Next Article in Journal
Synthesis of Alkynyl Ketones by Sonogashira Cross-Coupling of Acyl Chlorides with Terminal Alkynes Mediated by Palladium Catalysts Deposited over Donor-Functionalized Silica Gel
Next Article in Special Issue
Energy Transport of Photocatalytic Carbon Dioxide Reduction in Optical Fiber Honeycomb Reactor Coupled with Trough Concentrated Solar Power
Previous Article in Journal
Activation Treatments and SiO2/Pd Modification of Sol–Gel TiO2 Photocatalysts for Enhanced Photoactivity under UV Radiation
Previous Article in Special Issue
Photocatalytic H2 Evolution, CO2 Reduction, and NOx Oxidation by Highly Exfoliated g-C3N4
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Layered Double Hydroxide (LDH) Based Photocatalysts: An Outstanding Strategy for Efficient Photocatalytic CO2 Conversion

1
Department of Chemical Engineering, COMSATS University Islamabad, Lahore Campus, 1.5 KM Defence Road, Off Raiwind Road, Lahore 54000, Pakistan
2
Department of Energy Science & Engineering, DGIST, 333 Techno Jungang-daero, Hyeonpung-eup, Dalseong-gun, Daegu 42988, Korea
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(10), 1185; https://doi.org/10.3390/catal10101185
Submission received: 15 September 2020 / Revised: 8 October 2020 / Accepted: 9 October 2020 / Published: 14 October 2020
(This article belongs to the Special Issue Photocatalytic Reduction of CO2)

Abstract

:
CO2 conversion to solar fuels/chemicals is an alluring approach for narrowing critical issues of global warming, environmental pollution, and climate change, caused by excess atmospheric CO2 concentration. Amongst various CO2 conversion strategies, photocatalytic CO2 conversion (PCC) is considered as a promising approach, which utilizes inexpensive sunlight and water with a photocatalyst material. Hence, development of an efficient and a stable photocatalyst is an essential activity for the respective scientific community to upscale the PCC research domain. Until today, metal oxides, such as TiO2, ZnO, etc., are categorized as standard photocatalysts because of their relative stability, abundant availability and low cost. However, their performance is tethered by limited light absorption and somewhat physical properties. Recently, layered double hydroxides (LDHs) have offered an exciting and efficient way for PCC due to their superb CO2 adsorption and moderate photocatalytic properties. The LDH based photocatalysts show marvelous physiochemical and electrical properties like high surface area, stability, and excellent conductivity. In the present review article, a summarized survey is portrayed regarding latest development for LDH based photocatalysts with a focus on synthesis strategies employing various photocatalyst materials, influencing parameters and possible mechanism involved in PCC to useful fuels and chemicals like CO, CH4, CH3OH, and H2.

Graphical Abstract

1. Introduction

Ever increasing energy demand for rapid industrialization/urbanization, being fulfilled mainly from fossil fuels, is continuously stimulating the alarming and severe threats of environmental pollution and global warming [1]. It is believed that amongst various global house gases (GHG), increased atmospheric CO2 gas concentration contributes to the major portion for the respective critical issue [2,3,4]. Industrialization and utilization of fossil fuels are considered as the prime driving factors for excessive CO2 levels. Therefore, in accordance with the adverse sentiment, it is imperative to adopt and promote an alternative technology that provides an efficient, renewable, and sustainable pathway to counterbalance excessive atmospheric CO2 level. In this regard, photocatalytic CO2 conversion (PCC) or photo-reduction to useful/value added chemicals (CO, CH4, C2H6, C2H5OH, C2H4, CH3OH, HCOOH, etc.) is one of the most engaging and alluring approaches. PCC is based on the concept of natural photosynthesis, employing a semiconductor material acting as a photocatalyst with a reducing agent (e.g., H2O/H2 etc.) for CO2 conversion to useful chemicals under the light irradiation, a free energy source [5,6,7]. Generally, PCC is considered as a research domain under the umbrella of artificial photosynthesis, targeting normalization of CO2 level and providing chemicals/fuels in a sustainable manner.
Since the pioneering work of Inoue et al., the photocatalysis field has been researched extensively with respect to materials [8]. As well-established, titanium dioxide (TiO2) is counted on as one of the best photocatalytic materials due to its specific characteristics of abundant availability, high stability, better charge lifetime, favorable surface area, non-toxicity, and cost effectiveness [9]. However, despite certain benefits, TiO2 also possesses the major drawback of a limited light absorption due to its wide band gap and limited CO2 surface adsorption. Until now, enormous efforts and researches have been done to achieve an efficient photocatalytic CO2 conversion on the account of process development, such as materials with broadened light absorption, higher surface area, enhanced CO2 adsorption, and photocatalytic reactors advancement [10,11,12,13,14]. For the case of photocatalytic materials development, certain strategies have been executed such as doping [15,16], composite/hybrid/heterojunction formation [17,18,19,20,21], nano-architectures [22,23,24], and coupling with noble metals or carbon-based materials [22,25,26]. All such approaches are oriented towards attaining outstanding photocatalytic CO2 conversion efficiency.
Recently, layered double hydroxide (LDH), an interesting class of layered anionic clays, has captivated the researchers’ interest for their development and application in a variety of research domains such as adsorption [27], photocatalysis [28], electrochemistry [29,30,31,32], and biomedical science [33]. Generally, LDHs are synthetically produced and rarely occur in nature. The exceptional properties of LDH, i.e., their superb adsorption capacity extended light absorption via compositional variation, cheap synthesis procedure, and unique layered morphology, making them suitable materials for photocatalysis applications [28]. Until now, a moderate number of researches and studies have been reported employing LDH regarding their synthesis approaches and electronic and physical properties influencing photocatalytic applications [32,34]. However, specific to PCC, there exist limited reports employing bare LDH, surface modified LDH, or hybrid LDH based photocatalysts.
In the present review, the recent progress of bare LDH and LDH based photocatalysts employed for photocatalytic CO2 conversion to useful chemicals/fuels are briefly overviewed. The synthesis procedure involved, key reaction conditions, PCC mechanism, and prime parameters promoting the improvement of photocatalytic performance are detailed. In short, the review is a brief overview of LDH based photocatalysts offering excellent performance for PCC.

2. Layered Double Hydroxide (LDH) Structure and Synthesis

LDH are mainly synthetic materials with their structure mimicking natural mineral hydrotalcite (Mg6Al2(OH)16)CO3. H2O was discovered in 1842 and synthetically prepared in 1942 [35]. The detailed structural investigation of the LDH was conducted by Allmann and co-authors in 1969 [36]. Hence, LDH are composed of layered structure and are also termed hydrotalcite-like compounds comprising brucite like metal hydroxide layers and water molecules balanced by interlayer anions. The commonly accepted formula representing the LDH is [ M 1 x 2 + M x 3 + ( O H ) 2 A ? ] A x +   [ A n ] x / n m H 2 O , where M2+ (e.g., Mg2+, Cu2+, Ni2+, Zn2+ etc.) and M3+ (e.g., Al3+, Fe3+, Ga3+ etc.) represents the divalent and trivalent cations, respectively. An− indicates the interlayer gallery ions such as CO32−, NO32−, Cl, etc. along with water molecules. The illustration of commonly acknowledged and endorsed LDH is shown in Figure 1.
It is well established that the composition of the LDH, their structure, and certain properties can be controlled and manipulated by the nature of the synthesis procedure. Primarily the synthesis procedure can be classified based on the phase involved, i.e., (1) solid phase and (2) liquid phase synthesis approach. Solid state synthesis approach mainly consists of mechanochemical process employing wet and dry milling methods [37,38]. Such approach offers benefit of less pollution in terms of liquids used but again is an expensive method, with a major drawback of structure distortion and nonuniform LDH structure. On the contrary, liquid phase synthesis approach is a commonly adopted and dominant strategy because of its easiness, cost effectiveness, and provision of the prime benefits of complete and uniform LDH structural morphology. The liquid phase synthesis approach is further branched into variety of synthesis methods based on the reaction type [39,40,41]. Some of important and commonly practiced liquid phase synthesis methods can be listed as follows:
I.
Co-precipitation method;
II.
Anion exchange method;
III.
Reconstruction/rehydration method;
IV.
Solution mixing method.
Amongst the above-mentioned methods, co-precipitation method is the most frequently adopted synthesis approach offering a direct method for the synthesis of LDH. Co-precipitation method practices slow addition of anionic solution to the divalent or trivalent metal cations solution, with addition of alkali or urea, resulting in hydrolysis and finally precipitation of LDH. The addition of base or urea in metal cations solution leads to co-precipitation of metallic salts via condensation of hexa-aqua metal complexes finally resulting in brucite-like metal cations layers with solvated interlamellar anions.
Anion exchange method employs the exchange of the intercalated anions by other targeted anions. In this method, already prepared LDH is subjected to stirring in the targeted anions solution, which are to be intercalated to achieve the required LDH. Moreover, when the co-precipitation method is not feasible, anion exchange method is utilized, e.g., when either metal cations or anions are unstable in alkaline solution or when there is chance of direct reaction between the cations and intercalated anions. Under these specific circumstances, anion exchange method offers viable alternative approach to synthesize the desired LDH by exchanging the intercalated anions with the target anions.
Another interesting synthesis method is the reconstruction method. Like shape memory materials, LDH materials also tend to return to their original structure. Generally, when the LDH materials are calcined at higher temperatures (400–500 °C), the intercalated anions and water molecules are completely erased, and hydroxides are converted to respective metal oxides. These metal oxides are immersed in water or targeted anion solution, and the metal hydroxide and intercalated layers of anions are regenerated, and LDH is reformed. Such process is very fruitful for intercalating the specific targeted anions in the interlayer gallery of the LDH.
Simple solution mixing and then subjecting it to high temperature and pressure conditions such as hydrothermal or solvothermal approaches are another effective mean to fabricate the LDH and/or LDH based composites with other semiconductor materials.
In addition to the abovementioned methods, i.e., hydrothermal [42], there exist a variety of synthesis methods such as sol–gel [43], and microemulsion methods [44], which can be adopted based on the required features and properties of the LDH.

