possessing a new ZnCr-LDH/Ti3C2Tx Schottky junction structure by electric field coprecipitation.174 The introduction of Ti3C2Tx MXene increased the light absorption and efficiently induced light-induced electron separation, greatly improving the electron transfer efficiency. High conductive MXene species and nanoarray architectures were the main reasons for the improved photocatalytic activity (Table 7).

Table 7 MXene/LDH photocatalyst systems for CO2 photoreduction
Catalyst Light source Reaction system Reaction conditions Performance (μmol g−1 h−1) AQY (%) Reference
NiAl-LDH/Ti3C2 300 W Xe lamp Gas–solid CO2 + 0.4 mL H2O CO: 11.82 NA 172
CH4: 1.02
NiAl-LDH/Ti3C2 300 W Xe lamp (λ > 420 nm) Gas–liquid CO2 + 12 mL mixed solution (MeCN/H2O/TEOA) CO: 2128.46 NA 173
ZnCr-LDH/Ti3C2Tx 300 W Xe lamp Gas–solid CO2 + 0.4 mL H2O CO: 20.41 NA 174
CH4: 3.32
Co–Co LDH/Ti3C2 5 W LED Gas–liquid MeCN/H2O/TEOA (3 mL/2 mL/1 mL) CO: 1.25 × 104 CO: 0.92% (λ = 420 nm) 175


H. Shen et al.177 used monolayer ZnTi-LDHs with good CO2 adsorption property and modified Ti3C2 nanosheets with different coverage (θ) at their O-terminus (θ = 1, 8/9, 7/9, 1/3), as one MXene, to build a ZnTi-LDH/Ti3C2O2 heterostructure.176 The d-band centers of ZnTi-LDH/Ti3C2O2 at different O vacancy amounts were systematically investigated using DFT, focusing on the energy band arrangement, charge transfer and reaction pathway during the photocatalytic reduction of CO2. This Schottky heterostructure is more favorable for absorbing solar energy and rapidly exciting electrons. Further analysis of ZnTi-LDH/Ti3C2O2 (θ = 8/9) formed by Ti3C2O2 with 8/9 O-terminal coverage showed that the built-in electric field at the interface of ZnTi-LDH and Ti3C2O2 (θ = 8/9) induces band bending and charge separation, which favor the reduction of CO2 to CO through multi-electron consumption. Finally, the reaction mechanism for the photocatalytic reduction of CO2 to CO on ZnTi-LDH/Ti3C2O2 (θ = 8/9) was proposed (Fig. 15).


image file: d4se00405a-f15.tif
Fig. 15 Molecular electrostatic potentials for (a) CO2, (b) ZnTi-LDH/Ti3C2O2 (θ = 8/9) and (c) CO2 interacting with ZnTi-LDH/Ti3C2O2 (θ = 8/9). (d) d-band centers of ZnTi-LDH, Ti3C2O2 and ZnTi-LDH/Ti3C2O2. Band structures of (e) ZnTi-LDH/Ti3C2O2 and (g) ZnTi-LDH/Ti3C2O2 (θ = 8/9). TDOS and PDOS of (f) ZnTi-LDH/Ti3C2O2 and (h) ZnTi-LDH/Ti3C2O2 (θ = 8/9). Electrostatic potentials along the Z axis of (i) ZnTi-LDH and (j) Ti3C2O2 (θ = 8/9). (k) Planar-averaged electron density difference with Z direction for the ZnTi-LDH/Ti3C2O2 (θ = 8/9).177

Composites with a perovskite structure are a type of promising photocatalyst for reducing CO2.178 However, pristine perovskites suffer from low gas adsorption capacity and inefficient active sites, resulting in unsatisfactory photocatalytic activity. Recently, Chen et al. synthesized a Schottky heterojunction and FAPbBr3/Ti3C2 composite by anchoring FAPbBr3 quantum dots (QDs) to Ti3C2 nanosheets.179 The exciton separation effect of the Ti3C2 2D sheet was determined by the supplied catalyst site as well as the electron acceptor. According to the results, the FAPBBr3/0.2-Ti3C2 composite showed a 2.08 times higher electron consumption rate than that of bare FAPBr3. Then, the same group developed a 2D/2D FAPBBr3/Ti3C2 Schottky heterojunction using the hot-injection and in situ growth method.180 The size of FAPbBr3 was reduced to 2.5–4.5 nm and the CO yield of FPB/TC-2 using ethyl acetate/deionized water as a sacrificial reagent was 93.82 μmol g−1 h−1, which was 25% higher than that of pristine FAPbBr3. Liu et al. developed CsPbBr3/MXene nanosheet composites via the in situ growth of CsPbBr3 perovskite nanocrystals on 2D MXene nanosheets.181 The CsPbBr3/MXene-20 nanocomposite showed better CO and CH4 formation rates than the bare CsPbBr3 (<4.4 μg g−1 h−1 mol−1) of 26.32 and 7.25 μmol g−1 h−1, respectively.183 The Ti3C2 MXene nanosheets facilitated the formation of free charge carriers in Cs2AgBiBr6 and extended the charge carrier life. Therefore, a high electron consumption yield of 50.6 μmol g−1 h−1 was achieved by Cs2AgBiBr6/Ti3C2 (Table 8).

Table 8 MXene/perovskite photocatalyst systems for CO2 photoreduction
Catalyst Light source Reaction system Reaction condition Performance (μmol g−1 h−1) AQY (%) Reference
FAPbBr3 QDs/Ti3C2 300 W Xe lamp Gas–solid CO2 + 0.5 mL H2O CO: 283.41 NA 179
CH4: 17.67
FAPbBr3/Ti3C2 300 W Xe lamp Gas–solid CO2 + 0.5 mL H2O CO: 93.82 NA 180
CsPbBr3/Ti3C2 MXene 300 W Xe lamp (λ > 420 nm) Gas–liquid CO2 + ethyl acetate CO: 26.32 NA 181
CH4: 7.25
CsPbBr3/Ti3C2Tx 300 W Xe lamp (λ > 400 nm) Gas–solid CO2 + water vapor Total electron consumption: 112.6 NA 182
Cs2AgBiBr6/Ti3C2Tx Xe lamp (λ > 400 nm) Gas–solid CO2 + 20 μL H2O Total electron consumption: 50.6 NA 183


In 2018, a pioneering study on MXene-derived photocatalysts was conducted by the Yu group.46 TiO2 nanoparticles were grown on high-conductive Ti3C2 MXene by calcination. The obtained TiO2/Ti3C2 MXene complex showed a similar structure to a rice shell. The optimized sample, TT550, showed CH4 evolution of 0.22 μmol h−1, which was 3.7 times higher than that of P25. The close contact between TiO2 and Ti3C2 MXene accelerated the transfer of photogenerated electrons from TiO2 to Ti3C2. In addition, the large specific surface area effect is associated with many active sites, and the multi-layered structure, such as the unique rice shell, induces a large specific surface area and improves the CO2 reduction reaction by the photocatalyst. Cao et al. successfully synthesized a 3D hierarchical Ti3C2Tx/TiO2 heterojunction structure by calcining 3D hierarchical Ti3C2Tx in air.184 The TT500 sample showed a relatively high photocatalytic CH4 evolution activity of 4.41 μmol g−1 h−1. The strong interfacial interaction between Ti3C2Tx and TiO2 improved the photocatalytic activity by improving the electron–hole separation and CO2 adsorption. Lv et al. described the easy production of a (001) TiO2/Ti3C2Tx photocatalyst,185 where HF acted as a morphological regulator for the growth of the (001) TiO2 nanosheets as well as an etchant for the removal of Al. The resulting (001) TiO2/Ti3C2Tx heterojunction showed a CO yield of 13.45 μmol g−1 h−1. The spatial isolation of the photogenerated charge carrier induced by Ti3C2Tx was the major cause for the superior performance of the (001) TiO2/Ti3C2Tx heterojunction. The Ru–Ti3CN MXene-TiO2 photocatalyst was constructed by Wang et al., where Ti3CN MXene was synthesized by Lewis acid etching in the absence of toxic HF.111 Ti3CN MXene was later applied as a support for the field growth of TiO2 and Ru nanoparticles. After a 5 h experiment, Ru–Ti3CN–TiO2 showed the CO and CH4 production rates of 99.58 and 8.97 μmol g−1 h−1, which were 20.5 times and 9.3 times that of P25, respectively. The layered Ti3CN MXene could provide a rich path for electron transfer, and the addition of Ru further increased the separation and transfer of charges. 3D hierarchical TiO2/Ti3C2Tx was modified by a photo-reduction method by introducing a single Pt atom.186 The single-atom Pt could effectively capture photo-generated electrons through its atomic interface, Pt–O binding, and served as an active site for CO2 adsorption and activation simultaneously (Table 9).

Table 9 MXene-derived photocatalysts systems for CO2 photoreduction
Catalyst Light source Reaction system Reaction condition Performance (μmol g−1 h−1) AQY (%) Reference
TiO2/Ti3C2 300 W Xe lamp Gas–solid 84 mg NaHCO3 + 0.4 mL 4 M HCl CH4: 0.22 NA 46
Ti3C2Tx/TiO2 300 W Xe lamp Gas–solid 84 mg NaHCO3 + 0.3 mL 2 M H2SO4 CH4: 4.41 NA 184
(001)TiO2/Ti3C2Tx 300 W Xe lamp Gas–solid 2.1 g NaHCO3 + 7 mL 2 M H2SO4 CO: 3.17 NA 185
CH4: 10.28
Ru–Ti3CN–TiO2 300 W Xe lamp Gas–solid CO2 + water vapor CO: 19.92 NA 111
CH4: 1.79
Pt SA-TiO2/Ti3C2 300 W Xe lamp Gas–solid CO2 + water vapor CO: 20.5 NA 186
CH4: 1.79
Pt–TiO2 NW/Ti3C2 NA Gas–solid CO2 + water vapor CO: 3.81 NA 187
CH4: 3.62
Co3O4–TiO2/C 300 W Xe lamp (λ > 420 nm) Gas–liquid CO2 + 3 mL C2H3N + 2 mL H2O + 1 mL C6H15NO3 CO: 23[thin space (1/6-em)]040 NA 188
H2: 10[thin space (1/6-em)]170
CeO2/Ti3C2/TiO2 300 W Xe lamp Gas–liquid CO2 + 5 mL H2O CO: 0.986 NA 189
CH4: 4.342
TiO2/Ti3C2 500 W Hg lamp Gas–liquid CO2 + H2O (with 10% TEOA) CO: 604.15 NA 190
CH4: 79.55