3. Bare and Modified Layered Double Hydroxide (LDH) Photocatalysts

Layered double hydroxide (LDH) is mainly comprised of layered morphology, providing important aspects of (i) broader light absorption (generally in the visible light range) and (ii) better CO2 sorption capability because of their layered morphology. Moreover, the choice of metal cations within LDH can well tune the photocatalytic activity by influencing its light absorption and sorption attributes. Ahmed et al. [45] studied the effect of the metal cation for zinc-copper-M (III) LDH, where M stands for aluminum (Al) or gallium (Ga) metal. It was observed that samples composed of ternary metal ions, i.e., ZnCuAl and ZnCuGa LDH, show a red shift in their light absorption as compared to ZnAl and ZnGa LDHs. Hence, the insertion of Cu in both mentioned LDHs extends their absorption towards the visible range, which might be due to the generation of an electronic state induced by Cu ions, thus narrowing down the band gap. Under the UV–Vis photo irradiation and reactant mixture consisting of CO2 and H2, the ZnCuAl and ZnCuGa LDH produced mainly CH3OH as a key product component as compared to ZnAl and ZnGa LDHs, which mainly yielded CO as a key product. Hence, the selectivity of the product mainly from CO switches to CH3OH with the incorporation of the Cu in the LDH. The role of Cu addition is responsible for increased production of CH3OH as compared to CO. Cu sites within the LDH might be providing a binding site for CO2, thus interacting with photogenerated electrons, protons, and probably the CuI/CuII redox couple. The proposed mechanism is shown in Figure 2a, where the interlayer hydroxl group linked with Ga/Al and Cu ties with the CO2 molecule resulting in hydrogen carbonate, which undergoes a series of reduction reactions in presence of protons finally forming the CH3OH as a major product. Furthermore, the insertion of Cu narrows down the band gap and well, aligning it with the redox level of CO2/CH3OH as shown by energy level diagram in Figure 2b. Such a band gap alignment is favorable for producing CH3OH under light irradiation.
As investigated in their previous research [45], the interlayer space within LDH layers acts as an active site for the CO2 conversion to CH3OH with higher selectivity when Cu is inserted. Hence, extending their investigation with the modification of the interlayer space of the LDH for enhancing the photocatalytic performance via anionic species between the cationic LDH layers, Ahmed and co-authors [46] found an increased CH3OH formation rate. They found, during the synthesis process of Zn3Ga|CO3 and Zn1.5Cu1.5Ga|CO3 LDH, the (CO3)2− is replaced with [Cu(OH)4]2− anions by [CuCl4]2− ions hydrolyzed in the alkaline solution. It was observed with the intercalation of the [Cu(OH)4]2− anions that the light absorption is shifted towards the lower energy as compared to samples with (CO3)2− anions. Moreover, Cu insertion also exhibits the absorption in the visible range due to intrinsic property of Cu ions. The Zn3Ga|Cu(OH)4 LDH photocatalyst exhibits 5.9 times higher CH3OH production as compared to Zn3Ga|CO3 and similar to that of Zn1.5Cu1.5Ga|CO3. However, Zn1.5Cu1.5Ga|Cu(OH)4 shows an increase selectivity of CH3OH up to 88%. The investigation concludes that the modification of the respective LDH photocatalysts with the Cu(OH)4 anions instead of CO3 anions leads to increase of interlayer space between the LDH thus resulting in more reactive space. Secondly, the insertion of the Cu ions narrows the bandgap of the resulting LDH photocatalyst with more light absorption and thus better photocatalytic performance.
Another report investigates the role of the precursor anions for hydroxide sheets, for various LDH photocatalysts including Ni-Al, Zn-Al, Mg-In, and Ni-In [47]. The LDH were synthesized by both nitrates (LDH-NO3) and chlorides (LDH-Cl) and evaluated for their photocatalytic activity in terms of CO2 conversion to useful chemicals. It was observed that the LDH with Ni/Al ratio of 4 and synthesized using chlorides exhibited the maximum CO2 conversion to CO. It was noticed that except Ni-Al LDH-Cl, all other samples selectively generate H2; however, Ni-Al LDH-Cl mainly yielded CO from CO2 photo-reduction. The production rate of various products obtained from variety of LDH photocatalysts can be seen in Figure 3a. The reason for the selective formation of CO for Ni-Al LDH-Cl can be attributed to the co-catalytic behavior of the Ni species, which in turn leads to the enhancement of CO formation as the Ni ratio is increased. For the other LDH photocatalysts, H2 is mainly produced due to competition of H+ being reduced during the CO2 photo-reduction, a well-known competitive reaction phenomenon. Furthermore, when the Ni content is decreased within Ni-Al LDH-Cl, the CO formation is decreased, and H2 formation is increased, which endorses Ni species to act as a co-catalyst during the CO2 photo-reduction reaction. The Ni-Al LDH-NO3, on the contrary, showed the production of CH4 when employed for CO2 photo-reduction. The authors also found that Cl ions in the aqueous solution also act as hole scavengers, thus promoting the selective formation of CO instead of other products.
Continuing the role of Cl ions, Iguchi et al. researched the effect of Cl ions as a hole scavenger in CO2 photo-reduction employing Ni-Al LDH [48]. Various additives were added in the aqueous solution containing suspended LDH photocatalyst, bubbled with CO2 gas, and irradiated using 400 W high pressure mercury lamp. The additives were NaHCO3, Na2CO3, Na2SO4, NaNO3, and NaCl. It was observed that NaHCO3 and Na2CO3 exhibited a greater selectivity towards H2 generation due to favorable reduction of H+ ions generated in water. The Na2SO4 and NaNO3 did not contribute in the formation of either product, H2 or CO. However, when NaCl was used as an additive, a significant amount of CO was produced with higher selectivity as compared to all other additives. The selectivity along with production rate of H2 and CO with various inorganic additives can be viewed in Figure 3b (dark bar represents CO, grey bar H2 and circle represents selectivity towards CO formation). In order to understand the mechanism involved, DPD (N, N-dimethyl-p-phenylenediamine) test was performed. The DPD test indicated that when the NaCl was added to aqueous solution, the product was consisted of CO, H2 and HClO under photoirradiation as a reaction amongst CO2, H+ and Cl. Hence, the formation of HClO as an oxidant product during the photoreaction indicates that Cl is expected to be a strong reducing agent and is rapidly oxidized by holes in the solution leading to yield HClO and selective formation of CO.
Another report investigated the role of fluorinated surface of LDH photocatalysts and their role in the CO2 conversion enhancement [49]. The authors fabricated fluorinated Mg-Al and Ni-Al LDH photocatalysts and evaluated their photocatalytic performance for the CO2 conversions in aqueous solution. The fluorination was done by incorporating hexafluoroaluminate (AlF63−) units within the hydroxide sheets of the LDH. In addition, various samples of LDH were prepared by changing the concentration of fluorine precursor, with the optimized sample providing the maximum yield of CO as a main product of CO2 photo-reduction. It was observed and confirmed in authors previous works that the Ni-Al LDH is more efficient in selective formation of CO as compared to Mg-Al LDH as Ni species are supposed to act as co-catalyst, thus suppressing the formation of H2. However, the optimized fluorine content (which was estimated by ratio of fluorine precursor to the aluminum precursor of LDH and was found 7.5) significantly enhances the formation of CO with NaCl as an additive for the hole scavenger. Such an improvement is mainly associated with the increased surface area of the fluorinated LDH, which leads to higher sites for CO2 adsorption and thus increased CO2 photo-reduction to CO.
The work done by Flores et al. investigates the role of various Mg sources, organic precursor and a one-step microwave-hydrothermal method on the photocatalytic activity of Mg-Al LDH [50]. Two different salts of magnesium, i.e., magnesium acetate and magnesium nitrate, were used with organic urea as a source of CO3 and OH within LDH layers. The LDH photocatalysts were then subjected for two different times for microwave radiation for 20 and 40 min and at a temperature of 100 °C. All the samples were employed for photocatalytic CO2 conversion in gas as well as liquid phase reactions. It was observed that the crystallinity of the synthesized LDH was significantly influenced by the nature of the Mg precursor and microwave irradiation time. The greater microwave irradiation time leads to formation of a larger crystallite size along with decomposition of organic urea resulting in formation of NH4+, OH, H+, and CO32− ions, which will readily react with Mg2+ due to higher dipole moment, thus resulting in better crystallinity. When the prepared LDH photocatalysts were employed in liquid phase for CO2 conversion, the key product formed was CH3OH with the highest yield obtained from Mg-Al LDH prepared by respective acetate salt under 40 min microwave radiation. Hence, by increasing the microwave irradiation time, the crystallite size and crystallinity are improved, which leads to decreased impedance, whereas, the decomposition of acetate anions adsorbed on LDH surface in turn facilitated the electron transfer towards the CO2 sites within LDH. Such decomposition for LDH obtained by nitrates salts resulted in formation of NO3, which can occupy the CO2 active site. Figure 4a displays the CH3OH production by various LDH photocatalysts prepared in the reported work. When the optimized LDH photocatalysts were employed for gas phase reactions, the key products obtained were CO and CH4. Such formation of CO and CH4 under gas phase is more favorable due to its potential lowering than the CH3OH formation and deoxygenation of C-species with electrons. The band gap alignment of all the Mg-Al LDH prepared in the respective work is shown in Figure 4b.
A recent work done by Wang et al. presents synthesis of Co-Al LDH nanosheets for the conversion of very low concentration atmospheric CO2 and water to CH4 [51]. Furthermore, the investigation also focuses on the importance of alkaline OH group and divalent cobalt for efficient CO2 conversion to CH4. Figure 5a,b show the morphology of the prepared Co-Al LDH nanosheets, showing hexagonal sheet like morphology and thickness of 18 nm. The unique attribute was the uniform composition of Co, Al, and O elements. The UV–Vis absorption spectrum for the Co-Al LDH nanosheets suggests the light absorption of the synthesized photocatalyst lies in the visible range with an estimated band gap of 2.1 eV. Thus, the respective photocatalyst can capture the visible range of the light, which can probably lead to improve photocatalytic performance. The synthesized Co-Al LDH nanosheets were found to be active for 55 h of irradiation with atmospheric CO2 (400 ppm), yielding CH4 as a main product with rate of 4.3 µmol g−1 h−1, which is 13 times higher as compared to standard TiO2 sample, i.e., P25. The improved CH4 generation was mainly attributed to the enhanced CO2 adsorption by surface OH groups and unique effect of divalent cobalt. The divalent characteristic of cobalt was investigated by oxidation of the Co-Al LDH in oxygen atmosphere, and it was observed that the photocatalytic performance was sharply decreased when the LDH was oxidized. Moreover, the LDH photocatalyst exhibited a stable photocatalytic performance over five cycles of repeated reaction as shown in Figure 5c. The mechanistic view displaying the proposed reduction mechanism for CO2 conversion upon the surface of Co-Al LDH is shown in Figure 5d. The CO2 reduction is proposed to initiate with the chemisorption of CO2 and OH groups, which can bend linear CO2 molecule, thus activating them to form carbonate species. These carbonate species in the presence of electrons and protons undergo multistep proton assisted reaction forming CO, which is finally converted to CH4 and other useful chemicals.
In another work, Tokudome et al. reported synthesis of nanocrystalline Ni-Al LDH (20 nm) by simple homogenous supersaturation process induced via rapid increase of pH form concentrated aqueous salt solutions of precursors [52]. When Ni-Al LDH was employed for CO2 conversion in aqueous medium, the CO formation rate (selectivity of 80%) was found 7 times higher as compared to the standard Ni-Al LDH prepared by conventional method. The key for the improved photocatalytic performance was attributed to the metastable surface of the nano LDH due to high degree of supersaturation. Figure 6a shows the morphology of the nano LDH and standard LDH. Nano LDH mainly consists of 20 nm particle size in agglomerated form whereas the conventional LDH is mainly composed of various size ranges of platelets within a range of 40–200 nm. Figure 6b shows the photocatalytic CO2 conversion to CO in aqueous media. It was observed that nano LDH exhibited a significant increase in CO production as compared to standard LDH. This was mainly attributed to the surface affinity of the nano LDH towards the CO2 adsorption in gaseous phase. Such a specific surface nature is induced mainly due to synthesis process designed in the research work. Furthermore, for the purpose of comparison, various control samples were synthesized in order to investigate the role of morphology and crystallinity of nano LDH for CO2 conversion. However, it was observed that all those parameters, when improved, exhibited a decrease in photocatalytic performance; hence ultimately, the metastable surface of the nano LDH and its superb affinity towards CO2 adsorption can only be considered as the key to success of improved photocatalytic performance.
A recently published research work by Gao et al. presented an ultrathin Mg-Al LDH modified with Fe3O4 by a simple coprecipitation approach [53]. It was observed that Fe3O4/Mg-Al LDH (FMAL) exhibited significant improvement in the photocatalytic activity, which was due to synergetic effect of efficient separation of electron-hole pairs and reduction of resistance for charges transmission induced by Fe3O4 and ultrathin Mg-Al LDH, respectively. Figure 7a shows the transmission electron microscopy (TEM) image of Fe3O4/Mg-Al LDH, which exhibits thin 2D nanosheets of Mg-Al LDH with Fe3O4 nanoparticles loaded on nanosheets. The UV–Vis DRS for the Fe3O4/Mg-Al LDH is shown in Figure 7b. The light absorption spectrum covers visible region due to presence of Fe3O4 nanoparticles, which possess a narrow band gap and almost act like a conductor. Such conductive behavior leads to inhibition of photogenerated electron-hole recombination, thus contributing to improved photocatalytic performance. Figure 7c,d show the photocatalytic production of CO and CH4 by various Fe3O4/Mg-Al LDH samples (with varied content of Fe3O4). It was observed that the sample FMAL-10 with 10 wt.% of Fe3O4, exhibited the highest CO and CH4 yield, and hence, it was selected as an optimized sample. The key factors as explained above were supposed to be the efficient charge separation by the Fe3O4 content and the rapid transformation of the photoexcited charges towards the reactive sites on Fe3O4/Mg-Al LDH surface. Further increase of Fe3O4 content decreased the light transmission characteristic and hence reduced photocatalytic performance.
Recently, Bai et al. reported four different types of ultrathin MAl-LDH photocatalysts (u-MAl-LDH, where M stands for Mg2+, Ni2+, CO2+ and Zn2+) for CO2 conversion to mainly CO [54]. The synthesized u-MAl-LDH were tested under visible light irradiation (λ = 400–800 nm) and employed with a [Ru(bpy)3]Cl2·6H2O photosensitizer for photocatalytic CO2 conversion to CO. Amongst various samples synthesized, the u-CoAl-LDH showed the highest efficiency, Figure 8a shows the schematic view of proposed photocatalytic CO2 conversion employing u-CoAl-LDH with production of CO and H2 when irradiated with visible light (400–800 nm) and light at 600 nm wavelength. Figure 8b shows that the photocatalytic reactions were not initiated without [Ru(bpy)3]Cl2·6H2O and triethanolamine, thus endorsing the important roles of [Ru(bpy)3]Cl2·6H2O as a photosensitizer and triethanolamine as a hole scavenger for regeneration of LDH photocatalysts valence bands. Figure 8c illustrates the photocatalytic CO2 conversion products for all the synthesized u-MAl-LDH photocatalysts. It could be observed that all samples exhibit less amount of CO and H2 production as than u-CoAl-LDH, which exhibited the highest conversion rate. The u-NiAl-LDH sample also exhibited production of CH4 due to well-known selectivity of Ni towards CH4. The key reason for improved photocatalytic performance can be associated to (i) defects on the photocatalyst surface due to its ultrathin nature thus promoting the CO2 adsorption; (ii) improved light absorption, thus capturing the maximum of visible light region; and (iii) well matched energy levels of u-CoAl-LDH with [Ru(bpy)3]Cl2·6H2O photosensitizer, thus on light irradiation the photogenerated charges are drained towards the LDH surface where they can efficiently react with the adsorbed CO2.
Another work proposed by Xiong et al. presented the role of trivalent and tetravalent in Zn based LDH photocatalysts employed for photocatalytic CO2 conversion [55]. It was observed that choice of trivalent or tetravalent cation in the LDH photocatalyst alters the selectivity of the product. Authors synthesized a variety of ZnM-LDH (where M = Ti4+, Fe3+, Co3+, Ga3+, Al3+) and observed that ZnTi-LDH generates CH4 as main product, whereas ZnFe-LDH and ZnCo-LDH were unable to reduce CO2 and yield H2 by water splitting only. Moreover, ZnGa-LDH and ZnAl-LDH yielded CO as main product of CO2 photo-reduction. By help of in situ diffuse reflection, infrared Fourier transform spectroscopy (DRIFT), and computational calculations, it was revealed that the metal cations with d band level (ԑd) closer to the fermi level interact favorably with CO2, leading to its well adsorption, and hence, light irradiation converts the adsorbed CO2 into carbon containing products, i.e., CH4 and CO, whereas the metals with d band level far from the fermi level exhibited very less CO2 adsorption, thus yielding H2 due to water splitting. Figure 9a shows the mechanistic view of the synthesized ZnM-LDH photocatalysts and their selectivity to certain products. Figure 9b shows the photocatalytic product generated by the ZnM-LDH photocatalysts, and Figure 9c exhibits the respective selectivity.
In addition to the development of LDH photocatalysts, another important and beneficial approach is to use the LDH as a base material to convert it into a highly efficient photocatalyst. Recently, Wang et al. fabricated thin and defective NiO/Al2O3 composite from NiAl-LDH [56]. The NiAl-LDH photocatalyst was transformed to NiO with Ni and oxygen vacancies under oxidation and was investigated at various oxidation temperatures from 200–800 °C. The schematic view of the proposed mechanism for synthesis methodology is exhibited in Figure 10a. The photocatalytic CO2 conversion to various products with selectivity can be seen in Figure 10b,c, respectively. It could be observed, NiO sample synthesized at a temperature of 275 °C, yielded maximum CH4 under the irradiation of 600 nm wavelength light. Hence, this temperature meant to be an optimum condition for having the best oxygen and nickel vacancies.
The summary of the various bare/pure LDH photocatalysts regarding their research objective, reaction conditions, products yielded, and influential parameters are tabulated in Table 1.

4. Metals Loaded/Embedded Layered Double Hydroxide (LDH) Photocatalysts

Another effective strategy for the development of efficient LDH photocatalysts is to couple, embed, or dope the metals/non-metals to LDH structures, which will lead to improved photocatalytic performance in terms of CO2 photo-reduction. Until today, a limited amount of investigations has been done with respect to mentioned domain; however, there exists an enormous potential to develop metals embedded or doped LDH, which might result in an enhanced and efficient photocatalytic CO2 conversion to useful chemicals/fuels.
Zhao et al. reported a Ti embedded MgAl-LDH (MgAlTi-LDH) synthesized by three different methodologies and their performance evaluated by photocatalytic CO2 conversion with water to useful products [57]. The three different synthesis methodologies include (i) co-precipitation (CP), (ii) co-precipitation and hydrothermal (HT), and (iii) co-precipitation followed by calcination and reconstruction (R). It was observed that the two key factors influencing the photocatalytic activity were crystallinity and surface area of the prepared samples. Moreover, the effect of the temperature treatment was also investigated in accordance with the photocatalytic performance. Figure 11a shows the surface morphologies for the MgAlTi-LDH samples synthesized by coprecicpitation method followed by HT and R treatment at various temperatures, respectively. It was observed that the CP samples hydrothermally treated at different temperatures go under transformation of their small nanoflakes (50 nm) to large nanoflakes with increased size (500–1000 nm). Such transformation indicates the increase of crystallite size, which was also confirmed by XRD data. On the other hand, calcinated and reconstructed samples (R-MgAlTi LDH) do not exhibit any significant change in the crystallite size with the increase of the temperature. When employed for photocatalytic CO2 tests, all samples exhibited better CO yield as compared to standard TiO2-P25. Figure 11b,c show the photocatalytic CO2 conversion to CO normalized by catalyst mass and surface area, respectively. It was noticed that HT150 and HT200 samples (hydrothermally treated sample at 150 and 200 °C) exhibit best performance in accordance with both parameters, which can be attributed to the optimum condition of TiO2 crystallinity, crystallite size, band gap, and surface area of the LDH photocatalyst.
A research work reported by Iguchi et al. presented extension of their previously reported work by modifying the MgAl LDH with Ga2O3 and loading Ag nanoparticles on it [58]. It was observed that by loading 0.25 wt.% of Ag co-catalyst of Ga2O3-MgAl LDH, the photocatalytic CO2 conversion to CO was improved with enhanced selectivity towards CO. The molar concentration of already prepared MgAl LDH was varied in order to optimize the sample with the best performance. It was observed the sample with 95 molar concentration and 0.25 wt.% of loaded Ag co-catalyst, exhibited the best performance, yielding 211.7 µmol/h with 0.19 g of photocatalyst. Such an increased yield was associated to the enhanced CO2 adsorption by MgAl LDH, improving surface areas and selectivity induced by Ag nanoparticles.
Another research investigated by Chong et al. presents the role of MgAl-LDO as an interlayer in between Pt and TiO2 for CO2 photo-reduction to useful chemicals [59]. It was observed that the insertion of MgAl-LDO layer boosts the photocatalytic CO2 conversion to CO and CH4 as compared to reference samples. The authors used rutile TiO2 rods, as they are more active for water oxidation, thus providing more protons and electrons, which can be utilized for CO2 photo-reduction. The MgAl-LDH was in situ deposited and then calcined at high temperature to give LDO upon which the Pt (fixed 1 wt.%) was deposited. The sample optimization was performed by varying the loading amount of MgAl-LDH on the TiO2 surface. It was observed that there exist no significant changes in the sample crystallinity, surface morphology, and surface areas. Figure 12a–c shows TEM images for the Pt-MgAl-LDO-TiO2 photocatalyst with clear depiction of the MgAl-LDO interlayer and Pt deposition. Figure 12d shows the photocatalytic CO and CH4 evolution rate from CO2 photo-reduction and it can be noticed that Pt-MgAl-LDO-TiO2 exhibit the highest production rate as compared to all other samples. The key parameter for promoting and enhancing the photocatalytic CO2 conversion is synergetic effect of Pt deposition and interlayer of MgAl-LDO. It was proposed that the thin layer of MgAl-LDO optimized by MgAl-LDO deposition promotes the CO2 adsorption and efficient transfer of photogenerated electrons to Pt where the CO2 photo-reduction can occur. Moreover, the Pt nanoparticles act as electron sinks, and once they receive the photoexcited electrons, it is difficult for them to go back and recombine. The proposed reaction mechanism is well displayed in Figure 12e.
Another research conducted by Wang et al. developed series of palladium (Pd) loaded CoAl-LDH (Pd/CoAl-LDH) photocatalysts with varied content of Pd, in combination with a ruthenium-based complex acting as a photosensitizer [60]. When these photocatalysts employed for photocatalytic CO2 conversion, the syngas CO/H2 was the major product, where Pd promotes the production of H2 and CoAl-LDH yielded CO under visible light irradiation even when extended above 600 nm. It was observed that the ratio of CO/H2 can be tuned by varying the loading of Pd nanoparticles. Figure 13a,b show the production rate of CO/H2 and selectivity with varying the ratio of Pd. It can be noticed that with increasing the Pd amount, the relative amount of H2 is increased due to reduction of protons. As Pd is a well-known H2 producing cocatalyst, H2 is produced by Pd, which captures the photogenerated electrons from the LDH, whereas the photogenerated electrons within the LDH can directly react with the adsorbed CO2 to convert it into CO. The proposed mechanism of the CO2 photo-reduction over the prepared photocatalysts is shown in Figure 13c where triethanolamine (TEOA) assists to regenerate the Ru based photosensitizer.
The summarized overview of up to date developed metal embedded/loaded LDH photocatalysts employed for photocatalytic CO2 conversion to useful chemicals are presented in Table 2 in respect to research objective, reaction conditions, products yielded, and important influential parameters contributing to enhanced performance.