5.4 Sacrificial study for photocatalytic effects

The MXene-based photocatalyst materials studied for CO2 reduction with the sacrificial agent effect to hydrocarbon fuels and other products are summarized in Table 10. As shown in Table 10, MXene-based photocatalysts have high potential to improve the photocatalytic activity, charge carrier, and overall characteristics of the neighboring catalyst as a co-catalyst.
Table 10 Summary of the CO2 reduction study over MXene-based photocatalysts with sacrificial agents
No. Catalyst Synthesis method Reduction method Sacrificial agent Product Ref.
1 MXene/GO/PDI Impregnation Photocatalytic CH3OH C2H4O2 (771.1 μmol g−1 h−1) 101
2 TiO2/Ti3CN Facile hydrothermal Photo electrocatalytic 0.1 M KHCO3 45.6 μM cm−2 h−1 102
3 CdS/Ti3C2 MXene Hydrothermal Electrochemical 0.1 M KHCO3 CO (FE – 94%) 104
4 2D/3D PCN/Ti3C2TA-TiO2 Ultrasonication Photocatalytic H2O/CH3OH (5%) CO (317 μmol g−1 h−1) and CH4 (79 μmol g−1 h−1) 107
5 CeO2/Ti3C2-MXene Hydrothermal Photocatalytic H2O/NaHCO3 CO (40.2 μmol m−2 h−1) 108
6 2D/2D/OD TiO2/C3N4/Ti3C2 Ultrasonication-assisted hydrothermal Photocatalytic NaHCO3/H2SO4 CO (4.39 μmol g−1 h−1) and CH4 (1.20 μmol g−1 h−1) 109
7 Pd50–Ru50–Ti3C2Tx MXene Microwave irradiation C2H6O2/H2O with 1 M NaBH4 CH3OH (76%) 112
CH4 (0.5%)
CO (0.5%)
8 2D/2D/2D O-gC3N4/Bt/Ti3C2Tx Ultrasonic Photocatalytic H2O/acetic acid CO (365 μmol g-cat−1 h−1) and CH4 (955 μmol g-cat−1 h−1) 113
9 Cu2O/Ti3C2 MXene Hydrothermal Photocatalytic H2O CO (17.55 μmol g−1 h−1) and CH4 (0.96 μmol h−1 g−1) 114
10 Cu0.05Zn2.95In2S6@Ti3C2OH Hydrothermal Photocatalytic H2O/C2H3N with TEOA CO (51.848 μmol g−1 h−1) and CH4 (0.105 μmolg−1 h−1) 115
11 BiVO4/Ti3C2Tx Self-assembly Photocatalytic 0.2 M NaHCO3 CH3OH (20.13 μmol g−1 h−1) 116


5.5 MXene-based photoelectrocatalytic (PEC) reactions

In a photoanode-based PEC cell, the OER is driven by an n-type semiconductor photoanode with an appropriate counter electrode (platinum, etc.), as shown in Fig. 16a. Moreover, in a photoanode-based PEC cell, the p-type semiconductor accelerates the HER with the corresponding OER to perform the entire water splitting, as shown in Fig. 16b. CO2 reduction by PEC systems is rapidly advancing due to the fascinating properties of photocatalytic systems, including low operating cost, adjustable overpotential through applied bias, flexibility of photoelectrode selection, and high stability. PEC systems employed for CO2 reduction consist of anode and cathode compartments divided by a proton exchange membrane (e.g., Nafion), as shown in Fig. 16c. This figure shows the evolution of the synthesis of 2D-MXene (left), the discovery of (center) electrode fabrication methods, and (right) various MXene-based battery and supercapacitor configurations. The CO2 reduction reaction by PEC systems proceeds via a mechanism similar to water division, including photo-induced charge generation and movement to each electrode, especially the OER process. However, the movement of a large number of electrons to the cathode surface is required to reduce the thermodynamically stable CO2 to fuels such as CH3OH, CH3CH2OH, CH4, HCOOH, and HCHO in the potential window of 1.90 to 0.33 V vs. NHE. Hori et al. found in their pioneering study that 1.90 V vs. NHE CO2 was initially reduced to CO2 ions with massive kinetic barriers.192 Subsequently, these CO2 anions react in various ways depending on the properties of the catalyst. The detailed reaction route for CO2RR is as follows:
 
CO2 + e → CO2ads(9)
 
H + e → Hads(10)
 
CO2ads + Hads → HCOO(11)
 
CO2ads + H2O + e → HCOO + OH(12)
 
HCOO + e → CO + OH(13)
 
CO2ads + CO2ads → 2CO(14)
 
CO2ads + CO2 + e → CO + CO32−(15)

image file: d4se00405a-f16.tif
Fig. 16 Some types of PEC water-splitting/CO2 reduction with cell design: (a) photoanode-based cell, (b) photocathode-based cell, and (c) PEC cell.191

In addition to the above-mentioned reactions, the H2 generation reaction is also distinguished from the CO2RR process, which can be the rate-determining step. Obviously, CO2RR can proceed as a thermodynamic reaction, but some auxiliary energy must be provided to the electrodes to alleviate the activation energy barrier. Unlike water splitting, band-bending at the interface is not the only decisive parameter, where stabilization is much more important for the transition process, which includes CO2 adsorption and activation and complex multi-electron transfer.193 Insights into the selectivity and reaction pathways to produce products were well discussed in a professional review article on the reduction process of CO2.194–196 Although this section deals with the basic principles related to the reduction of CO2 by PEC systems, they are discussed in a recent literature on BiVO4-based photoelectrodes for CO2 reduction systems, which play an important role in promoting redox reactions in water splitting as well as CO2 capture systems. Therefore, a basic but noteworthy requirement for semiconductors to achieve a photocatalytic performance for CO2 reduction by competitive PEC is that photocatalytic semiconductors should be able to absorb a large portion of the solar spectrum and their bandgap energy should be greater than the water dissociation energy (1.23 eV) for efficient water division reactions. In addition, in this system, there must be minimal loss of carriers at the interface and on the semiconductor–electrolyte surface. Semiconductors should be able to provide a minimum overpotential for the H2 or O2 evolution reaction. Also, they should have long-term stability and resistance to corrosion under illumination, be easily synthesized for scalable device fabrication, and be cost effective and abundant.

Fig. 17 shows the reaction used in PEC for CO2R reduction by WE and CO2 reduction by photoelectrons generated on WE. By supplying electricity, the surface of TCCuFe is activated, electrons are excited, and the electrons accumulated on the surface are used to reduce CO2 to methanol. WE was manufactured using photocatalysts, and Ni foam, which can improve electron collection in WE and greatly improve the CO2R performance by PEC, was used as the current collector. Alternatively, excited electrons were efficiently transferred to the interface between WE and the electrolyte, driving the reduction reaction of CO2 on the WE surface. The electrolyte effect on CO2R by PEC is expressed in Fig. 17a, while the chemically absorbed halide ions contributed or retained negative changes in TCCuFe, changing the local environment on the surface of TCCuFe. As a result, the partial positively charged C-atom of CO2 has the ability to interact strongly with the TCCuFe surface, and in addition to the above-mentioned process, the electrons generated by the light supply were consumed in CO2R to obtain alcohol in high yield. The mechanism is shown in Fig. 17b to better understand the improvement of CO2R activity by PEC of CO2 to methanol and ethanol by the TCCuFe catalyst.


image file: d4se00405a-f17.tif
Fig. 17 (a) Representative experiment setup of PEC CO2 reduction. (b) Mechanism for PEC-CO2R over Ti3C2/Cu2O/Fe3O4 ternary nanocomposites under visible light irradiation.197

Ti3C2 QDS has been demonstrated to contribute significantly to the migration behavior of electro–hole pairs excited by light in the host photocatalysts. Compared to Ti3C2 sheet/Cu2O NW/Cu, and Ti3C2 QD/Cu2O NWs/Cu heterostructures, they showed excellent results in terms of improved light absorption, charge separation, and small band gaps, respectively. In addition, the current density is increased by promoting the movement of light-generated electrons from Cu2OWs/Cu to Ti3C2 QDs in these complex photocatalysts. The high current density indicates that Ti3C2 QDs accelerate the charge separation. Therefore, it can be concluded that Ti3C2 QDs act as an auxiliary catalyst, attract the generated electrons and provide active sites for the CO2 reduction reaction. As shown in Fig. 18E for photocatalytic performance, a high yield is shown for the formation of methanol (CH3OH). Therefore, CH3OH production by Ti3C2 QDs/Cu2O NWs/Cu showed 8.25 times and 2.15 times higher yield than Cu2O NWs/Cu and Ti3C2 sheet/Cu2O NWs/Cu after 6 h of reaction, respectively. As shown in Fig. 18F and G, this catalyst also showed high selectivity for CH3OH and improved CO2 conversion efficiency, where its high selectivity can be attributed to the essential role of Ti3C2 QDs in promoting light absorption, charge separation, transport and carrier density, as identified by electrochemical impedance spectroscopy and Mott–Schottky measurements. The above-mentioned examples show the promising properties of Ti-MXene for highly selective photocatalytic CO2 conversion, which demonstrates the future-oriented development of these materials in both theoretical and computational terms. Therefore, combining Ti3C2 auxiliary catalysts with various photocatalysts requires more advanced and intensive research. In addition, the simulations of these structures will assist the in-depth understanding of the approach of Ti3C2 towards adsorption, activation, and charge transfer mechanisms. V vs. NHE at pH = 7 and the standard reduction potential of H2O/O2 is 1.23 V vs. NHE and of ˙OH/H2O is 2.3 V vs. NHE. Under light irradiation, the activated catalyst of CuFe and the photon electrode electron can be transferred toward the conductive Ti3C2 due to its higher work function. The interaction of CuFe and MXene in the Ti3C2 MXene CuFe complex has the effect of promoting the separation and transportation of photo-excited carriers on the CuFe mixed metal oxide surface, which results in the efficient reduction of CO2. The generated electrons can be used to reduce CO2 to activate radicals. The analysis results showed that the structure and relationship between the light source and the catalyst supplied had a significant impact on the CO2RR and alcohol production rates.


image file: d4se00405a-f18.tif
Fig. 18 (A) Schematic depiction of the preparation method and (B) TEM image of Ti3C2 QD/Cu2O NW/Cu heterostructure. (C) HRTEM image of Ti3C2 QDs/Cu2O NWs/Cu. (D) PEC-chemical performance of Cu2O NWs/Cu, Ti3C2 sheets/Cu2O NWs/Cu, and Ti3C2 QDs/Cu2O NWs/Cu. (E) Yield of methanol with time. (F) Nyquist plots from EIS. G) Mott–Schottky plots.155

6. Conclusion and perspectives

In conclusion, the photocatalytic performance, crystal structure, and charge transfer characteristics of photocatalysts differ according to their synthesis method. Herein, the photocatalytic materials widely used for the photocatalytic reduction of CO2 were presented based on the principle aspects of photocatalytic CO2 reduction, such as thermodynamics, material transfer, selectivity, and reaction mechanisms. In addition, to further improve their CO2 reduction performance, the possible ways for their improvement were suggested in the following two directions. Exploring the composition and structure of new materials will continue to be the core of research on CO2 reduction by electrocatalysts and photocatalysts. Essentially, all existing photocatalysts for water splitting can be converted to CO2 reduction photocatalysts if a strategy can be applied to significantly convert the side reaction selectivity from CO2 to CO2RR by introducing an appropriate cocatalyst. The bonding of semiconductors to MXene substrates, charge transfer properties, band gaps, recombination coefficients, and conditions of the reduction process are the main factors in the photocatalytic reduction of carbon dioxide. The electrochemical reduction of CO2 is a process involving the movement of many electrons and protons. In this case, give that the solubility of CO2 in the electrolyte is low, effectively transferring CO2 to the anode surface in the actual CO2 electrolytic bath is a key reaction process to achieve both a high current density and high CO2 reduction selectivity. In this review, we aimed to provide references for researchers in this field by summarizing the recent progress on MXene-based catalysts for photocatalytic CO2 reduction. Undoubtedly, MXene-based catalysts exhibit increased photocatalytic activity due to their extraordinary properties. This reaction shows great potential in improving the selectivity of the target product and lowering the cost of the catalyst. However, the research on MXene-based catalysts is still in its infancy. Furthermore, there are many problems to be solved before their large-scale commercialization. The problem of quality improvement and unit cost reduction of MXene used in MXene-based photocatalysts should be addressed first. It is necessary to expand the type of MXenes used for photocatalytic CO2 reduction in the future. Most of the currently used MXenes are manufactured based on Ti3C2 MXene, and thus it is urgent to manufacture other types of MXenes such as V2C, Zr2C, and Mo2C. This can provide more choices for their applications in the future, and nitrogen-based MXenes are characterized by having higher electrical conductivity than their corresponding carbide precursors. It can be predicted that nitrogen-based MXenes may be more promising cocatalysts for CO2 reduction by photoelectric catalysts. However, MXenes are known to be easily oxidized into the corresponding metal oxides in the presence of water and oxygen. More seriously, radicals such as ·OH and ·O2 and oxygen produced during photocatalytic CO2 reduction with a strong oxidizing capacity are expected to oxidize MXene easily. Therefore, improving the antioxidant ability of MXene-based photocatalysts is an important factor in future studies. Also, MXenes can be an inefficient factor in effectively removing pollutants from water because of their reactivity in more complex environments. Thus, it is essential to investigate their reactivity and interactions in complex multi-component fractions to fully evaluate their environmental performance. To achieve the effective application of MXenes in CO2 reduction, a thorough investigation of pH and the dissolved amount of CO2 is required. In addition, in the case of MXenes, their ability to reduce recycling and regeneration should be considered. Furthermore, given that photocatalytic CO2 reduction systems typically contain different carbon-containing parts that can be degraded during photocatalytic processes, they are one of the important factors in verifying the carbon source of the product from CO2 reduction by designing a suitable 13CO2 isotope experiment.

Conflicts of interest

The authors declare no conflict of interest.