5. Composite/Hybrid/Heterojunctioned Layered Double Hydroxide (LDH) Based Photocatalysts

The fabrication of hybrid/heterojunctioned photocatalyst materials has always been an attractive and fascinating domain for the material scientists/researchers. As a well-explored domain, a similar approach is also implicated for the synthesis of hybrid/heterojunctioned photocatalysts employing LDH as one of the components with another established semiconductor material. The LDH support can act as an adsorbent and/or photocatalyst support to capture/convert CO2 into useful products. Hence, the benefits and gains of the hybrid/heterojunctioned materials can be a harvested for CO2 reduction by designing a suitable LDH based hybrid photocatalysts. With plenty of research space, until now, very limited investigations have been done with the respective concept.
A research work reported by Zhao et al. displays hybrid photocatalysts composed of MgAl-LDO coupled with TiO2 cuboids (MgAl-LDO/TiO2) [61]. The photocatalyst was synthesized by the combined approach of hydrothermal and coprecipitation methodologies, and a series of MgAl-LDO/TiO2 was prepared with varied molar ratio of Mg + Al to Ti, in order to optimize the photocatalyst with best performance. As the MgAl-LDO are reported to be in micrometer size platelets, it was required to synthesize micrometer size TiO2 which were synthesized in micrometer sized cuboids. Figure 14a–f shows the morphology of various MgAl-LDO/TiO2 photocatalysts with varied ratio of Mg + Al. It is obvious that MgAl-LDO platelets were successfully anchored on the TiO2 cuboids even after calcination and do not change the morphology. Hence, this report is also unique in terms of morphology of the photocatalyst obtained. When employed for CO2 photo-reduction with water vapors under UV light irradiation, CO was found to be the main product with small amount of CH4. Figure 14g shows the CO production rate by photocatalytic CO2 conversion. It was observed that at a condition of UV light irradiation (4 h) and 50 °C, MgAl-LDO alone had no activity, and also, when coupled with the TiO2, it did not show any significant photocatalytic activity towards CO2 conversion to CO. This might be due to weak CO2 adsorption at lower temperatures by MgAl-LDO. However, at higher temperature, i.e., 150 °C, the 10% Mg-Al LDO/TiO2 showed a 5 times higher CO production as compared to bare TiO2 cuboids. Moreover, 10% MgAl-LDO/TiO2 also exhibited the best performance, hence optimizing the Mg + Al ratio to Ti. Such an enhanced CO production can be attributed to the improved adsorption capacity of MgAl-LDO at higher temperatures. Upon illumination, the photogenerated electron in TiO2 can easily transfer to the nearby interface of MgAl-LDO and TiO2, where it can react with adsorbed CO2 converting it into CO. Figure 14h shows control tests of 10% MgAl-LDO/TiO2 under He gas and water vapors as reactants. It was observed that, at a temperature of 50 °C, 10% MgAl-LDO/TiO2 does not show any CO yield; however, at a temperature of 150 °C, a considerable amount of CO was observed. It was proposed that such CO might be a result of carbonate species on the surface of photocatalyst, which was stable at lower temperature, but at higher temperature, it was converted into CO.
Another work performed by Kumar et al. investigated the heterojunction of CoAl-LDH with P25, a well-known standard TiO2 [62]. P25 nanoparticles were encapsulated within microporous CoAl-LDH by a simple one step hydrothermal approach. It was observed that light absorption of the composite CoAl-LDH -P25 was red shifted, thus capturing the solar spectrum which might lead to improved photocatalytic activity. Furthermore, the CO2 adsorption capacity and matched band alignment also promote the efficient conversion of CO2 into useful chemicals. Figure 15a,b shows the morphology of the P25@CoAl-LDH photocatalyst with clear depiction of 2D platelets of CoAl LDH with thickness of approximately 20 nm and irregular nanoparticles of P25 within size range of 20–50 nm. Figure 15c shows the UV–Vis DRS for all the samples investigated in the research work. P25 showed its characteristic absorption in the UV range with the absorption peak centered around 300 nm, whereas CoAl-LDH exhibited two absorption peaks appearing in the UV and visible range, respectively. The absorption peak appearing around 500 nm was attributed to the d–d transitions of octahedral Co2+. Figure 15d shows the steady state photoluminescence (PL) spectroscopy for the samples. The PL peaks for P25@CoAl-LDH were intermediate between pure P25 and CoAl-LDH, thus endorsing the reduced charge recombination within the composite P25@CoAl-LDH photocatalyst. When employed for photocatalytic CO2 conversion, only three products were formed, i.e., CO, H2, and O2, shown in Figure 16a–d. Pure P25 showed the poorest activity with negligible production; however, CoAl-LDH exhibited certain amount of CO and O2 production. P25@CoAl-LDH showed a production in between P25 and CoAl-LDH, whereas 20 wt.% P25@CoAl-LDH showed a superior production of CO with negligible amount of CH4. Such an increased CO production and higher selectivity can be attributed to the well heterojunction formation within P25@CoAl-LDH and was associated to extended light absorption, enhanced CO2 adsorption, and effective photogenerated charges separation. Furthermore, the P25@CoAl-LDH sample also exhibited a stable performance up to 3 cycles of CO2 photo-reduction. Figure 15e shows the band gap alignment and heterojunction formation for the P25@CoAl-LDH photocatalyst. As well established, band gap of P25 is suitable for proton reduction but not suitable for CO2 photo-reduction. On the contrary, the bandgap of CoAl-LDH with its conduction band edge at −0.75 eV can easily produce CO. Thus, the combination of both P25 and CoAl-LDH leads to formation of p-n junction which is well aligned for the efficient production of CO and photogenerated charge separation, ultimately resulting in enhanced photocatalytic activity for CO2 conversion into useful product.
A research work executed by Tang et al. demonstrated the synthesis of unique homo-heterojunction of BiOCl nanoplates coupled with ZnCr-LDH by a simple electrostatic interaction method [63]. As well-known BiOCl generally consists of two directions, that is, {001} and {110} surfaces. Both surfaces form a homojunction within the BiOCl, whereas the ZnCr-LDH can assemble on {001} surface via electrostatic interaction. A built-in field will allow the electrons generated under irradiation to flow from one facet to another facet of BiOCl, from where they can easily flow down to the coupled ZnCr-LDH and react with adsorbed CO2. Various samples were synthesized with different content of the ZnCr-LDH represented by “Sc” wt.%. Figure 17a shows the photocatalytic CO2 conversion for various BiOCl-ZnCr-LDH samples with different contents of ZnCr yielding CH4 as a main product. The maximum yield was obtained when ZnCr-LDH amount was 10 wt.%. The proposed mechanism for charge transfer within BiOCl homojunction and heterojunction with ZnCr-LDH is shown in Figure 17b. The key factor indulged in the performance improvement was the efficient charge transfer.
Another research work executed by Yang and coauthors proposed an urchinlike Z-scheme photocatalyst by hierarchical structure of CoZnAl-LDH/RGO/g-C3N4 (LDH/RGO/CN) [64]. The LDH was synthesized by simple hydrothermal method, while for the synthesis of LDH/RGO/CN, the already prepared RGO and CN were added into the ionic precursors of LDH and then subjected for hydrothermal process. For the sake of optimization and achieving of best sample, the weight ratio of CN was varied among the LDH/RGO/CN hybrid photocatalyst. The schematic view of the synthesis procedure designed to achieve LDH/RGO/CN is shown in Figure 18a. The morphologies of CN and LDH/RGO/CN are shown in Figure 18b,c, respectively. The CN exhibited a bulky morphology with somewhat lamellar structures, whereas LDH/RGO/CN exhibited a well observable urchin like nanostructure. When employed for photocatalytic CO2 conversion, CO was observed as the key product, with CH4 up to 5 h of irradiation; it is shown in Figure 18d,e. It could be noticed that insertion of RGO between CoZnAl-LDH and CN significantly enhanced the CO yield, specifically for sample LDH/RGO/CN-2 (with 0.1 g of CN in the composite) whose CO yield was 3.4 and 8.5 times higher than LDH/CN and bare CN, respectively. As well established, CO2 adsorption and photocatalytic activity are two prime contributors for efficient photocatalytic CO2 conversion. The ternary composite of CoZnAl-LDH shows improved CO2 adsorption whereas CN is well known for its visible light activity. In addition, the insertion of RGO promotes the efficient charge separation from CoZnAl-LDH to valence band of CN, whereas the holes of CoZnAl-LDH are generated by water oxidation. The proposed schematic based on several analysis techniques is displayed in Figure 18f.
Another excellent work done by Ziarati et al. reported an efficient architecture of 3D yolk@shell TiO2−x/ CoAl LDH (Y@S TiO2−x/LDH) for photocatalytic CO2 conversion to solar fuels [65]. Such an architecture was synthesized by sequential designing of solvothermal, hydrogen treatment, and hydrothermal steps. When such a 3D architecture was employed for CO2 photo-reduction, it yielded CH3OH as a main product for the first 2 h of irradiation; however, the CH3OH production decreased with the passage of time, and CH4 yield increased. Figure 19a shows the CH3OH production for the 3D architecture photocatalyst along with other reference samples. Figure 19b exhibits the time dependent photocatalytic CO2 conversion to solar fuels. Such a conversion of product from CH3OH to CH4 is proposed, since, with the passage of time, CH3OH gets adsorbed on the surface, and photocatalytic is converted into CH3 radical, which reacts with electron and protons to form CH4 as the key product. The enhanced performance of the 3D photocatalyst was attributed to the combination of better CO2 sorption capacity, improved light absorption, and enhanced charge separation at interface of TiO2−x and LDH.
Another approach represented an LDH based composite photocatalyst with doped strategy. Jo et al. developed a N doped C dots/ CoAl LDH/gC3N4 (NCD/LDH/CN, represented as NLC) hybrid photocatalyst with efficient and selective photocatalytic CO2 conversion to CH4 [66]. The N doped carbon dots (NCD) are well researched materials as an alternate to noble metals cocatalysts. In their work, co-authors designed a strategy to employ NCD in replacement of noble metals and found a significant increase and selective production of CH4. Various NLC photocatalysts were synthesized by varying the contents of LDH to CN as 5, 10, 15, and 20 wt.%, whereas the amount of NCD was kept fixed for all samples, which was 2 wt.%. Figure 20a shows the SEM image of the NLC sample exhibiting a flower like morphology with clear depiction of LDH and CN. Figure 20b shows a high resolution TEM image of NLC with clear presence of all three components, i.e., LDH, CN, and NCD in the hybrid sample. When utilized for photocatalytic CO2 conversion, the key product obtained was CH4 with small amounts of CO and H2, as shown in Figure 20c. It was observed when NLC is formed by 2 wt.% NCD and with LDH to CN ratio of 10, the sample NLC-10 exhibited the maximum production rate as compared to all other samples. The most remarkable selectivity was 99% towards CH4 for NLC-10 sample. The reusability of the NLC-10 was also evaluated (shown in Figure 20d), which displayed a durable performance in each cycle, thus representing the stable performance during the prolonged reaction possibly to a good structural stability of the photocatalyst. The prime factors contributing to significantly improved photocatalytic performance of NLC consist of broadened light absorption, optimized surface area, and improved charge separation due to 2D junction in between LDH and CN. Moreover, the product selectivity was attributed to the well aligned band gap with corresponding product redox potential. The proposed mechanism for the photocatalytic CO2 conversion employing NLC-10 sample is shown in Figure 20e. Under the visible light irradiation, both LDH and CN photogenerate the charges and electron flowing down to conduction band of LDH whereas holes from LDH towards the valence band of CN where they are regenerated by water oxidation. Due to formation of well-defined 2D–2D heterojunction between LDH and CN, the photogenerated charges are efficiently transferred and react with CO2 adsorbed on the LDH surface. In addition, NCD also acts as electron sinks for efficient removal of electrons from conduction band of LDH to get reacted with adsorbed CO2 and converted to CH4.
Recently Wu and co-authors developed an efficient NiFe-LDH wrapped Cu2O nanocube (NFC) heterostructure for enhanced photocatalytic CO2 conversion [67]. The NFC photocatalysts were synthesized by a simple co-precipitation approach with varied time of aging temperature. Figure 21a–c shows the SEM images of NiFe-LDH, Cu2O nanocubes and the NFC heterostructure photocatalyst, respectively. The NiFe-LDH consisted of interconnected layers resulting in a flower like morphology, whereas Cu2O were in fine nanocubes. The NFC heterostructure shows a smooth structure of nanocubes covered with layers of LDH. The photocatalytic CO2 conversion for NFC photocatalyst yielded mainly CH4 as a key product. Figure 21d shows the CH4 production rate for various samples with varied aging time, i.e., 1, 2, and 4 h represented by NFC-1, NFC-2, and NFC-4, respectively. The NFC-4 displayed the highest efficiency amongst all the samples. The authors proposed a Z-scheme mechanism for conversion of CO2 to CH4, shown in Figure 21e. It can be assumed that upon light irradiation, the electrons generated in Cu2O react to the adsorbed CO2 on NFC surface and are converted to CH4, while the holes in valence band are filled by the photogenerated electrons from LDH. Thus, the key factors involved for the enhanced efficiency include visible light absorption, efficient charge separation, and improved CO2 adsorption.
Similarly, Jiang et al. reported Cu2O loaded ZnCr LDH for CO2 photo-reduction to useful fuels [68]. The Cu2O@ZnCr LDH was synthesized by a ternary CuZnCr LDH via in situ reduction process. A variety of samples were prepared by in situ reduction of Cu2−xZn2−2xCr LDH (where x = 0.05, 0.1, 0.2, and 0.4) resulting in the formation of xCu2O@Zn2−2xCr LDH. The photocatalytic CO2 conversion employing xCu2O@Zn2−2xCr LDH sample is shown in Figure 22a,b. The CO2 conversion to mainly CO was observed from the samples with 0.1Cu2O@Zn1.8Cr LDH sample exhibiting the highest yield as compared to its corresponding reference samples (Figure 22a). The key behind the significantly enhanced activity after Cu loading was its superb electron extraction property from the photocatalyst. Figure 22c exhibits the CO2 conversion of various xCu2O@Zn2−2xCr LDH samples, indicating the 0.1Cu2O@Zn1.8Cr LDH sample is the most efficient photocatalyst with optimized loading of Cu content. Moreover, the effect of various additives was also investigated for 0.1Cu2O@Zn1.8Cr LDH sample and is shown in Figure 22d. It was observed that Na2SO3 and Na2CO3 additives suppressed the formation of CO and enhanced the formation of H2, due to their better hole scavenging properties and occupying the CO2 adsorption sites. Figure 22e shows the proposed mechanism of CO2 photo-reduction for 0.1Cu2O@Zn1.8Cr LDH sample. Upon light irradiation, the photogenerated electrons are extracted by Cu2O nanoparticles where they react with adsorbed CO2 species to yield CO. Hence the key role of Cu2O nanoparticles on the surface of 0.1Cu2O@Zn1.8Cr LDH sample is an efficient electron separator providing CO2 active sites.
In another manner, LDH materials can be utilized as based material to synthesize efficient composite/heterojunctioned/hybrid photocatalysts. There are very limited reports regarding the utilization of LDH materials as photocatalyst precursors. To our best knowledge, in our investigation, we found a couple of research works taking advantage of mentioned approach and providing a novel pathway towards efficient photocatalysts development. Ye et al. reported CeO2−x platelets from monometallic cerium LDH (MCe-LDH) and evaluated their performance for photocatalytic CO2 conversion [69]. The synthesis approach consists of two steps including synthesis of MCe-LDH in the first step followed by second step of heat treatment at various temperatures up to 800 °C. Figure 23a shows the SEM image of a synthesized un-calcined MCe-LDH displaying the quasi-hexagonal platelets. Figure 23b shows the SEM image of the MCe-LDH calcined at a temperature of 800 °C, thus indicating that thermal treatment does not change significantly the structure except for the increased density of the nanopores. Figure 23c exhibits the UV–Vis DRS for various samples. It was observed that MCe-LDH exhibits broad hump at 320 nm, whereas when MCe-LDH were calcined, the absorption spectra showed two absorption peaks at 280 and 320 nm, corresponding to the Ce3+ and Ce4+, respectively. The inset to UV–Vis DRS shows the absorption ratio for Ce3+ to Ce4+ with temperature. The MCe-LDH sample calcined at 800 °C (MCe-LDH-800) shows the maximum ratio, whereas upon increasing the temperature up to 1000 °C, the absorption ratio decreased due to structure distortion. Figure 23d displays the photocatalytic CO2 conversion test with the synthesized and reference samples. It was observed that only CO was yielded, and sample calcined at 800 °C yielded the maximum CO production rate, 2.5 times higher than reference CeO2. Moreover, the optimized sample MCe-LDH-800 shows a stable photocatalytic performance for 4 cycles (8 h of each cycle) of testing. The photocatalytic performance was believed to be enhanced by two key factors of Ce3+/Ce4+ redox couple present at the surface of MCe-LDH and enhanced surface areas. The Ce3+/Ce4+ redox couple supports in the efficient charge separation, whereas improved surface areas lead to increased active sites for CO2 adsorption and conversion. The proposed mechanism by authors is shown in Figure 23e; under light irradiation, Ce4+ ions and oxygen vacancies will trap the photogenerated electrons. Such oxygen vacancy will promote the CO2 adsorption which will react with Ce3+ ions and get converted to CO2, which will further react with electron and proton to give CO. On the other hand, Ce3+ ions after reducing adsorbed CO2 will be converted to Ce4+ ions, and thus, the reaction cycle continues.
The summary of the reviewed composite photocatalysts with LDH with respective research aims, reaction condition, solar fuels production, and various contributing parameters to catalytic performance is tabulated in Table 3.