References

  1. S. Back, M. S. Yeom and Y. Jung, Active Sites of Au and Ag Nanoparticle Catalysts for CO2 Electroreduction to CO, ACS Catal., 2015, 5, 5089 CrossRef CAS .
  2. A. Salehi-Khojin, H. M. Jhong, B. A. Rosen, W. Zhu, S. Ma, P. J. A. Kenis and R. I. Masel, Nanoparticle Silver Catalysts That Show Enhanced Activity for Carbon Dioxide Electrolysis, J. Phys. Chem. C, 2013, 117, 1627 CrossRef CAS .
  3. M. Zeng and Y. Li, Recent advances in heterogeneous electrocatalysts for the hydrogen evolution reaction, J. Mater. Chem. A, 2015, 3, 14942 RSC .
  4. M. Asadi, B. Kumar, A. Behranginia, B. A. Rosen, A. Baskin, N. Repnin, D. Pisasale, P. Phillips, W. Zhu, R. Haasch, R. F. Klie, P. Kral, J. Abiade and A. Salehi-Khojin, Robust carbon dioxide reduction on molybdenum di-sulphide edges, Nat. Commun., 2014, 5, 4470 CrossRef CAS PubMed .
  5. G. Xi, S. Ouyang, P. Li, J. Ye, Q. Ma, N. Su, H. Bai and C. Wang, Ultrathin W18O49 Nanowires with Diameters below 1 nm: Synthesis, Near-Infrared Absorption, Photoluminescence, and Photochemical Reduction of Carbon Dioxide, Angew. Chem., Int. Ed., 2012, 51, 2395 CrossRef CAS PubMed .
  6. T. Wang, X. Meng, G. Liu, K. Chang, P. Li, Q. Kang, L. Liu, M. Li, S. Ouyang and J. Ye, In situ synthesis of ordered mesoporous Co-doped TiO2 and its enhanced photocatalytic activity and selectivity for the reduction of CO2, J. Mater. Chem. A, 2015, 3, 9491 RSC .
  7. V. P. Indrakanti, J. D. Kubicki and H. H. Schobert, Photoinduced activation of CO2 on Ti-based heterogeneous catalysts: Current state, chemical physics-based insights and outlook, Energy Environ. Sci., 2009, 2, 745 RSC .
  8. Z. Zhang, Y. Huang, K. Liu, L. Guo, Q. Yuan and B. Dong, Multichannel-Improved Charge-Carrier Dynamics in Well-Designed Hetero-nanostructural Plasmonic Photocatalysts toward Highly Efficient Solar-to-Fuels Conversion, Adv. Mater., 2015, 27, 5906 CrossRef CAS PubMed .
  9. R. Reske, H. Mistry, F. Behafarid, B. Roldan Cuenya and P. Strasser, Particle Size Effects in the Catalytic Electroreduction of CO2 on Cu Nanoparticles, J. Am. Chem. Soc., 2014, 136, 6978 CrossRef CAS PubMed .
  10. M. P. Pechini, Method of Preparing Lead and Alkaline-Earth Titanates and Niobates and Coating Method Using the Same to Form a Capacitor, US Pat., 3.330.697, 1967 Search PubMed .
  11. X. Chen, S. Shen, L. Guo and S. S. Mao, Semiconductor-based Photocatalytic Hydrogen Generation, Chem. Rev., 2010, 110, 6503 CrossRef CAS PubMed .
  12. Y. Sohn, W. Huang and F. Taghipour, Recent progress and perspectives in the photocatalytic CO2 reduction of Ti-oxide-based nanomaterials, Appl. Surf. Sci., 2017, 396, 1696 CrossRef CAS .
  13. J. Pan, G. Liu, G. Q. M. Lu and H. Cheng, On the True Photo reactivity Order of {001}, {010}, and {101} Facets of Anatase TiO2 Crystals, Angew. Chem., Int. Ed., 2011, 50, 2133 CrossRef CAS PubMed .
  14. B. R. Eggins, J. T. Irvine, E. P. Murphy and J. Grimshaw, Formation of two-carbon acids from carbon dioxide by photoreduction on cadmium sulphide, J. Chem. Soc., Chem. Commun., 1988, 16, 1123 RSC .
  15. B. AlOtaibi, X. Kong, S. Vanka, S. Y. Woo, A. Pofelski, F. Oudjedi, S. Fan, M. G. Kibria, G. A. Botton, W. Ji, H. Guo and Z. Mi, Photochemical Carbon Dioxide Reduction on Mg-Doped Ga(In)N Nanowire Arrays under Visible Light Irradiation, ACS Energy Lett., 2016, 1, 246 CrossRef CAS .
  16. Y. Izumi, Recent advances in the photocatalytic conversion of carbon dioxide to fuels with water and/or hydrogen using solar energy and beyond, Coord. Chem. Rev., 2013, 257, 171 CrossRef CAS .
  17. X. Luo, Z. Yin, M. Zeng and M. Kurmoo, The construction, structures, and functions of pillared layer metal–organic frameworks, Inorg. Chem. Front., 2016, 3, 1208 RSC .
  18. J. Sun, G. Chen, J. Wu, H. Dong and G. Xiong, Bismuth vanadate hollow spheres: Bubble template synthesis and enhanced photocatalytic properties for photodegradation, Appl. Catal., B, 2013, 132, 304–314 CrossRef .
  19. Y. Li, J. Liu, X. Huang and G. Li, Hydrothermal synthesis of Bi2WO6 uniform hierarchical microspheres, Cryst. Growth Des., 2007, 7, 1350–1355 CrossRef CAS .
  20. R. Buonsanti, V. Grillo, E. Carlino, C. Giannini, T. Kipp, R. Cingolani and P. D. Cozzoli, Nonhydrolytic synthesis of high-quality anisotropically shaped brookite TiO2 nanocrystals, J. Am. Chem. Soc., 2008, 130, 11223–11233 CrossRef CAS PubMed .
  21. X. Wang, C. Wang, W. Jiang, W. Guo and J. Wang, Sonochemical synthesis and characterization of Cl-doped TiO2 and its application in the photodegradation of phthalate ester under visible light irradiation, Chem. Eng. J., 2012, 189, 288–294 CrossRef .
  22. P. Saha, S. Amanullah and A. Dey, Selectivity in Electrochemical CO2 Reduction, Acc. Chem. Res., 2022, 55, 134 CrossRef CAS PubMed .
  23. D. Bhalothia, W.-H. Hsiung, S.-S. Yang, C. Yan, P. Chen, T. Lin, S. Wu, P. Chen, K. Wang, M. Lin and T. Chen, Submillisecond Laser Annealing Induced Surface and Subsurface Restructuring of Cu–Ni–Pd Trimetallic Nanocatalyst Promotes Thermal CO2 Reduction, ACS Appl. Energy Mater., 2021, 4, 14043 CrossRef CAS .
  24. S. Kondaveeti, I. M. Abu-Reesh, G. Mohanakrishna, M. Bulut and D. Pant, Advanced Routes of Biological and Bio-electrocatalytic Carbon Dioxide (CO2) Mitigation Toward Carbon Neutrality, Front. Energy Res., 2020, 8 DOI:10.3389/fenrg.2020.00094 .
  25. A. Wagner, C. D. Sahm and E. Reisner, Towards molecular understanding of local chemical environment effects in electro- and photocatalytic CO2 reduction, Nat. Catal., 2020, 3, 775 CrossRef CAS .
  26. X. Li, W. He, C. Li, B. Song and S. Liu, Synergetic surface modulation of ZnO/Pt@ZIF-8 hybrid nanorods for enhanced photocatalytic CO2 valorization, Appl. Catal., B, 2021, 287, 119934,  DOI:10.1016/j.apcatb.2021.119934 .
  27. M. Tahir, A. A. Khan, S. Tasleem, R. Mansoor and W. K. Fan, Titanium Carbide (Ti3C2) MXene as a Promising Co-catalyst for Photocatalytic CO2 Conversion to Energy-Efficient Fuels: A Review, Energy Fuels, 2021, 35, 10374 CrossRef CAS .
  28. P. Kuang, J. Low, B. Cheng, J. Yu and J. Fan, MXene-based photocatalysts, J. Mater. Sci. Technol., 2020, 56, 18,  DOI:10.1016/j.jmst.2020.02.037 .
  29. M. Li, C. Bian, G. Yang and X. Qiang, Facile fabrication of water-based and non-fluorinated superhydrophobic sponge for efficient separation of immiscible oil/water mixture and water-in-oil emulsion, Chem. Eng. J., 2019, 368, 350–358 CrossRef CAS .
  30. J. Mao, K. Li and T. Peng, Recent advances in the photocatalytic CO2 reduction over semiconductors, Catal. Sci. Technol., 2013, 3, 2481–2498 RSC .
  31. L. Xiao, C. Yuan, P. Chen, Y. Liu, J. Sheng, S. Zhang, F. Dong and Y. Sun, Cu–S Bonds as an Atomic-Level Transfer Channel to Achieve Photocatalytic CO2 Reduction to CO on Cu-Substituted ZnIn2S4, ACS Sustainable Chem. Eng., 2022, 10, 11902 CrossRef CAS .
  32. X. Li, J. Wen, J. Low, Y. Fang and J. Yu, Design and fabrication of semiconductor photocatalyst for photocatalytic reduction of CO2 to solar fuel, Sci. China Mater., 2014, 57, 70 CrossRef .
  33. J. Ran, M. Jaroniec and S. Z. Qiao, Cocatalysts in Semiconductor-based Photocatalytic CO2 Reduction: Achievements, Challenges, and Opportunities, Adv. Mater., 2018, 30 DOI:10.1002/adma.201704649 .
  34. L. Yuan and Y.-J. Xu, Photocatalytic conversion of CO2 into value-added and renewable fuels, Appl. Surf. Sci., 2015, 342, 154–167 CrossRef CAS .
  35. H. Takeda, K. Koike, H. Inoue and O. Ishitani, Development of an efficient photocatalytic system for CO2 reduction using rhenium (i) complexes based on mechanistic studies, J. Am. Chem. Soc., 2008, 130, 2023–2031 CrossRef CAS PubMed .
  36. A. J. Morris, G. J. Meyer and E. Fujita, Molecular approaches to the photocatalytic reduction of carbon dioxide for solar fuels, Acc. Chem. Res., 2009, 42, 1983–1994 CrossRef CAS PubMed .
  37. J. Hawecker, J.-M. Lehn and R. Ziessel, Efficient photochemical reduction of CO2 to CO by visible light irradiation of systems containing re (bipy)(CO)3X or Ru(bipy)32+–CO2+ combinations as homogeneous catalysts, J. Chem. Soc. Chem. Commun., 1983, 536–538 RSC .
  38. K. Sayama, Solar hydrogen production on photocatalysis-electrolysis hybrid system using redox mediator and porous oxide photoelectrodes, Solar to Chemical Energy Conversion, Springer, Berlin/Heidelberg, Germany, 2016, pp. 345–365 Search PubMed .
  39. J. Yang, D. Wang, H. Han and C. Li, Roles of co-catalysts in photocatalysis and photoelectrocatalytic, Acc. Chem. Res., 2013, 46, 1900–1909 CrossRef CAS PubMed .
  40. J. Hong, W. Zhang, J. Ren and R. Xu, Photocatalytic reduction of CO2: A brief review on product analysis and systematic methods, Anal. Methods, 2013, 5, 1086–1097 RSC .
  41. S. Neatu, J. A. Maciá-Agulló and H. Garcia, Solar light photocatalytic CO2 reduction: General considerations and selected bench-mark photocatalysts, Int. J. Mol. Sci., 2014, 15, 5246–5262 CrossRef PubMed .
  42. K. Hashimoto, H. Irie and A. Fujishima, TiO2 photocatalysis: A historical overview and future prospects, Jpn. J. Appl. Phys., 2005, 44, 8269 CrossRef CAS .
  43. P. Wang, B. Huang, Y. Dai and M. H. Whangbo, Plasmonic photocatalysts: Harvesting visible light with noble metal nanoparticles, Phys. Chem. Chem. Phys., 2012, 14, 9813–9825 RSC .