6. Conclusions

The review mainly focusses on the progress of LDH and LDH based photocatalysts for photocatalytic CO2 conversion. The influential properties, reaction conditions, proposed mechanisms, and various influential aspects for photocatalytic CO2 conversion were comprehensively explained. The interesting and mind captivating behavior of LDH photocatalysts for improved photocatalytic performance can be attributed to a combination of prime parameters which include (i) improved CO2 adsorption, (ii) improved light absorption, (iii) improved separation of photogenerated charges, and (iv) better product selectivity due to band gap alignment. Furthermore, surface modification of LDH, effect of various additives, noble metals loading/embedding, wrapping with graphene derivatives, and hybrid LDH based heterojunctioned photocatalysts were studied and explained principally. Conclusively, it can be inferred that LDH and/or LDH based photocatalysts open up a new way towards the alternate and efficient photocatalysts with a great potential for efficient for photocatalytic CO2 conversion into useful chemicals/fuels.

Author Contributions

A.R. collected the data, analyzed it, and wrote the review. S.A. reviewed and modified the manuscript. M.A. analyzed and tabulated the data, and S.-I.I. analyzed, modified, and rewrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are thankful towards the Ministry of Science and ICT and the Technology Development Program to Solve Climate Changes of the National Research Foundation (NRF) funded by the Ministry of Science and ICT for financial support under the grant numbers 2017R1E1A1A01074890 and 2015M1A2A2074670, respectively.

Acknowledgments

The authors gratefully acknowledge the support of the Ministry of Science and ICT, and the Technology Development Program to Solve Climate Changes of the National Research Foundation (NRF) funded by the Ministry of Science. The authors also acknowledge the Flux Photon Corporation for their support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cherniwchan, J. Economic growth, industrialization, and the environment. Resour. Energy Econ. 2012, 34, 442–467. [Google Scholar] [CrossRef]
  2. Aznar-Márquez, J.; Ruiz-Tamarit, J.R. Environmental pollution, sustained growth, and suf fi cient conditions for sustainable development. Econ. Model. 2016, 54, 439–449. [Google Scholar] [CrossRef]
  3. Li, K.; Lin, B. Impacts of urbanization and industrialization on energy consumption/CO2 emissions: Does the level of development matter? Renew. Sustain. Energy Rev. 2015, 52, 1107–1122. [Google Scholar] [CrossRef]
  4. Rabab, S.; Zaman, K.; Mushtaq, M.; Ahmad, M. Energy for economic growth, industrialization, environment and natural resources: Living with just enough. Renew. Sustain. Energy Rev. 2013, 25, 580–595. [Google Scholar]
  5. Yujie, S. As featured in: Steering charge kinetics in photocatalysis: Intersection of materials syntheses. Chem. Soc. Rev. 2015, 44, 2893–2939. [Google Scholar]
  6. White, J.L.; Baruch, M.F.; Iii, J.E.P.; Hu, Y.; Fortmeyer, I.C.; Park, J.E.; Zhang, T.; Liao, K.; Gu, J.; Yan, Y. Light-Driven Heterogeneous Reduction of Carbon Dioxide: Photocatalysts and Photoelectrodes. Chem. Rev. 2015, 115, 12888–12935. [Google Scholar] [CrossRef] [PubMed]
  7. Tu, W.; Zhou, Y.; Zou, Z. Photocatalytic Conversion of CO2 into Renewable Hydrocarbon Fuels: State-of-the-Art Accomplishment, Challenges and Prospects. Adv. Mater. 2014, 26, 4607–4626. [Google Scholar] [CrossRef] [PubMed]
  8. Inoue, T.; Fujishima, A.; Konishi, S.; Honda, K. Photoelectrocatalytic reduction of carbon dioxide in aqueous suspensions of semiconductor powders. Nature 1979, 277, 637–638. [Google Scholar] [CrossRef]
  9. Mo, S.; Ching, W. Electronic and optical properties of three phases of titaniumdioixde: Rutile, anatase, and brookite. Phys. Rev. 1995, 51, 23–32. [Google Scholar] [CrossRef]
  10. Mao, J.; Peng, T.; Zhang, X.; Li, K.; Ye, L.; Zan, L. Effect of graphitic carbon nitride microstuctures on the activity and selectivity of photocatalytic CO2 reduction under visible light. Catal. Sci. Technol. 2013, 5, 1253–1260. [Google Scholar] [CrossRef]
  11. Peng, F.; Wang, J.; Ge, G.; He, T.; Cao, L.; He, Y.; Ma, H.; Sun, S. Photochemical reduction of CO2 catalyzed by silicon nanocrystals produced by high energy ball milling. Mater. Lett. 2013, 92, 65–67. [Google Scholar] [CrossRef]
  12. Hsu, H.-C.; Shown, I.; Chang, Y.-C.; Du, H.-Y.; Lin, Y.-G.; Tseng, C.-A.; Wang, C.-H.; Chen, L.-C.; Lin, Y.-C.; Chen, K.-H. Graphene oxide as a promising photocatalyst for CO2 to methanol conversion. Nanosclae 2013, 5, 262–268. [Google Scholar] [CrossRef]
  13. Liu, L.; Zhao, H.; Andino, J.M.; Li, Y. Photocatalytic CO2 Reduction with H2O on TiO2 Nanocrystals: Comparison of Anatase, Rutile, and Brookite Polymorphs and Exploration of Surface Chemistry. ACS Catal. 2012, 8, 1817–1822. [Google Scholar] [CrossRef]
  14. Ali, S.; Flores, M.C.; Razzaq, A.; Sorcar, S.; Hiragond, C.B.; Kim, H.R.; Park, Y.H.; Hwang, Y.; Kim, H.S.; Kim, H.; et al. Gas Phase Photocatalytic CO2 Reduction, “A Brief Overview for Benchmarking”. Catalysts 2019, 9, 727. [Google Scholar] [CrossRef] [Green Version]
  15. Parayil, S.K.; Razzaq, A.; Park, S.; Kim, H.R.; Grimes, C.A.; In, S.I. Photocatalytic conversion of CO2 to hydrocarbon fuel using carbon and nitrogen co-doped sodium titanate nanotubes. Appl. Catal. A Gen. 2015, 498, 205–213. [Google Scholar] [CrossRef]
  16. Razzaq, A.; Sinhamahapatra, A.; Kang, T.; Grimes, C.A.; Yu, J.; In, S.-I. Efficient solar light photoreduction of CO2 to hydrocarbon fuels via magnesiothermally reduced TiO2 photocatalyst. Appl. Catal. B Environ. 2017, 215, 28–35. [Google Scholar] [CrossRef]
  17. In, S.-I.; Ii, D.D.V.; Schaak, R.E. Hybrid CuO-TiO2-xNx Hollow Nanocubes for Photocatalytic Conversion of CO 2 into Methane under Solar Irradiation. Angew. Chem. 2012, 3915–3918. [Google Scholar] [CrossRef]
  18. Park, S.-M.; Razzaq, A.; Park, Y.H.; Sorcar, S.; Park, Y.; Grimes, C.A.; In, S.-I. Hybrid CuxO-TiO2 Heterostructured Composites for Photocatalytic CO2 Reduction into Methane Using Solar Irradiation: Sunlight into Fuel. ACS Omega 2016, 1, 868–875. [Google Scholar] [CrossRef] [Green Version]
  19. Zubair, M.; Razzaq, A.; Grimes, C.A.; In, S.-I. Cu2ZnSnS4(CZTS)-ZnO: A noble metal-free hybrid Z-scheme photocatalyst for enhanced solar-spectrum photocatalytic conversion of CO2 to CH4. J. CO2 Util. 2017, 20, 301–311. [Google Scholar] [CrossRef]
  20. Parayil, S.K.; Razzaq, A.; In, S.-I. Formation of titania-silica mixed oxides in solvent mixtures and their influences for the photocatalytic CO2 conversion to hydrocarbon. J. Nanosci. Nanotechnol. 2015, 15, 7285–7292. [Google Scholar] [CrossRef]
  21. Kim, K.; Razzaq, A.; Sorcar, S.; Park, Y.; Grimes, A. RSC Advances conversion of CO2 into CH4 under solar irradiation. RSC Adv. 2016, 6, 38964–38971. [Google Scholar] [CrossRef]
  22. Razzaq, A.; Grimes, C.A.; In, S.-I. Facile fabrication of a noble metal-free photocatalyst: TiO2 nanotube arrays covered with reduced graphene oxide. Carbon 2016, 98, 537–544. [Google Scholar] [CrossRef]
  23. Rim, H.; Razzaq, A.; Grimes, C.A.; In, S.-I. Heterojunction p-n-p Cu2O/S-TiO2/CuO: Synthesis and application to photocatalytic conversion of CO2 to methane. J CO2 Util. 2017, 20, 91–96. [Google Scholar]
  24. Hiragond, C.; Ali, S.; Sorcar, S.; In, S.-I. Hierarchical Nanostructured Photocatalysts for CO2 Photoreduction. Catalysts 2019, 9, 370. [Google Scholar] [CrossRef] [Green Version]
  25. Ali, S.; Razzaq, A.; In, S.-I. Development of graphene based photocatalysts for CO2 reduction to C1 chemicals: A brief overview. Catalysts 2018, 335, 39–54. [Google Scholar] [CrossRef]
  26. Ali, S.; Lee, J.; Kim, H.; Hwang, Y.; Razzaq, A.; Jung, J.; Cho, C.; In, S.-I. Sustained, photocatalytic CO2 reduction to CH4 in a continuous flow reactor by earth-abundant materials: Reduced titania-Cu2O Z-scheme heterostructures. Appl. Catal. B Environ. 2020, 279, 119344. [Google Scholar] [CrossRef]
  27. Daud, M.; Hai, A.; Banat, F.; Wazir, M.B.; Habib, M.; Bharath, G.; Al-harthi, M.A. A review on the recent advances, challenges and future aspect of layered double hydroxides (LDH)—Containing hybrids as promising adsorbents for dyes removal. J. Mol. Liq. 2019, 288, 110989. [Google Scholar] [CrossRef]
  28. Mohapatra, L.; Parida, K. A review on the recent progress, challenges and perspective of layered double hydroxides as promising photocatalysts. J. Mater. Chem. A Mater. Energy Sustain. 2016, 4, 10744–10766. [Google Scholar] [CrossRef]
  29. Long, X.; Wang, Z.; Xiao, S.; An, Y.; Yang, S. Transition metal based layered double hydroxides tailored for energy conversion and storage. Biochem. Pharmacol. 2016, 19, 213–226. [Google Scholar] [CrossRef]
  30. Zhang, W.; Li, N.; Xie, Z.; Liu, Z.; Huang, Q. ScienceDirect Defective layered double hydroxide formed by H2O2 treatment act as highly efficient electrocatalytic for oxygen evolution reaction. Int. J. Hydrog. Energy 2019, 44, 21858–21864. [Google Scholar] [CrossRef]
  31. Hou, L.; Du, Q.; Su, L.; Di, S.; Ma, Z.; Chen, L.; Shao, G. Ni-Co layered double hydroxide with self-assembled urchin like morphology for asymmetric supercapacitors. Mater. Lett. 2019, 237, 262–265. [Google Scholar] [CrossRef]
  32. Prasad, C.; Tang, H.; Qin, Q.; Zul, S.; Shah, S.; Bahadur, I. An overview of semiconductors/layered double hydroxides composites: Properties, synthesis, photocatalytic and photoelectrochemical applications. J. Mol. Liq. 2019, 289, 111114. [Google Scholar] [CrossRef]
  33. Chatterjee, A.; Bharadiya, P.; Hansora, D. Layered double hydroxide based bionanocomposites. Appl. Clay Sci. 2019, 177, 19–36. [Google Scholar] [CrossRef]
  34. Cao, Y.; Li, G.; Li, X. Graphene/layered double hydroxide nanocomposite: Properties, synthesis and applications. Chem. Eng. J. 2016, 292, 207–223. [Google Scholar] [CrossRef]
  35. Feitknecht, W.; Gerber, M. Acknowledgement of the double hydroxide and base double salt III Manesium-aluminium double hydroxide. Helv. Chim. Acta 1942, 27, 131–137. [Google Scholar] [CrossRef]
  36. Allmann, R. The crystal structure of pyroaurite. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1968, 24, 972–977. [Google Scholar] [CrossRef]
  37. Liu, X.; Shao, Y.; Zhang, Y.; Meng, G.; Zhang, T.; Wang, F. Using high-temperature mechanochemistry treatment to modify iron oxide and improve the corrosion performance of epoxy coating. High-temperature ball milling treatment. Corros. Sci. 2015, 90, 451–462. [Google Scholar] [CrossRef]
  38. Zhang, Q.; Wang, J.; Yin, S.; Sato, T.; Saito, F. Synthesis of a visible-light active TiO2-xSx photocatalyst by means of mechanochemical doping. J. Am. Ceram. Soc. 2004, 87, 1161–1163. [Google Scholar] [CrossRef]
  39. Olfs, H.W.; Torres-Dorante, L.O.; Eckelt, R.; Kosslick, H. Comparison of different synthesis routes for Mg-Al layered double hydroxides (LDH): Characterization of the structural phases and anion exchange properties. Appl. Clay Sci. 2009, 43, 459–464. [Google Scholar] [CrossRef]
  40. Theiss, F.L.; Ayoko, G.A.; Frost, R.L. Synthesis of layered double hydroxides containing Mg 2+, Zn 2+, Ca 2+ and Al 3+ layer cations by co-precipitation methods—A review. Appl. Surf. Sci. 2016, 383, 200–213. [Google Scholar] [CrossRef]
  41. Wang, Q.; Ohare, D. Recent advances in the synthesis and application of layered double hydroxide (LDH) nanosheets. Chem. Rev. 2012, 112, 4124–4155. [Google Scholar] [CrossRef] [PubMed]
  42. Arai, Y.; Ogawa, M. Preparation of Co-Al layered double hydroxides by the hydrothermal urea method for controlled particle size. Appl. Clay Sci. 2009, 42, 601–604. [Google Scholar] [CrossRef]
  43. Han, S.H.; Zhang, C.G.; Hou, W.G.; Sun, D.J.; Wang, G.T. Study on the preparation and structure of positive sol composed of mixed metal hydroxide. Colloid Polym. Sci. 1996, 274, 860–865. [Google Scholar] [CrossRef]
  44. Yuan, X.; Wang, Y.; Zhou, C. One-pot hydrothermal synthesis of graphene/MgAl-LDH composite by urea hydrolysis. Nanomater. Energy 2014, 3, 30–38. [Google Scholar] [CrossRef]
  45. Ahmed, N.; Shibata, Y.; Taniguchi, T.; Izumi, Y. Photocatalytic conversion of carbon dioxide into methanol using zinc-copper-M(III) (M = aluminum, gallium) layered double hydroxides. J. Catal. 2011, 279, 123–135. [Google Scholar] [CrossRef]
  46. Ahmed, N.; Morikawa, M.; Izumi, Y. Photocatalytic conversion of carbon dioxide into methanol using optimized layered double hydroxide catalysts. Catal. Today 2012, 185, 263–269. [Google Scholar] [CrossRef]
  47. Iguchi, S.; Teramura, K.; Hosokawa, S.; Tanaka, T. Photocatalytic conversion of CO2 in an aqueous solution using various kinds of layered double hydroxides. Catal. Today 2015, 251, 140–144. [Google Scholar] [CrossRef] [Green Version]
  48. Iguchi, S.; Teramura, K.; Hosokawa, S.; Tanaka, T. Effect of the chloride ion as a hole scavenger on the photocatalytic conversion of CO2 in an aqueous solution over Ni-Al layered double hydroxides. Phys. Chem. Chem. Phys. 2015, 17, 17995–18003. [Google Scholar] [CrossRef] [Green Version]
  49. Iguchi, S.; Teramura, K.; Hosokawa, S.; Tanaka, T. Photocatalytic conversion of CO2 in water using fluorinated layered double hydroxides as photocatalysts. Appl. Catal. A Gen. 2016, 521, 160–167. [Google Scholar] [CrossRef] [Green Version]
  50. Flores-Flores, M.; Luévano-Hipólito, E.; Martínez, L.M.T.; Morales-Mendoza, G.; Gómez, R. Photocatalytic CO2 conversion by MgAl layered double hydroxides: Effect of Mg2+ precursor and microwave irradiation time. J. Photochem. Photobiol. A Chem. 2018, 363, 68–73. [Google Scholar] [CrossRef]
  51. Wang, K.; Zhang, L.; Su, Y.; Shao, D.; Zeng, S.; Wang, W. Photoreduction of carbon dioxide of atmospheric concentration to methane with water over CoAl-layered double hydroxide nanosheets. J. Mater. Chem. A 2018, 6, 8366–8373. [Google Scholar] [CrossRef]
  52. Tokudome, Y.; Fukui, M.; Iguchi, S.; Hasegawa, Y.; Teramura, K.; Tanaka, T.; Takemoto, M.; Katsura, R.; Takahashi, M. A nanoLDH catalyst with high CO2 adsorption capability for photo-catalytic reduction. J. Mater. Chem. A 2018, 6, 9684–9690. [Google Scholar] [CrossRef] [Green Version]
  53. Gao, G.; Zhu, Z.; Zheng, J.; Liu, Z.; Wang, Q.; Yan, Y. Ultrathin magnetic Mg-Al LDH photocatalyst for enhanced CO2 reduction: Fabrication and mechanism. J. Colloid Interface Sci. 2019, 555, 1–10. [Google Scholar] [CrossRef] [PubMed]