  44. M. Naguib, V. N. Mochalin, M. W. Barsoum and Y. Gogotsi, 25th anniversary article: MXenes: a new family of two-dimensional materials, Adv. Mater., 2014, 26, 992 CrossRef CAS PubMed .
  45. M. Naguib, M. W. Barsoum and Y. Gogotsi, Ten Years of Progress in the Synthesis and Development of MXenes, Adv. Mater., 2021, 33, 2103393 CrossRef CAS PubMed .
  46. J. Low, L. Zhang, T. Tong, B. Shen and J. Yu, TiO2/MXene Ti3C2 composite with excellent photocatalytic CO2 reduction activity, J. Catal., 2018, 361, 255–266,  DOI:10.1016/j.jcat.2018.03.009 .
  47. R. Khaledialidusti, M. Khazaei, S. Khazaei and K. Ohno, High-throughput computational discovery of ternary-layered MAX phases and prediction of their exfoliation for formation of 2D MXenes, Nanoscale, 2021, 13, 7294 RSC .
  48. M. Khazaei, M. Arai, T. Sasaki, M. Estili and Y. Sakka, Trends in electronic structures and structural properties of MAX phases: a first-principles study on M2AlC (M = Sc, Ti, Cr, Zr, Nb, Mo, Hf, or Ta), M2AlN, and hypothetical M2AlB phases, J. Phys.: Condens. Matter, 2014, 26, 505503 CrossRef PubMed .
  49. M. Naguib, O. Mashtalir, J. Carle, V. Presser, J. Lu, L. Hultman, Y. Gogotsi and M. W. Barsoum, Two-Dimensional Transition Metal Carbides, ACS Nano, 2012, 6, 1322 CrossRef CAS PubMed .
  50. M. Naguib, M. Kurtoglu, V. Presser, J. Lu, J. Niu, M. Heon, L. Hultman, Y. Gogotsi and M. W. Barsoum, Two-dimensional nanocrystals produced by exfoliation of Ti3 AlC2, Adv. Mater., 2011, 23, 4248 CrossRef CAS PubMed .
  51. O. Mashtalir, M. Naguib, V. N. Mochalin, Y. Dall'Agnese, M. Heon, M. W. Barsoum and Y. Gogotsi, Intercalation and delamination of layered carbides and carbonitrides, Nat. Commun., 2013, 4, 1716 CrossRef PubMed .
  52. J. Halim, M. R. Lukatskaya, K. M. Cook, J. Lu, C. R. Smith, L. Å. Näslund, S. J. May, L. Hultman, Y. Gogotsi, P. Eklund and M. W. Barsoum, Transparent Conductive Two-Dimensional Titanium Carbide Epitaxial Thin Films, Chem. Mater., 2014, 26, 2374 CrossRef CAS PubMed .
  53. M. Ghidiu, M. R. Lukatskaya, M. Zhao, Y. Gogotsi and M. W. Barsoum, Conductive two-dimensional titanium carbide ‘clay’ with high volumetric capacitance, Nature, 2014, 516, 78 CrossRef CAS PubMed .
  54. M. Naguib, R. R. Unocic, B. L. Armstrong and J. Nanda, Large-scale delamination of multi-layers transition metal carbides and carbonitrides “MXenes”, Dalton Trans., 2015, 44, 9353 RSC .
  55. F. Shahzad, M. Alhabeb, C. B. Hatter, B. Anasori, S. Man Hong, C. M. Koo and Y. Gogotsi, Electromagnetic interference shielding with 2D transition metal carbides (MXenes), Science, 2016, 353, 1137 CrossRef CAS PubMed .
  56. P. Urbankowski, B. Anasori, K. Hantanasirisakul, L. Yang, L. Zhang, B. Haines, S. J. May, S. J. L. Billinge and Y. Gogotsi, 2D molybdenum and vanadium nitrides synthesized by ammoniation of 2D transition metal carbides (MXenes), Nanoscale, 2017, 9, 17722 RSC .
  57. Y. Li, H. Shao, Z. Lin, J. Lu, L. Liu, B. Duployer, P. O. Å. Persson, P. Eklund, L. Hultman, M. Li, K. Chen, X.-H. Zha, S. Du, P. Rozier, Z. Chai, E. Raymundo-Piñero, P.-L. Taberna, P. Simon and Q. Huang, A general Lewis acidic etching route for preparing MXenes with enhanced electrochemical performance in non-aqueous electrolyte, Nat. Mater., 2020, 19, 894 CrossRef CAS PubMed .
  58. J. Mei, G. A. Ayoko, C. Hu and Z. Sun, Thermal reduction of sulfur-containing MAX phase for MXene production, Chem. Eng. J., 2020, 395, 125111 CrossRef CAS .
  59. W. Sun, S. A. Shah, Y. Chen, Z. Tan, H. Gao, T. Habib, M. Radovic and M. J. Green, Electrochemical etching of Ti2AlC to Ti2CTx (MXene) in low-concentration hydrochloric acid solution, J. Mater. Chem. A, 2017, 5, 21663 RSC .
  60. T. Li, L. Yao, Q. Liu, J. Gu, R. Luo, J. Li, X. Yan, W. Wang, P. Liu, B. Chen, W. Zhang, W. Abbas, R. Naz and D. Zhang, Fluorine-Free Synthesis of High-Purity Ti3 C2 Tx (T=OH, O) via Alkali Treatment, Angew. Chem., Int. Ed., 2018, 57, 6115 CrossRef CAS PubMed .
  61. H. Shi, P. Zhang, Z. Liu, S. Park, M. R. Lohe, Y. Wu, A. Shaygan Nia, S. Yang and X. Feng, Ambient-Stable Two-Dimensional Titanium Carbide (MXene) Enabled by Iodine Etching, Angew. Chem., Int. Ed., 2021, 60, 8689 CrossRef CAS PubMed .
  62. M. Alhabeb, K. Maleski, T. S. Mathis, A. Sarycheva, C. B. Hatter, S. Uzun, A. Levitt and Y. Gogotsi, Selective Etching of Silicon from Ti3SiC2 (MAX) To Obtain 2D Titanium Carbide (MXene), Angew. Chem., Int. Ed., 2018, 57, 5444 CrossRef CAS PubMed .
  63. H. Deng, Z. Li, L. Wang, L. Yuan, J. Lan, Z. Chang, Z. Chai and W. Shi, Nanolayered Ti3C2 and SrTiO3 Composites for Photocatalytic Reduction and Removal of Uranium(VI), ACS Appl. Nano Mater., 2019, 2, 2283 CrossRef CAS .
  64. Y. Gao, L. Wang, A. Zhou, Z. Li, J. Chen, H. Bala, Q. Hu and X. Cao, Hydrothermal synthesis of TiO2/Ti3C2 nanocomposites with enhanced photocatalytic activity, Mater. Lett., 2015, 150, 62 CrossRef CAS .
  65. X. Lu, J. Zhu, W. Wu and B. Zhang, Hierarchical architecture of PANI@TiO2/Ti3C2Tx ternary composite electrode for enhanced electrochemical performance, Electrochim. Acta, 2017, 228, 282 CrossRef CAS .
  66. W. Zhou, J. Zhu, F. Wang, M. Cao and T. Zhao, One-step synthesis of Ceria/Ti3C2 nanocomposites with enhanced photocatalytic activity, Mater. Lett., 2017, 206, 237 CrossRef CAS .
  67. P. Tian, X. He, L. Zhao, W. Li, W. Fang, H. Chen, F. Zhang, Z. Huang and H. Wang, Enhanced charge transfer for efficient photocatalytic H2 evolution over UiO-66-NH2 with annealed Ti3C2Tx MXenes, Int. J. Hydrogen Energy, 2019, 44, 788 CrossRef CAS .
  68. H. Fang., Y. Pan, M. Yin, L. Xu, Y. Zhu and C. Pan, Facile synthesis of ternary Ti3C2–OH/ln2S3/CdS composite with efficient adsorption and photocatalytic performance towards organic dyes, J. Solid State Chem., 2019, 280, 120981 CrossRef CAS .
  69. H. Fang, Y. Pan, M. Yin and C. Pan, Enhanced photocatalytic activity and mechanism of Ti3C2−OH/Bi2WO6:Yb3+, Tm3+ towards degradation of RhB under visible and near infrared light irradiation, Mater. Res. Bull., 2020, 121, 110618 CrossRef CAS .
  70. Z. Li, H. Zhang, L. Wang., M. Xiangchao., J. Shi., C. Qi., Z. Zhang, L. Feng and C. Li, 2D/2D BiOBr/Ti3C2 heterojunction with dual applications in both water detoxification and water splitting, J. Photochem. Photobiol., A, 2020, 386, 112099 CrossRef CAS .
  71. Y. Li, Z. Yin, G. Ji, Z. Liang, Y. Xue, Y. Guo, J. Tian, X. Wang and H. Cui, 2D/2D/2D heterojunction of Ti3C2 MXene/MoS2 nanosheets/TiO2 nanosheets with exposed (001) facets toward enhanced photocatalytic hydrogen production activity, Appl. Catal., B, 2019, 246, 12 CrossRef CAS .
  72. S. Sun, M. Wang, X. Chang, Y. Jiang, D. Zhang, D. Wang, Y. Zhang and Y. Lei, W18O49/Ti3C2Tx Mxene nanocomposites for highly sensitive acetone gas sensor with low detection limit, Sens. Actuators, B, 2020, 304, 127274 CrossRef CAS .
  73. N. Hao, Y. Wei, J. Wang, Z. Wang, Z. Zhu, S. Zhao, M. Han and X. Huang, In-situ hybridization of an MXene/TiO2/NiFeCo-layered double hydroxide composite for electrochemical and photoelectrochemical oxygen evolution, RSC Adv., 2018, 8, 20576 RSC .
  74. L. Tie, S. Yang, C. Yu, H. Chen, Y. Liu, S. Dong, J. Sun and J. Sun, In-situ decoration of ZnS nanoparticles with Ti3C2 MXene nanosheets for efficient photocatalytic hydrogen evolution, J. Colloid Interface Sci., 2019, 545, 63 CrossRef CAS PubMed .
  75. Y. Li, D. Zhang, X. Feng, Y. Liao, Q. Wen and Q. Xiang, Truncated octahedral bipyramidal TiO2/MXene Ti3C2 hybrids with enhanced photocatalytic H2 production activity, Nanoscale Adv., 2019, 1, 1812 RSC .
  76. X. Ding, Y. Li, C. Li, W. Wang, L. Wang, L. Feng and D. Han, 2D visible-light driven TiO2@Ti3C2/g-C3N4 ternary heterostructure for high photocatalytic activity, J. Mater. Sci., 2019, 54, 9385 CrossRef CAS .
  77. C. Su, B. Y. Hong and C. M. Tseng, Sol–gel preparation and photocatalysis of titanium dioxide, Catal. Today, 2004, 96, 119 CrossRef CAS .
  78. M. A. Iqbal, S. I. Ali, F. Amin, A. Tariq, M. Z. Iqbal and S. Rizwan, La- and Mn-Codoped Bismuth Ferrite/Ti3C2 MXene Composites for Efficient Photocatalytic Degradation of Congo Red Dye, ACS Omega, 2019, 4, 8661 CrossRef CAS PubMed .
  79. J. B. Rivest and P. K. Jain, exchange on the nanoscale: an emerging technique for new material synthesis, device fabrication, and chemical sensing, Chem. Soc. Rev., 2013, 42, 89 RSC .
  80. Y. Li, X. Deng, J. Tian, Z. Liang and H. Cui, Ti3C2 MXene-derived Ti3C2/TiO2 nanoflowers for noble-metal-free photocatalytic overall water splitting, Appl. Mater. Today, 2018, 13, 217 CrossRef .
  81. Q. Huang, Y. Liu, T. Cai and X. Xia, Simultaneous removal of heavy metal ions and organic pollutant by BiOBr/Ti3C2 nanocomposite, J. Photochem. Photobiol., A, 2019, 375, 201 CrossRef CAS .
  82. H. Zhang, M. Li, J. Cao, Q. Tang, P. Kang, C. Zhu and M. Ma, 2D α-Fe2O3 doped Ti3C2 MXene composite with enhanced visible light photocatalytic activity for degradation of Rhodamine B, Ceram. Int., 2018, 44, 19958 CrossRef CAS .
  83. T. Cai, L. Wang, Y. Liu, S. Zhang, W. Dong, H. Chen, X. Yi, J. Yuan, X. Xia, C. Liu and S. Luo, Ag3PO4/Ti3C2 MXene interface materials as a Schottky catalyst with enhanced photocatalytic activities and anti-photocorrosion performance, Appl. Catal., B, 2018, 239, 545 CrossRef CAS .