  54. Bai, S.; Wang, Z.; Tan, L.; Waterhouse, G.I.N.; Zhao, Y.; Song, Y.F. 600 nm Irradiation-Induced Efficient Photocatalytic CO2 Reduction by Ultrathin Layered Double Hydroxide Nanosheets. Ind. Eng. Chem. Res. 2020, 59, 5848–5857. [Google Scholar] [CrossRef]
  55. Xiong, X.; Zhao, Y.; Shi, R.; Yin, W.; Zhao, Y.; Waterhouse, G.I.N.; Zhang, T. Selective photocatalytic CO2 reduction over Zn-based layered double hydroxides containing tri or tetravalent metals. Sci. Bull. 2020, 65, 987–994. [Google Scholar] [CrossRef]
  56. Wang, Z.; Xu, S.M.; Tan, L.; Liu, G.; Shen, T.; Yu, C.; Wang, H.; Tao, Y.; Cao, X.; Zhao, Y.; et al. 600 nm-driven photoreduction of CO2 through the topological transformation of layered double hydroxides nanosheets. Appl. Catal. B Environ. 2020, 270, 118884. [Google Scholar] [CrossRef]
  57. Zhao, H.; Xu, J.; Liu, L.; Rao, G.; Zhao, C.; Li, Y. CO2 photoreduction with water vapor by Ti-embedded MgAl layered double hydroxides. J. CO2 Util. 2016, 15, 15–23. [Google Scholar] [CrossRef] [Green Version]
  58. Iguchi, S.; Hasegawa, Y.; Teramura, K.; Kidera, S.; Kikkawa, S.; Hosokawa, S.; Asakura, H.; Tanaka, T. Drastic improvement in the photocatalytic activity of Ga2O3 modified with Mg-Al layered double hydroxide for the conversion of CO2 in water. Sustain. Energy Fuels 2017, 1, 1740–1747. [Google Scholar] [CrossRef]
  59. Chong, R.; Su, C.; Du, Y.; Fan, Y.; Ling, Z.; Chang, Z.; Li, D. Insights into the role of MgAl layered double oxides interlayer in Pt/TiO2 toward photocatalytic CO2 reduction. J. Catal. 2018, 363, 92–101. [Google Scholar] [CrossRef]
  60. Wang, X.; Wang, Z.; Bai, Y.; Tan, L.; Xu, Y.; Hao, X.; Wang, J.; Mahadi, A.H.; Zhao, Y.; Zheng, L.; et al. Tuning the selectivity of photoreduction of CO2 to syngas over Pd/layered double hydroxide nanosheets under visible-light up to 600 nm. J. Energy Chem. 2020, 46, 1–7. [Google Scholar] [CrossRef]
  61. Zhao, C.; Liu, L.; Rao, G.; Zhao, H.; Wang, L.; Xu, J.; Li, Y. Synthesis of novel MgAl layered double oxide grafted TiO2 cuboids and their photocatalytic activity on CO2 reduction with water vapor. Catal. Sci. Technol. 2015, 5, 3288–3295. [Google Scholar] [CrossRef] [Green Version]
  62. Kumar, S.; Isaacs, M.A.; Trofimovaite, R.; Durndell, L.; Parlett, C.M.A.; Douthwaite, R.E.; Coulson, B.; Cockett, M.C.R.; Wilson, K.; Lee, A.F. P25@CoAl layered double hydroxide heterojunction nanocomposites for CO2 photocatalytic reduction. Appl. Catal. B Environ. 2017, 209, 394–404. [Google Scholar] [CrossRef]
  63. Tang, L.; Chen, R.; Meng, X.; Lv, B.; Fan, F.; Ye, J.; Wang, X.; Zhou, Y.; Li, C.; Zou, Z. Unique homo-heterojunction synergistic system consisting of stacked BiOCl nanoplate/Zn-Cr layered double hydroxide nanosheets promoting photocatalytic conversion of CO2 into solar fuels. Chem. Commun. 2018, 54, 5126–5129. [Google Scholar] [CrossRef] [PubMed]
  64. Yang, Y.; Wu, J.; Xiao, T.; Tang, Z.; Shen, J.; Li, H.; Zhou, Y.; Zou, Z. Urchin-like hierarchical CoZnAl-LDH/RGO/g-C3N4 hybrid as a Z-scheme photocatalyst for efficient and selective CO2 reduction. Appl. Catal. B Environ. 2019, 255, 117771. [Google Scholar] [CrossRef]
  65. Ziarati, A.; Badiei, A.; Grillo, R.; Burgi, T. 3D Yolk@Shell TiO2−x/LDH Architecture: Tailored Structure for Visible Light CO2 Conversion. ACS Appl. Mater. Interfaces 2019, 11, 5903–5910. [Google Scholar] [CrossRef]
  66. Jo, W.K.; Kumar, S.; Tonda, S. N-doped C dot/CoAl-layered double hydroxide/g-C3N4 hybrid composites for efficient and selective solar-driven conversion of CO2 into CH4. Compos. Part B Eng. 2019, 176, 107212. [Google Scholar] [CrossRef]
  67. Wu, Y.; Gong, Y.; Liu, J.; Chen, T.; Liu, Q.; Zhu, Y.; Niu, L.; Li, C.; Liu, X.; Sun, C.Q.; et al. Constructing NiFe-LDH wrapped Cu2O nanocube heterostructure photocatalysts for enhanced photocatalytic dye degradation and CO2 reduction via Z-scheme mechanism. J. Alloys Compd. 2020, 831, 154723. [Google Scholar] [CrossRef]
  68. Jiang, H.; Katsumata, K.I.; Hong, J.; Yamaguchi, A.; Nakata, K.; Terashima, C.; Matsushita, N.; Miyauchi, M.; Fujishima, A. Photocatalytic reduction of CO2 on Cu2O-loaded Zn-Cr layered double hydroxides. Appl. Catal. B Environ. 2018, 224, 783–790. [Google Scholar] [CrossRef]
  69. Ye, T.; Huang, W.; Zeng, L.; Li, M.; Shi, J. CeO2−x platelet from monometallic cerium layered double hydroxides and its photocatalytic reduction of CO2. Appl. Catal. B Environ. 2017, 210, 141–148. [Google Scholar] [CrossRef]
Figure 1. Structural illustration of layered double hydroxide (LDH).
Figure 1. Structural illustration of layered double hydroxide (LDH).
Catalysts 10 01185 g001
Figure 2. (a) Photocatalytic reactions proposed for the CO2 photo-reduction to CH3OH employing ZnCuGa LDH. (b) Band gap alignment of ZnCuAl LDH (with varied Zn, Cu and Al ratio) against the redox potential of CH3OH, CO and H2. (taken with permission from reference [45]. Copyright 2011, Elsevier).
Figure 2. (a) Photocatalytic reactions proposed for the CO2 photo-reduction to CH3OH employing ZnCuGa LDH. (b) Band gap alignment of ZnCuAl LDH (with varied Zn, Cu and Al ratio) against the redox potential of CH3OH, CO and H2. (taken with permission from reference [45]. Copyright 2011, Elsevier).
Catalysts 10 01185 g002
Figure 3. (a) Photocatalytic CO2 conversion to useful chemicals over various M-Al or M-In LDH, where M stands for Mg, Ni, Zn. (b) Photocatalytic CO evolution rate and its selectivity for Ni-Al LDH with various salt additives (Figures taken with permission from reference [47] copyright 2011, Elsevier and [48] copyright 2015, The royal society of chemistry).
Figure 3. (a) Photocatalytic CO2 conversion to useful chemicals over various M-Al or M-In LDH, where M stands for Mg, Ni, Zn. (b) Photocatalytic CO evolution rate and its selectivity for Ni-Al LDH with various salt additives (Figures taken with permission from reference [47] copyright 2011, Elsevier and [48] copyright 2015, The royal society of chemistry).
Catalysts 10 01185 g003
Figure 4. (a) Photocatalytic CO2 conversion to CH3OH for Mg-Al LDH photocatalyst; A represents acetate precursor, and N stands for nitrate precursor; 20 and 40 stands for microwave radiation time in minutes. (b) Band gap alignment of respective Mg-Al LDH against the redox potential of CH3OH, CO, CH4, and H2O. (Figures taken with permission from reference [50] Copyright 2018, Elsevier)).
Figure 4. (a) Photocatalytic CO2 conversion to CH3OH for Mg-Al LDH photocatalyst; A represents acetate precursor, and N stands for nitrate precursor; 20 and 40 stands for microwave radiation time in minutes. (b) Band gap alignment of respective Mg-Al LDH against the redox potential of CH3OH, CO, CH4, and H2O. (Figures taken with permission from reference [50] Copyright 2018, Elsevier)).
Catalysts 10 01185 g004
Figure 5. (a,b) TEM images of the Co-Al LDH Nanosheets displaying a hexagonal type morphology. (c) Photocatalytic CO2 conversion to CH4 for Co-Al LDH photocatalyst over more than 5 cycles of evaluation. (d) Mechanistic view of the photocatalytic process occurring at the surface of the Co-Al LDH. (Figures taken with permission from reference [51]. Copyright 2018, The Royal Society of Chemistry).
Figure 5. (a,b) TEM images of the Co-Al LDH Nanosheets displaying a hexagonal type morphology. (c) Photocatalytic CO2 conversion to CH4 for Co-Al LDH photocatalyst over more than 5 cycles of evaluation. (d) Mechanistic view of the photocatalytic process occurring at the surface of the Co-Al LDH. (Figures taken with permission from reference [51]. Copyright 2018, The Royal Society of Chemistry).
Catalysts 10 01185 g005
Figure 6. (a) Surface morphology of the nano Ni-Al LDH and bulk Ni-Al LDH with its surface area and particle size. (b) Photocatalytic CO2 conversion to CO and H2 for nano Ni-Al LDH (NLDH) and bulk Ni-Al LDH (reference LDH) with respective selectivity (Figures taken with permission from reference [52]. Copyrights 2018, The Royal Society of Chemistry).
Figure 6. (a) Surface morphology of the nano Ni-Al LDH and bulk Ni-Al LDH with its surface area and particle size. (b) Photocatalytic CO2 conversion to CO and H2 for nano Ni-Al LDH (NLDH) and bulk Ni-Al LDH (reference LDH) with respective selectivity (Figures taken with permission from reference [52]. Copyrights 2018, The Royal Society of Chemistry).
Catalysts 10 01185 g006
Figure 7. (a) Surface morphology of the Fe3O4/Mg-Al LDH. (b) UV–Vis DRS for Fe3O4/Mg-Al LDH, Mg-AL LDH and Fe3O4. (c,d) Photocatalytic CO2 conversion to CO and CH4, respectively, for various Fe3O4/Mg-Al LDH, represented by FMAL with varied amount of Fe3O4 (Figures taken with permission from reference [53]. Copyright 2019, Elsevier).
Figure 7. (a) Surface morphology of the Fe3O4/Mg-Al LDH. (b) UV–Vis DRS for Fe3O4/Mg-Al LDH, Mg-AL LDH and Fe3O4. (c,d) Photocatalytic CO2 conversion to CO and CH4, respectively, for various Fe3O4/Mg-Al LDH, represented by FMAL with varied amount of Fe3O4 (Figures taken with permission from reference [53]. Copyright 2019, Elsevier).
Catalysts 10 01185 g007
Figure 8. Photocatalytic CO2 conversion to CO, CH4, and H2 over (a) ultrathin CoAl-LDH, (b) under various conditions, and (c) with different ultrathin LDH photocatalysts. (Figures taken with permission from reference [54]. Copyrights 2020, American chemical society).
Figure 8. Photocatalytic CO2 conversion to CO, CH4, and H2 over (a) ultrathin CoAl-LDH, (b) under various conditions, and (c) with different ultrathin LDH photocatalysts. (Figures taken with permission from reference [54]. Copyrights 2020, American chemical society).
Catalysts 10 01185 g008
Figure 9. (a) Mechanistic view of various products obtained from different Zn based LDH (ZnM-LDH, where M = Ti, Fe, Co, Ga, Al) photocatalysts. (b) Photocatalytic CO2 conversion to useful chemicals. (c) Selectivity for various ZnM-LDH photocatalysts. (Figures taken with permission from reference [55]. Copyright 2020, Elsevier).
Figure 9. (a) Mechanistic view of various products obtained from different Zn based LDH (ZnM-LDH, where M = Ti, Fe, Co, Ga, Al) photocatalysts. (b) Photocatalytic CO2 conversion to useful chemicals. (c) Selectivity for various ZnM-LDH photocatalysts. (Figures taken with permission from reference [55]. Copyright 2020, Elsevier).
Catalysts 10 01185 g009
Figure 10. (a) Schematic view of the synthesis approach developed for obtaining NiO photocatalyst. (b) Photocatalytic CO2 conversion to useful chemicals. (c) Selectivity for various NiAl-LDH photocatalysts, oxidized at various temperatures as represented by temperature value after name. (Figures taken with permission from reference [56]. Copyright 2020, Elsevier).
Figure 10. (a) Schematic view of the synthesis approach developed for obtaining NiO photocatalyst. (b) Photocatalytic CO2 conversion to useful chemicals. (c) Selectivity for various NiAl-LDH photocatalysts, oxidized at various temperatures as represented by temperature value after name. (Figures taken with permission from reference [56]. Copyright 2020, Elsevier).
Catalysts 10 01185 g010
Figure 11. (a) SEM images for the MgAlTi-LDH samples synthesized by coprecipitation approach followed by hydrothermal treatment (HT100, HT150, and HT200) and coprecipitation followed by calcination and reconstruction (R400, R500, and R600); the number represents the temperature value. (b) Photocatalytic CO production rate normalized as amount of TiO2 present in the LDH photocatalysts and (c) normalized per surface area of the LDH photocatalysts synthesized. (Figures taken from [57]. Copyright 2016, Elsevier).
Figure 11. (a) SEM images for the MgAlTi-LDH samples synthesized by coprecipitation approach followed by hydrothermal treatment (HT100, HT150, and HT200) and coprecipitation followed by calcination and reconstruction (R400, R500, and R600); the number represents the temperature value. (b) Photocatalytic CO production rate normalized as amount of TiO2 present in the LDH photocatalysts and (c) normalized per surface area of the LDH photocatalysts synthesized. (Figures taken from [57]. Copyright 2016, Elsevier).
Catalysts 10 01185 g011
Figure 12. (ac) TEM images for the Pt-MgAl-LDO-TiO2 indicating the presence of interlayer MgAl-LDO between Pt nanoparticles and rutile TiO2 nanorods. (d) CO and CH4 evolution rate for all samples synthesized when employed for photocatalytic CO2 conversion. (e) Proposed mechanism involved in the photocatalytic CO2 conversion to CO and CH4 on the surface of Pt-MgAl-LDO-TiO2. (Figures taken from [59]. Copyright 2018, Elsevier).