  84. Y. Yang, Z. Zeng, G. Zeng, D. Huang, R. Xiao, C. Zhang, C. Zhoua, W. Xiong, W. Wang, M. Cheng, W. Xue, H. Guo, X. Tang and D. He, Ti3C2 Mxene/porous g-C3N4 interfacial Schottky junction for boosting spatial charge separation in photocatalytic H2O2 production, Appl. Catal., B, 2019, 258, 117956 CrossRef CAS .
  85. N. Liu, N. Lu, Y. Su, P. Wang and X. Quan, Fabrication of g-C3N4/Ti3C2 composite and its visible-light photocatalytic capability for ciprofloxacin degradation, Sep. Purif. Technol., 2019, 211, 782 CrossRef CAS .
  86. H. Zhang, M. Li, C. Zhu, Q. Tang, P. Kang and J. Cao, Preparation of magnetic α- Fe2O3/ZnFe2O4@Ti3C2 MXene with excellent photocatalytic performance, Ceram. Int., 2020, 46(1), 81–88 CrossRef CAS .
  87. Y. Zhuang, Y. Liu and X. Meng, Fabrication of TiO2 nanofibers/MXene Ti3C2 nanocomposites for photocatalytic H2 evolution by electrostatic self-assembly, Appl. Surf. Sci., 2019, 496, 143647 CrossRef CAS .
  88. G. Huang, S. Li, L. Liu, L. Zhu and Q. Wang, Ti3C2 MXene-modified Bi2WO6 nanoplates for efficient photodegradation of volatile organic compounds, Appl. Surf. Sci., 2020, 503, 144183 CrossRef CAS .
  89. Y. Fang, Y. Cao and Q. Chen, Synthesis of an Ag2WO4/Ti3C2 Schottky composite by electrostatic traction and its photocatalytic activity, Ceram. Int., 2019, 45, 22298 CrossRef CAS .
  90. P. Lin, J. Shen, X. Yu, Q. Liu, D. Li and H. Tang, Construction of Ti3C2 MXene/O-doped g-C3N4 2D-2D Schottky-junction for enhanced photocatalytic hydrogen evolution, Ceram. Int., 2019, 45, 24656 CrossRef CAS .
  91. Y. Li, W. Wang, Z. Zhan, M. H. Woo, C. Y. Wu and P. Biswas, Photocatalytic reduction of CO2 with H2O on mesoporous silica supported Cu/TiO2 catalysts, Appl. Catal., B, 2010, 100, 386–392 CrossRef CAS .
  92. S. C. Yan., S. X. Ouyang., J. Gao, M. Yang., J. Y. Feng, X. X. Fan, L. J. Wan, Z. S. Li, J. H. Ye and Y. Zhou, A room-temperature reactive-template route to mesoporous ZnGa2O4 with improved photocatalytic activity in reduction of CO2, Angew. Chem., 2010, 122, 6544–6548 CrossRef .
  93. Q. Liu, Y. Zhou, J. Kou, X. Chen, Z. Tian, J. Gao, S. Yan and Z. Zou, High-yield synthesis of ultralong and ultrathin Zn2GeO4 nanoribbons toward improved photocatalytic reduction of CO2 into renewable hydrocarbon fuel, J. Am. Chem. Soc., 2010, 132, 14385–14387 CrossRef CAS PubMed .
  94. K. Iizuka, T. Wato, Y. Miseki, K. Saito and A. Kudo, Photocatalytic reduction of carbon dioxide over Ag co-catalyst-loaded ALa4Ti4O15 (A = Ca, Sr, and Ba) using water as a reducing reagent, J. Am. Chem. Soc., 2011, 133, 20863–20868 CrossRef CAS PubMed .
  95. Q. Zhang, Y. Li, E. A. Ackerman, M. Gajdardziska-Josifovska and H. Li, Visible light responsive iodine-doped TiO2 for photocatalytic reduction of CO2 to fuels, Appl. Catal., A, 2011, 400, 195–202 CrossRef CAS .
  96. M. Stock and S. Dunn, LiNbO3—A polar material for solid-gas artificial photosynthesis, Ferroelectrics, 2011, 419, 9–13 CrossRef CAS .
  97. W. Tu, Y. Zhou, Q. Liu, Z. Tian, J. Gao, X. Chen, H. Zhang, J. Liu and Z. Zou, Robust hollow spheres consisting of alternating titania nanosheets and graphene nanosheets with high photocatalytic activity for CO2 conversion into renewable fuels, Adv. Funct. Mater., 2012, 22, 1215–1221 CrossRef CAS .
  98. S. Wang, B. Guan and X. Lou, Xiong Wen David Lou*Construction of ZnIn2S4–In2O3 Hierarchical Tubular Heterostructures for Efficient CO2 Photoreduction, J. Am. Chem. Soc., 2018, 140(15), 5037–5040 CrossRef CAS PubMed .
  99. L. Ran, Z. Li, B. Ran, J. Cao, Y. Zhao, T. Shao, Y. Song, M. Leung, L. Sun and J. Hou, Engineering Single-Atom Active Sites on Covalent Organic Frameworks for Boosting CO2 Photoreduction, J. Am. Chem. Soc., 2022, 144(37), 17097–17109 CrossRef CAS PubMed .
  100. K. Wang, M. Cheng, N. Wang, Q. Zhang, Y. Liu, J. Liang, J. Guan, M. Liu, J. Zhou and N. Li, Inter-plane 2D/2D ultrathin La2Ti2O7/Ti3C2 MXene Schottky heterojunctions toward high-efficiency photocatalytic CO2 reduction, Chin. J. Catal., 2022, 44, 146–159,  DOI:10.1016/S1872-2067(22)64155-X .
  101. W. Wu, H. Bi, Z. Zhang, L. Sun, R. Wei, L. Gao, X. Pan, J. Zhang and G. Xiao, Z-scheme π-π stacking MXene/GO/PDI composite aerogels to construct interface electron transport network for photocatalytic CO2 reduction, Colloids Surf., A, 2023, 620, 130486,  DOI:10.1016/j.colsurfa.2022.130486 .
  102. Y. Xu, F. Wang, S. Lei, Y. Wei, D. Zhao, Y. Gao, X. Ma, S. Li, S. Chang, M. Wang and H. Jing, In-situ grown two-dimensional TiO2/Ti3CN MXene heterojunction rich in Ti3+ species for highly efficient photoelectrocatalytic CO2 reduction, Chem. Eng. J., 2023, 452, 139392,  DOI:10.1016/j.cej.2022.139392 .
  103. K. Wang, Q. Wang, K. Zhang, G. Wang and H. Wang, Selective solar-driven CO2 reduction mediated by 2D/2D Bi2O2SiO3/MXene nanosheets heterojunction, J. Mater. Sci. Technol., 2022, 124, 202–208,  DOI:10.1016/j.jmst.2021.10.059 .
  104. Y. Wang, R. Du, Z. Li, H. Song, Z. Chao, D. Zu, D. Chong, N. Gao and C. Li, Rationally designed CdS/Ti3C2 MXene electrocatalysts for efficient CO2 reduction in aqueous electrolyte, Ceram. Int., 2021, 47(20), 28321–28327,  DOI:10.1016/j.ceramint.2021.06.249 .
  105. M. Madi, M. Tahir and Z. Y. Zakaria, 2D/2D V2C mediated porous g-C3N4 heterojunction with the role of monolayer/multilayer MAX/MXene structures for stimulating photocatalytic CO2 reduction to fuels, J. CO2 Util., 2022, 65, 102238,  DOI:10.1016/j.jcou.2022.102238 .
  106. X. Li, Y. Bai, X. Shi, J. Huang, K. Zhang, R. Wang and L. Ye, Mesoporous g-C3N4/MXene (Ti3C2Tx) heterojunction as a 2D electronic charge transfer for efficient photocatalytic CO2 reduction, Appl. Surf. Sci., 2021, 546, 149111,  DOI:10.1016/j.apsusc.2021.149111 .
  107. M. Tahir and B. Tahir, In-situ growth of TiO2 imbedded Ti3C2TA nanosheets to construct PCN/Ti3C2TA MXenes 2D/3D heterojunction for efficient solar driven photocatalytic CO2 reduction towards CO and CH4 production, J. Colloid Interface Sci., 2021, 591, 20–37,  DOI:10.1016/j.jcis.2021.01.099 .
  108. J. Shen, J. Shen, W. Zhang, X. Yu, H. Tang, M. Zhang, Zulfiqar and Q. Liu, Built-in electric field induced CeO2/Ti3C2-MXene Schottky-junction for coupled photocatalytic tetracycline degradation and CO2 reduction, Ceram. Int., 2019, 45(18), 24146–24153,  DOI:10.1016/j.ceramint.2019.08.123 .
  109. F. He, B. Zhu, B. Cheng, J. Yu, W. Ho and W. Macyk, 2D/2D/0D TiO2/C3N4/Ti3C2 MXene composite S-scheme photocatalyst with enhanced CO2 reduction activity, Appl. Catal., B, 2020, 272(5), 119006,  DOI:10.1016/j.apcatb.2020.119006 .
  110. A. Saeed, W. Chen, A. H. Shah, Y. Zhang, I. Mehmood and Y. Liu, Enhancement of photocatalytic CO2 reduction for novel Cd0.2Zn0.8S@Ti3C2 (MXenes) nanocomposites, J. CO2 Util., 2021, 47, 101501,  DOI:10.1016/j.jcou.2021.101501 .
  111. Q. Tang, P. Xiong, H. Wang and Z. Wu, Boosted CO2 photoreduction performance on Ru-Ti3CN MXene-TiO2 photocatalyst synthesized by non-HF Lewis acidic etching method, J. Colloid Interface Sci., 2022, 619, 179–187,  DOI:10.1016/j.jcis.2022.03.137 .
  112. G. Bharath, K. Rambabu, A. Hai, I. Othman, N. Ponpandian, F. Banat and P. L. Show, Hybrid Pd50-Ru50/MXene (Ti3C2Tx) nanocatalyst for effective hydrogenation of CO2 to methanol toward climate change control, Chem. Eng. J., 2021, 414, 128869,  DOI:10.1016/j.cej.2021.128869 .
  113. M. Tahir and B. Tahir, 2D/2D/2D O-C3N4/Bt/Ti3C2Tx heterojunction with novel MXene/clay multi-electron mediator for stimulating photo-induced CO2 reforming to CO and CH4, Chem. Eng. J., 2020, 400, 125868,  DOI:10.1016/j.cej.2020.125868 .
  114. J. Zhang, J. Shi, S. Tao, L. Wu and J. Lu, Cu2O/Ti3C2MXene heterojunction photocatalysts for improved CO2 photocatalytic reduction performance, Appl. Surf. Sci., 2021, 542, 148685,  DOI:10.1016/j.apsusc.2020.148685 .
  115. J. Yang, Y. Hou, J. Sun, T. Liang, T. Zhu, J. Liang, X. Lu, Z. Yu, H. Zhu and S. Wang, Cu doping and Ti3C2OH quantum dots co-modifications of Zn3In2S6 boosted photocatalytic reduction of CO2 to CO via accelerating carriers transfer and enhancing CO2 adsorption and activation, Chem. Eng. J., 2023, 452, 139522,  DOI:10.1016/j.cej.2022.139522 .
  116. L. Li, Y. Yang, B. Zhou, Y. Zhou and Z. Zou, Dimensional matched ultrathin BiVO4/Ti3C2Tx heterosystem for efficient photocatalytic conversion of CO2 to methanol, Mater. Lett., 2022, 306, 130937,  DOI:10.1016/j.matlet.2021.130937 .
  117. Z. Du, H. Cai, Z. Guo, Z. Z Shao, J. Lin, Y. Huang, C. Tang, G. Chen and Y. Fang, Synergistic photocatalytic of CO2-to-CO conversion by 2D/1D Ti3C2Tx/p-BN heterojunction with interfacial chemical bonding, J. Alloys Compd., 2022, 920, 165933,  DOI:10.1016/j.jallcom.2022.165933 .
  118. Q. Zhai, S. Xie, W. Fan, Q. Zhang, Y. Wang, W. Deng and Y. Wang, Photocatalytic conversion of carbon dioxide with water into methane: Platinum and copper (i) oxide co-catalysts with a core–shell structure, Angew. Chem., 2013, 125, 5888–5891 CrossRef .
  119. J. L. Giocondi and G. S. Rohrer, Spatial separation of photochemical oxidation and reduction reactions on the surface of ferroelectric BaTiO3, J. Phys. Chem. B, 2001, 105, 8275–8277 CrossRef CAS .
  120. S. V. Kalinin, D. A. Bonnell, T. Alvarez, X. Lei, Z. Hu, J. Ferris, Q. Zhang and S. Dunn, Atomic polarization and local reactivity on ferroelectric surfaces: A new route toward complex nanostructures, Nano Lett., 2002, 2, 589–593 CrossRef CAS .
  121. H. Yang, S. Qin, H. Wang and J. Lu, Organically doped palladium: a highly efficient catalyst for electroreduction of CO2 to methanol, Green Chem., 2015, 17, 5144–5148 RSC .
  122. J. H. Q. Lee, S. J. L. Lauw and R. D. Webster, The electrochemical reduction of carbon dioxide (CO2) to methanol in the presence of pyridoxine (vitamin B6), Electrochem. Commun., 2016, 64, 69–73 CrossRef CAS .
  123. J. Bi, P. Hou, F. Liu and P. Kang, Electrocatalytic reduction of CO2 to methanol by iron tetradentate phosphine complex through amidation strategy, ChemSusChem, 2019, 12, 2195–2201 CrossRef CAS PubMed .
  124. M. Liu, Y. J. Zhang, K. Cheng, X. Quan, X. F. Fan, Y. Su, S. Chen, H. M. Zhao, Y. B. Zhang, H. T. Yu and M. R. Hoffmann, Selective electrochemical reduction of carbon dioxide to ethanol on a boron- and nitrogen-Co-doped nano diamond, Angew. Chem., Int. Ed., 2017, 56, 15607–15611 CrossRef PubMed .
  125. Y. Yan, E. L. Zeitler, J. Gu, Y. Hu and A. B. Bocasly, Electrochemistry of aqueous pyridinium: exploration of a key aspect of electrocatalytic reduction of CO2 to methanol, J. Am. Chem. Soc., 2013, 135, 14020–14023 CrossRef CAS PubMed .
  126. S. Sun, W. Wang, L. Zhang, M. Shang and L. Wang, Ag@C core/shell nanocomposite as a highly efficient plasmonic photocatalyst, Catal. Commun., 2009, 11, 290–293 CrossRef CAS .
  127. P.-Y. Olu, Q. Li and K. Krischer, The true fate of pyridinium in the reportedly pyridinium-catalyzed carbon dioxide electroreduction on platinum, Angew. Chem., Int. Ed., 2018, 57, 14769–14772 CrossRef CAS PubMed .
  128. K. Awazu, M. Fujimaki, C. Rockstuhl, J. Tominaga, H. Murakami, Y. Ohki, N. Yoshida and T. Wantanabe, A Plasmonic Photocatalyst Consisting of Silver Nanoparticles Embedded in Titanium Dioxide, J. Am. Chem. Soc., 2008, 130(5), 1676–1680 CrossRef CAS PubMed .
  129. J. R. Bolton, S. J. Strickler and J. S. Connolly, Limiting and realizable efficiencies of solar photolysis of water, Nature, 1985, 316, 495 CrossRef CAS .
  130. S. Feng, X. Chen, Y. Zhou, W. Tu, P. Li, H. Li and Z. Zou, Na2V6O16·xH2O nanoribbons: large-scale synthesis and visible-light photocatalytic activity of CO2 into solar fuels, Nanoscale, 2014, 6, 1896–1900 RSC .
  131. B. Qin, Y. Li, H. Wang, G. Yang, Y. Cao, H. Yu, Q. Zhang, H. Liang and F. Peng, Efficient electrochemical reduction of CO2 into CO promoted by sulfur vacancies, Nano Energy, 2019, 60, 43–51 CrossRef CAS .
  132. H. Wang, L. Zhang, K. Wang, X. Sun and W. Wang, Enhanced photocatalytic CO2 reduction to methane over WO3·0.33H2O via Mo doping, Appl. Catal., B, 2019, 243, 771–779 CrossRef CAS .
  133. J. Albo, M. Alvarez-Guerra, P. Castaño and A. Irabien, Towards the electrochemical conversion of carbon dioxide into methanol, Green Chem., 2015, 17, 2304–2324 RSC .
  134. M. Subrahmanyam, S. Kaneco and N. Alonso-Vante, A screening for the photo reduction of carbon dioxide supported on metal oxide catalysts for C1–C3 selectivity, Appl. Catal., B, 1999, 23, 169–174 CrossRef CAS .
  135. N. Sasirekha, S. J. S. Basha and K. Shanthi, Photocatalytic performance of Ru doped anatase mounted on silica for reduction of carbon dioxide, Appl. Catal., B, 2006, 62, 169–180 CrossRef CAS .
  136. E. Karamian and S. Sharifnia, On the general mechanism of photocatalytic reduction of CO2, J. CO2 Util., 2016, 16, 194–203 CrossRef CAS .
  137. J.-Y. Liu, X.-Q. Gong and A. N. Alexandria, Mechanism of CO2 photocatalytic reduction to methane and methanol on defected anatase TiO2 (101): a density functional theory study, J. Phys. Chem. C, 2019, 123(6), 3505–3511 CrossRef CAS .
  138. P. Sharma, S. Kumar, O. Tomanec, M. Petr, J. Z. Chen, J. T. Miller, R. S. Varma, M. B. Gawande and R. Zbořil, Carbon Nitride-Based Ruthenium Single Atom Photocatalyst for CO2 Reduction to Methanol, Small, 2021, 17, 2006478 CrossRef CAS PubMed .
  139. M. Morikawa, Y. Ogura, N. Ahmed, S. Kawamura, G. Mikami, S. Okamoto and Y. Izumi, Photocatalytic conversion of carbon dioxide into methanol in reverse fuel cells with tungsten oxide and layered double hydroxide photocatalysts for solar fuel generation, Catal. Sci. Technol., 2014, 4, 1644 RSC .
  140. P. Periasamy, R. Ravindranath, S. M. S. Kumar, W. Wu, T. Jian and H. Chang, Facet- and Structure-Dependent Catalytic Activity of Cuprous Oxide/Polypyrrole Particles Towards the Efficient Reduction of Carbon Dioxide to Methanol, Nanoscale, 2018, 10, 25 RSC .
  141. S. Mou, T. Wu, J. Xie, Y. Zhang, L. Ji, H. Hung, T. Wang, Y. Luo, X. Xiong, B. Tang and X. Sun, Boron Phosphide Nanoparticles: A Nonmetal Catalyst for High-Selectivity Electrochemical Reduction of CO2 to CH3OH, Adv. Mater., 2019, 31(36), 1903499 CrossRef PubMed .
  142. M. Alhabeb, K. Maleski, B. Anasori, P. Lelyukh, L. Clark, S. Sin and Y. Gogotsi, Guidelines for synthesis and processing of two-dimensional titanium carbide (Ti3C2TX MXene), Chem. Mater., 2017, 29(18), 7633–7644 CrossRef CAS .
  143. B. M. Jun, S. Kim, J. Heo, N. Her, M. Jang, C. M. Park and Y. Yoon, Enhanced sonocatalytic degradation of carbamazepine and salicylic acid using a metalorganic framework, Ultrason. Sonochem., 2019, 56, 174–182 CrossRef CAS PubMed .
  144. S. N. Xiao, D. L. Pan, R. Liang, W. R. Dai, Q. T. Zhang, G. Q. Zhang, C. L. Su, H. X. Li and W. Chen, Bimetal MOF derived mesocrystal ZnCo2O4 on rGO with high performance in visible-light photocatalytic NO oxidization, Appl. Catal., B, 2018, 236, 304–313 CrossRef CAS .
  145. Y. L. Sun, X. Meng, Y. Dall'Agnese, C. Dall'Agnese, S. N. Duan, Y. Gao, G. Chen and X. F. Wang, 2D MXenes as co-catalysts in photocatalysis: synthetic methods, Nano-Micro Lett., 2019, 11(1), 79 CrossRef CAS PubMed .
  146. M. H. Ye, X. Wang, E. Z. Liu, J. H. Ye and D. F. Wang, Boosting the photocatalytic activity of P25 for carbon dioxide reduction by using a surface-alkalinized titanium carbide MXene as cocatalyst, ChemSusChem, 2018, 11(10), 1606–1611 CrossRef CAS PubMed .
  147. J. R. Ran, G. P. Gao, F. T. Li, T. Y. Ma, A. J. Du and S. Z. Qiao, Ti3C2 MXene co-catalyst on metal sulfide photo-absorbers for enhanced visible-light photocatalytic hydrogen production, Nat. Commun., 2017, 8, 13907 CrossRef CAS PubMed .
  148. C. Peng, H. J. Wang, H. Yu and F. Peng, (111) TiO2-x/Ti3C2: synergy of active facets, interfacial charge transfer and Ti3p doping for enhance photocatalytic activity, Mater. Res. Bull., 2017, 89, 16–25 CrossRef CAS .
  149. A. D. Handoko., H. Chen, Y. Lum, Q. Zhang, B. Anasori and Z. W. She, Two-Dimensional Titanium and Molybdenum Carbide MXenes as Electrocatalysts for CO2 Reduction, iScience, 2020, 23, 101181 CrossRef CAS PubMed .
  150. B. Anasori, M. R. Lukatskaya and Y. Gogotsi, 2D metal carbides and nitrides (MXenes) for energy storage, Nat. Rev. Mater., 2017, 2, 16098 CrossRef CAS .
  151. J. Hu, J. Ding and Q. Zhong, Ultrathin 2D Ti3C2 MXene Co-catalyst anchored on porous g-C3N4 for enhanced photocatalytic CO2 reduction under visible-light irradiation, J. Colloid Interface Sci., 2021, 582, 647 CrossRef CAS PubMed .
  152. J. Li, Z. Wang, H. Chen, Q. Zhang, H. Hu, L. Liu, J. Ye and D. Wang, A surface-alkalinized Ti3C2 MXene as an efficient cocatalyst for enhanced photocatalytic CO2 reduction over ZnO, Catal. Sci. Technol., 2021, 11, 4953 RSC .
  153. Y. Gao, K. Qian, B. Xu, Z. Li, J. Zheng, S. Zhao, F. Ding, Y. Sun and Z. Xu, Recent advances in visible-light-driven conversion of CO2 by photocatalysts into fuels or value-added chemicals, Carbon Resour. Convers., 2020, 3, 46–59 CrossRef CAS .
  154. S. Zhang, W. Xiong, J. Long., Y. Si., Y. Xu., L. Yang., J. Zou, W. Dai, X. Luo. and S. Luo, High-throughput lateral and basal interface in CeO2@Ti3C2TX: Reverse and synergistic migration of carrier for enhanced photocatalytic CO2 reduction, J. Colloid Interface Sci., 2022, 615, 716 CrossRef CAS PubMed .
  155. Z. Zeng, Y. Yan, J. Chen, P. Zan, Q. Tian and P. Chen, Boosting the Photocatalytic Ability of Cu2O Nanowires for CO2 Conversion by MXene Quantum Dots, Adv. Funct. Mater., 2019, 29, 1806500 CrossRef .
  156. S. Cao, B. Shen, T. Tong, J. Fu and J. Yu, 2D/2D Heterojunction of Ultrathin MXene/Bi2WO6 Nanosheets for Improved Photocatalytic CO2 Reduction, Adv. Funct. Mater., 2018, 28, 1800136 CrossRef .
  157. L. Li, Y. Yang, L. Yang, X. Wang, Y. Zhou and Z. Zou, 3D Hydrangea-like InVO4/Ti3C2Tx Hierarchical Heterosystem Collaborating with 2D/2D Interface Interaction for Enhanced Photocatalytic CO2 Reduction, ChemNanoMat, 2021, 7, 815 CrossRef CAS .
  158. J. Bai, R. Shen, Z. Jiang, P. Zhang, Y. Li and X. Li, Integration of 2D layered CdS/WO3 S-scheme heterojunctions and metallic Ti3C2 MXene-based Ohmic junctions for effective photocatalytic H2 generation, Chin. J. Catal., 2022, 43(2), 359–369 CrossRef CAS .
  159. L. Wang, X. Fei, L. Zhang, J. Yu, B. Cheng and Y. Ma, Solar fuel generation over nature-inspired recyclable TiO2/g-C3N4 S-scheme hierarchical thin-film photocatalyst, J. Mater. Sci. Technol., 2022, 112, 1–10 CrossRef CAS .
  160. K. Wang, X. Li, N. Wang, Q. Shen, M. Liu, J. Zhou and N. Li, Z-Scheme Core–Shell meso-TiO2@ZnIn2S4/Ti3C2 MXene Enhances Visible Light-Driven CO2-to-CH4 Selectivity, Ind. Eng. Chem. Res., 2021, 60, 8720 CrossRef CAS .
  161. Z. Sun, H. Wang, Z. Wu and L. Wang, g-C3N4 based composite photocatalysts for photocatalytic CO2 reduction, Catal. Today, 2018, 300, 160 CrossRef CAS .
  162. Q. Xu, Z. Xia, J. Zhang, Z. Wei, Q. Guo, H. Jin, H. Tang, S. Li, X. Pan, Z. Su and S. Wang, Recent advances in solar-driven CO2 reduction over g-C3N4-based photocatalysts, Carbon Energy, 2023, 5, e205 CrossRef CAS .
  163. C. Yang, Q. Tan, Q. Li, J. Zhou, J. Fan, B. Li, J. Sun and K. Lv, 2D/2D Ti3C2 MXene/g-C3N4 nanosheets heterojunction for high efficient CO2 reduction photocatalyst: Dual effects of urea, Appl. Catal., B, 2020, 268, 118738 CrossRef CAS .
  164. H. Wang, Q. Tang and Z. Wu, Construction of Few-Layer Ti3C2 MXene and Boron-Doped g-C3N4 for Enhanced Photocatalytic CO2 Reduction, ACS Sustainable Chem. Eng., 2021, 9, 8425 CrossRef CAS .
  165. Q. Song, J. Hu, Y. Zhou, Q. Ye, X. Shi, D. Li and D. Jiang, Carbon vacancy-mediated exciton dissociation in Ti3C2Tx/g-C3N4 Schottky junctions for efficient photoreduction of CO2, J. Colloid Interface Sci., 2022, 623, 487 CrossRef CAS PubMed .
  166. Q. Tang, Z. Sun, S. Deng, H. Wang and Z. Wu, Decorating g-C3N4 with alkalinized Ti3C2 MXene for promoted photocatalytic CO2 reduction performance, J. Colloid Interface Sci., 2020, 564, 406 CrossRef CAS PubMed .
  167. R. Ikreedeeeegh and M. Tahir, A critical review in recent developments of metal-organic-frameworks (MOFs) with band engineering alteration for photocatalytic CO2 reduction to solar fuels, J. CO2 Util., 2021, 43, 101381 CrossRef .
  168. Y. Yang, D. Zhang, J. Fan, Y. Liao and Q. Xiang, Construction of an Ultrathin S-Scheme Heterojunction Based on Few-Layer g-C3N4 and Monolayer Ti3C2Tx MXene for Photocatalytic CO2 Reduction, Sol. RRL, 2021, 5, 2000351 CrossRef CAS .
  169. Y. Xiao, C. Men, B. Chu, Z. Qin, H. Ji, J. Chen and T. Su, Spontaneous reduction of copper on Ti3C2Tx as fast electron transport channels and active sites for enhanced photocatalytic CO2 reduction, Chem. Eng. J., 2022, 446, 137028 CrossRef .
  170. X. Chen, W. G. Pan, L. F. Hong, X. Hu, J. Wang, Z. Bi and R. Guo, Ti3C2-modified g-C3N4/MoSe2 S-Scheme Heterojunction with Full-Spectrum Response for CO2 Photoreduction to CO and CH4, ChemSusChem, 2023, 16, e202300179 CrossRef CAS PubMed .
  171. Z. Otgonbayar and W.-C. Oh, Synthesis of 2D/2D structural Ti3C2 MXene/g-C3N4 via the Schottky junction with metal oxides: Photocatalytic CO2 reduction with a cationic scavenger, Appl. Mater. Today, 2023, 32, 101814 CrossRef .
  172. Q. Shi, X. Zhang, Y. Yang, J. Huang, X. Fu, T. Wang, X. Liu, A. Sun, J. Ge, J. Shen, Y. Zhou and Z. Liu, 3D hierarchical architecture collaborating with 2D/2D interface interaction in NiAl-LDH/Ti3C2 nanocomposite for efficient and selective photoconversion of CO2, J. Energy Chem., 2021, 59, 9 CrossRef CAS .
  173. S. Zhao, D. Pan, Q. Liang, M. Zhou, C. Yao, S. Xu and Z. Li, Ultrathin NiAl-Layered Double Hydroxides Grown on 2D Ti3C2Tx MXene to Construct Core–Shell Heterostructures for Enhanced Photocatalytic CO2 Reduction, J. Phys. Chem. C, 2021, 125, 10207 CrossRef CAS .
  174. B. Zhou, Y. Yang, Z. Liu, N. Wu, Y. Yan, Z. Wenhua, H. He, J. Du, Y. Zhang, Y. Zhou and Z. Zou, Boosting photocatalytic CO2 reduction via Schottky junction with ZnCr layered double hydroxide nanoflakes aggregated on 2D Ti3C2Tx cocatalyst, Nanoscale, 2022, 14, 7538 RSC .
  175. W. Chen, B. Han, Y. Xie, S. Liang, H. Deng and Z. Lin, Ultrathin Co-Co LDHs nanosheets assembled vertically on MXene: 3D nanoarrays for boosted visible-light-driven CO2 reduction, Chem. Eng. J., 2020, 391, 123519 CrossRef CAS .
  176. X. Y. Xiong, Y. F. Zhao, R. Shi, W. J. Yin, Y. X. Zhao, G. I. N. Waterhouse and T. R. Zhang, Selective photocatalytic CO2 reduction over Zn-based layered double hydroxides containing tri or tetravalent metals, Sci. Bull., 2020, 65, 987–994 CrossRef CAS PubMed .
  177. H. Shen, T. Ouyang, J. Guo, M. Mu and X. Yin, A perspective LDHs/Ti3C2O2 design by DFT calculation for photocatalytic reduction of CO2 to C2 organics, Appl. Surf. Sci., 2023, 609, 155445 CrossRef CAS .
  178. F. Xu, K. Meng, B. Cheng, S. Wang, J. Xu and J. Yu, Unique S-scheme heterojunctions in self-assembled TiO2/CsPbBr3 hybrids for CO2 photoreduction, Nat. Commun., 2020, 11, 4613 CrossRef CAS PubMed .
  179. M. Que, Y. Zhao, Y. Yang, L. Pan, W. Lei, W. Cai, H. Yuan, J. Chen and G. Zhu, Anchoring of Formamidinium Lead Bromide Quantum Dots on Ti3C2 Nanosheets for Efficient Photocatalytic Reduction of CO2, ACS Appl. Mater. Interfaces, 2021, 13, 6180 CrossRef CAS PubMed .
  180. M. Que, W. Cai, Y. Zhao, Y. Yang, B. Zhang, S. Yun, J. Chen and G. Zhu, 2D/2D Schottky heterojunction of in-situ growth FAPbBr3/Ti3C2 composites for enhancing photocatalytic CO2 reduction, J. Colloid Interface Sci., 2022, 610, 538 CrossRef CAS PubMed .
  181. A. Pan, X. Ma, S. Huang, Y. Wu, M. Jia, Y. Shi, Y. Liu, P. Wangyang, L. He and Y. Liu, CsPbBr3 Perovskite Nanocrystal Grown on MXene Nanosheets for Enhanced Photoelectric Detection and Photocatalytic CO2 Reduction, J. Phys. Chem. Lett., 2019, 10, 6590 CrossRef CAS PubMed .
  182. X. Li, J. Liu, G. Jiang, X. Lin, J. Wang and Z. Li, Self-supported CsPbBr3/Ti3C2Tx MXene aerogels towards efficient photocatalytic CO2 reduction, J. Colloid Interface Sci., 2023, 643, 174 CrossRef CAS PubMed .
  183. Z. Zhang, B. Wang, H.-B. Zhao, J.-F. Liao, Z.-C. Zhou, T. Liu, B. He, Q. Wei, S. Chen, H.-Y. Chen, D.-B. Kuang, Y. Li and G. Xing, Self-assembled lead-free double perovskite-MXene heterostructure with efficient charge separation for photocatalytic CO2 reduction, Appl. Catal., B, 2022, 312, 121358 CrossRef CAS .
  184. Q. Song, B. Shen, J. Yu and S. Cao, A 3D Hierarchical Ti3C2Tx/TiO2 Heterojunction for Enhanced Photocatalytic CO2 Reduction, ChemNanoMat, 2021, 7, 910 CrossRef CAS .
  185. J. Li, K. Li, Q. Tan, Q. Li, J. Fan, C. Wu and K. Lv, Facile Preparation of Highly Active CO2 Reduction (001) TiO2/Ti3C2Tx Photocatalyst from Ti3AlC2 with Less Fluorine, Catalysts, 2022, 12, 785 CrossRef CAS .
  186. H. Li, Q. Song, S. Wan, C.-W. Tung, C. Liu, Y. Pan, G. Luo, H. M. Chen, S. Cao, J. Yu and L. Zhang, Atomic Interface Engineering of Single-Atom Pt/TiO2-Ti3C2 for Boosting Photocatalytic CO2 Reduction, Small., 2023, 19, 2301711 CrossRef CAS PubMed .
  187. K. C. Devarayapalli, S. V. P. Vattikuti, D. J. Kim, Y. Lim, B. Kim, G. Kim and D. S. Lee, Platinum quantum dots-decorated MXene-derived titanium dioxide nanowire/Ti3C2 heterostructure for use in solar-driven gas-phase carbon dioxide reduction to yield value-added fuels, J. Energy Chem., 2023, 82, 627 CrossRef .
  188. H. Liu, K. Chen, Y.-N. Feng, Z. Zhuang, F.-F. Chen and Y. Yu, In Situ Confined Growth of Co3O4–TiO2/C S-Scheme Nanoparticle Heterojunction for Boosted Photocatalytic CO2 Reduction, J. Phys. Chem., 2023, 127, 5289 CAS .
  189. Y. Wu, W. Xu, W. Tang, Z. Wang, Y. Wang, Z. Lv, Y. Zhang, W. Zhong, H.-L. Cai, R. Yang and X. S. Wu, In-situ annealed “M-scheme” MXene-based photocatalyst for enhanced photoelectric performance and highly selective CO2 photoreduction, Nano Energy, 2021, 90, 106532 CrossRef CAS .
  190. L. Chen, K. Huang, Q. Xie, S. M. Lam, J. C. Sin, T. Su, H. Ji and Z. Qin, The enhancement of photocatalytic CO2 reduction by the in situ growth of TiO2 on Ti3C2 MXene, Catal. Sci. Technol., 2021, 11, 1602 RSC .
  191. P. Mane, I. V. Bagal, H. Bae, A. N. Kadam, V. Burungale, J. Heo, S. W. Ryu and J. S. Ha, Recent trends and outlooks on engineering of BiVO4 photoanodes toward efficient photoelectrochemical water splitting and CO2 reduction: A comprehensive review, Int. J. Hydrogen Energy, 2022, 47, 39796–39828 CrossRef CAS .
  192. Y. Hori, H. H. I. Wakebe, T. Tsukamoto and O. Koga, Electrocatalytic process OF CO selectivity in electrochemical reduction OF CO2 at metal electrodes in aqueous media, Electrochim. Acta, 1994, 39, 1833–1839 CrossRef CAS .
  193. H. Pang, T. Masuda and J. Ye, Semiconductor-based photoelectrochemical conversion of carbon dioxide: stepping towards artificial photosynthesis, Chem. - Asian J., 2018, 13, 127–142,  DOI:10.1002/asia.201701596 .
  194. S. Castro, J. Albo and A. Irabien, Photoelectrochemical reactors for CO2 utilization, ACS Sustainable Chem. Eng., 2018, 6, 15877–15894,  DOI:10.1021/acssuschemeng.8b03706 .
  195. X. Chang, T. Wang, P. Yang, G. Zhang and J. Gong, The development of Cocatalysts for photoelectrochemical CO2 reduction, Adv. Mater., 2019, 31, 1–13,  DOI:10.1002/adma.201804710 .
  196. A. U. Pawar, C. W. Kim, M. T. Nguyen-Le and Y. S. Kang, General review on the components and parameters of photoelectrochemical system for CO2 reduction with in situ analysis, ACS Sustainable Chem. Eng., 2019, 7, 7431–7455,  DOI:10.1021/acssuschemeng.8b06303 .
  197. Z. Otgonbayar, C. M. Yoon and W. C. Oh, Photoelectrocatalytic CO2 reduction with ternary nanocomposite of MXene (Ti3C2)-Cu2O-Fe3O4: Comprehensive utilization of electrolyte and light-wavelength, Chem. Eng. J., 2023, 464, 142716 CrossRef CAS .

This journal is © The Royal Society of Chemistry 2024