Figure 12. (ac) TEM images for the Pt-MgAl-LDO-TiO2 indicating the presence of interlayer MgAl-LDO between Pt nanoparticles and rutile TiO2 nanorods. (d) CO and CH4 evolution rate for all samples synthesized when employed for photocatalytic CO2 conversion. (e) Proposed mechanism involved in the photocatalytic CO2 conversion to CO and CH4 on the surface of Pt-MgAl-LDO-TiO2. (Figures taken from [59]. Copyright 2018, Elsevier).
Catalysts 10 01185 g012
Figure 13. (a) Photocatalytic evolution syngas via CO2 photo-reduction for varied amount Pd loaded CoAl-LDH (Pd/CoAl-LDH) photocatalysts. (b) Selectivity of prepared LDH photocatalysts towards CO and H2 production with increased Pd content. (c) Proposed mechanism involved in the photocatalytic CO2 conversion to CO and H2 on the surface of Pd/CoAl-LDH photocatalyst. (Figures taken from [60]. Copyright 2019, Elsevier).
Figure 13. (a) Photocatalytic evolution syngas via CO2 photo-reduction for varied amount Pd loaded CoAl-LDH (Pd/CoAl-LDH) photocatalysts. (b) Selectivity of prepared LDH photocatalysts towards CO and H2 production with increased Pd content. (c) Proposed mechanism involved in the photocatalytic CO2 conversion to CO and H2 on the surface of Pd/CoAl-LDH photocatalyst. (Figures taken from [60]. Copyright 2019, Elsevier).
Catalysts 10 01185 g013
Figure 14. SEM images of MgAl-LDO/TiO2 composites: (a,b) 8% MgAl-LDO/TiO2, (c,d) 10% MgAl-LDO/TiO2, and (e,f) 12 wt.% MgAl-LDO/TiO2. Photocatalytic CO2 conversion into CO employing (g) all samples at varied temperature, and (h) 10 wt.% MgAl-LDO/TiO2under CO2 + H2O and He + H2O gaseous composition. (Figures taken with permission from reference [61]. Copyrights 2015, The Royal Society of Chemistry).
Figure 14. SEM images of MgAl-LDO/TiO2 composites: (a,b) 8% MgAl-LDO/TiO2, (c,d) 10% MgAl-LDO/TiO2, and (e,f) 12 wt.% MgAl-LDO/TiO2. Photocatalytic CO2 conversion into CO employing (g) all samples at varied temperature, and (h) 10 wt.% MgAl-LDO/TiO2under CO2 + H2O and He + H2O gaseous composition. (Figures taken with permission from reference [61]. Copyrights 2015, The Royal Society of Chemistry).
Catalysts 10 01185 g014
Figure 15. (a) High resolution SEM image and (b) TEM image of 20 wt.% P25@CoAl-LDH. (c) UV–Vis DRS and (d) PL spectra for all the samples employed in the research. (Figures taken with permission from reference [62]. Copyright 1969, Elsevier).
Figure 15. (a) High resolution SEM image and (b) TEM image of 20 wt.% P25@CoAl-LDH. (c) UV–Vis DRS and (d) PL spectra for all the samples employed in the research. (Figures taken with permission from reference [62]. Copyright 1969, Elsevier).
Catalysts 10 01185 g015
Figure 16. (a) Schematic diagram of the heterojunction formation between P25 and CoAl-LDH. Photocatalytic CO2 conversion to (b) CO, (c) H2, (d) O2, and (e) stability evaluation for 20 wt.% P25@CoAl-LDH. (Figures taken with permission from reference [62]. Copyright 1969, Elsevier).
Figure 16. (a) Schematic diagram of the heterojunction formation between P25 and CoAl-LDH. Photocatalytic CO2 conversion to (b) CO, (c) H2, (d) O2, and (e) stability evaluation for 20 wt.% P25@CoAl-LDH. (Figures taken with permission from reference [62]. Copyright 1969, Elsevier).
Catalysts 10 01185 g016
Figure 17. (a) Photocatalytic CO2 conversion to CH4 for various hybrid BiOCl-ZnCr-LDH photocatalysts, S represents the BiOCl-ZnCr-LDH sample with subscript indicating the composition of the ZnCr-LDH. (b) Schematic diagram of the heterojunction formation between BiOCl and ZnCr LDH. (Figures taken with permission from reference [63]. Copyrights 2018, The Royal Society of Chemistry).
Figure 17. (a) Photocatalytic CO2 conversion to CH4 for various hybrid BiOCl-ZnCr-LDH photocatalysts, S represents the BiOCl-ZnCr-LDH sample with subscript indicating the composition of the ZnCr-LDH. (b) Schematic diagram of the heterojunction formation between BiOCl and ZnCr LDH. (Figures taken with permission from reference [63]. Copyrights 2018, The Royal Society of Chemistry).
Catalysts 10 01185 g017
Figure 18. (a) Synthesis scheme of hybrid urchin like LDH/RGO/CN photocatalyst. SEM Images of (b) graphitic carbon nitride (CN), and (c) urchin-like CoZnAl-LDH. (d) Photocatalytic CO2 conversion to CO for all samples. (e) Production rate of CO and minor CH4 per hour from all samples. (f) Proposed mechanism for the heterojunctioned formation and photocatalytic CO2 conversion to CO, mainly. (Figures taken with permission from reference [64]. Copyright 2019, Elsevier).
Figure 18. (a) Synthesis scheme of hybrid urchin like LDH/RGO/CN photocatalyst. SEM Images of (b) graphitic carbon nitride (CN), and (c) urchin-like CoZnAl-LDH. (d) Photocatalytic CO2 conversion to CO for all samples. (e) Production rate of CO and minor CH4 per hour from all samples. (f) Proposed mechanism for the heterojunctioned formation and photocatalytic CO2 conversion to CO, mainly. (Figures taken with permission from reference [64]. Copyright 2019, Elsevier).
Catalysts 10 01185 g018
Figure 19. (a) Photocatalytic CO2 conversion into CH3OH (various samples). (b) Production rate of CH3OH and CH4 as a function of time over the hybrid LDH based photocatalyst. (Figures taken with permission from reference [65]. Copyright 2019 American Chemical Society).
Figure 19. (a) Photocatalytic CO2 conversion into CH3OH (various samples). (b) Production rate of CH3OH and CH4 as a function of time over the hybrid LDH based photocatalyst. (Figures taken with permission from reference [65]. Copyright 2019 American Chemical Society).
Catalysts 10 01185 g019
Figure 20. (a,b) TEM images for the hybrid NLC-10 photocatalyst. (c) Photocatalytic CO2 conversion from various samples to CO, H2, and CH4, with (d) stable performance evaluation. (e) Schematic for the proposed mechanism of CO2 conversion to useful chemicals over the hybrid NLC photocatalyst surface. (Figures taken with permission from reference [66]. Copyright 2019, Elsevier).
Figure 20. (a,b) TEM images for the hybrid NLC-10 photocatalyst. (c) Photocatalytic CO2 conversion from various samples to CO, H2, and CH4, with (d) stable performance evaluation. (e) Schematic for the proposed mechanism of CO2 conversion to useful chemicals over the hybrid NLC photocatalyst surface. (Figures taken with permission from reference [66]. Copyright 2019, Elsevier).
Catalysts 10 01185 g020
Figure 21. SEM images of (a) NiFe-LDH, (b) Cu2O Cuboids, and (c) hybrid NiFe-LDH-Cu2O (NFC-4 Sample). (d) Photocatalytic CO2 conversion from reference and various NFC samples to CH4. (e) Proposed Z-scheme mechanism involved in PCC with band gap alignment. (Figures taken with permission from reference [67]. Copyright 2020, Elsevier).
Figure 21. SEM images of (a) NiFe-LDH, (b) Cu2O Cuboids, and (c) hybrid NiFe-LDH-Cu2O (NFC-4 Sample). (d) Photocatalytic CO2 conversion from reference and various NFC samples to CH4. (e) Proposed Z-scheme mechanism involved in PCC with band gap alignment. (Figures taken with permission from reference [67]. Copyright 2020, Elsevier).
Catalysts 10 01185 g021
Figure 22. Photocatalytic CO2 conversion to CO and H2, (a) employing 0.1Cu2O@ZnCr LDH samples under CO2 and Ar gaseous reactants; (b) amount of CO and H2 yielded after 24 h of reaction for employing 0.1Cu2O@ZnCr LDH samples, (c) under CO2 gaseous reactant for various xCu2O@ZnCr LDH samples; and (d) amount of CO and H2 produced after 24 h for various xCu2O@ZnCr LDH samples. (e) Proposed mechanism involved in PCC with band gap alignment. (Figures taken with permission from reference [68]. Copyright 2017, Elsevier).
Figure 22. Photocatalytic CO2 conversion to CO and H2, (a) employing 0.1Cu2O@ZnCr LDH samples under CO2 and Ar gaseous reactants; (b) amount of CO and H2 yielded after 24 h of reaction for employing 0.1Cu2O@ZnCr LDH samples, (c) under CO2 gaseous reactant for various xCu2O@ZnCr LDH samples; and (d) amount of CO and H2 produced after 24 h for various xCu2O@ZnCr LDH samples. (e) Proposed mechanism involved in PCC with band gap alignment. (Figures taken with permission from reference [68]. Copyright 2017, Elsevier).
Catalysts 10 01185 g022
Figure 23. SEM images of (a) MCe-LDH and, (b) MCe-LDH calcined at 800 °C. (c) UV–Vis DRS spectra of the various reference and synthesized samples in the respective research work. (d) Photocatalytic CO2 conversion from various MCe-LDH samples to CO. (e) Schematic of the mechanism involved in PCC with H2O. (Figures taken with permission from reference [69]. Copyright 2017, Elsevier).
Figure 23. SEM images of (a) MCe-LDH and, (b) MCe-LDH calcined at 800 °C. (c) UV–Vis DRS spectra of the various reference and synthesized samples in the respective research work. (d) Photocatalytic CO2 conversion from various MCe-LDH samples to CO. (e) Schematic of the mechanism involved in PCC with H2O. (Figures taken with permission from reference [69]. Copyright 2017, Elsevier).
Catalysts 10 01185 g023
Table 1. Summary of various pure/bare LDH photocatalysts with research objective, reaction conditions, value added chemicals production by photocatalytic CO2 conversion, and key parameters for improved performance.
Table 1. Summary of various pure/bare LDH photocatalysts with research objective, reaction conditions, value added chemicals production by photocatalytic CO2 conversion, and key parameters for improved performance.
LDH PhotocatalystsResearch ObjectiveLight Source and Reactants EmployedPhotocatalytic Activity, CO2 Conversion to Value Added ChemicalsParameters Contributing towards Improved PerformanceRef
ZnCuGa LDH and
ZnCuAl LDH
  • Development of UV-Visible light active LDH with tailoring the product selectivity via Zn and Cu sites
  • UV–visible light obtained from 500 W Xenon arc lamp
  • CO2 with H2 gas were used as reactants
  • ZnCuGa LDH selectively produces CH3OH: 179 nmol g−1 h−1 (68 mol.%, selectivity) and CO: 79 nmol g−1 h−1
  • ZnCuAl LDH was selective to produce CO: 82 nmol g−1 h−1 with a selectivity of 94 mol.%
  • LDH prepared were UV–Vis light active material with improved CO2 adsorption
  • The inclusion of Cu sites improves the CH3OH selectivity
  • The key to CH3OH vs. CO selectivity was attributed to the binding nature of CO2 at Cu species in form of carbonates
[45]
ZnCuGa|Cu(OH)4 LDH
  • Development of ZnCuGa|Cu(OH)4 LDH by replacing the anions in ZnCuGa|CO3 and investigation of metal cations composition ratio of product selectivity and CO2 photoreduction efficiency
  • UV-visible light obtained from 500 W Xenon arc lamp
  • CO2 with H2 gas as a reductant
  • Photoreduction of CO2 to CH3OH for Zn3Ga|Cu(OH)4 LDH was obtained 0.49 µmol g−1 h−1 with 88 mol.% selectivity which is 5.9 times higher as compared to ZnCuGa|CO3
  • The stearic availability for the [Cu(OH)4]2− is attributed to play the role in enhanced CH3OH formation rate
  • The narrower bandgap value also contributes to the photocatalytic performance
  • The hydroxy groups bounded to Cu sites plays an important role in the CH3OH production
[46]
Various M-Al (M = Mg, Zn and Ni) LDH and M-Ni LDH (where M = Mg, Zn) prepared from aqueous solution of metal nitrates and chlorides
  • Investigation of the effect of different metal salts precursor
  • Analysis of metals cations ratio variation of photocatalytic activity
  • Study of CO2 photoreduction in aqueous solution of salts (NaCl, KCl, Na2CO3, NaHCO3)
  • UV–visible light
  • Obtained from 400 W Hg high pressure lamp
  • CO2 with H2O and aqueous solution of salts (NaCl, KCl, Na2CO3, NaHCO3) were used in a closed system
  • Maximum yield was obtained for Ni-Al LDH (Ni/Al = 4), synthesized by chloride metal salt employing 500 mg of photocatalyst in 350 mL of 0.1 M NaCl aqueous solution
  • CO: 36.6 µmol
  • CH4: 0.7 µmol
  • H2: 12.2 µmol
  • The Ni-Al LDH prepared by chloride metal salt precursor compensate the photogenerated charges well than LDH prepared from NO3 salts
  • CO2 photoreduction in NaCl solution exhibits maximum performance which was attributed to the presence of interlayer Cl ion in aqueous solution acting as a hole scavenger under UV light irradiation
[47]
Ni-Al LDH (Chloride ion, Cl effect)
  • Investigation regarding the Cl ion effect as a holes scavenger on CO2 photoreduction
  • UV–visible light
  • Obtained from 400 W high pressure Hg lamp
  • CO2 bubbled in aqueous solution of the NaCl containing suspended photocatalyst
  • Ni-Al LDH within 0.1 M NaCl solution, 0.5 g of suspended photocatalysts and 8 h of light irradiation produces:
  • CO: 56.4 µmol
  • H2: 9.3 µmol
  • with selectivity towards CO of 86 mol.%
  • After 29 h of irradiation CO: 110.9 µmolwith selectivity towards CO of 88.4 mol.%
  • The chloride ions (Cl) in the aqueous solution of NaCl improved the CO2 reduction via clear suppression of H2 production and acts as an efficient hole scavenger
[48]
Fluorinated
Ni-Al LDH
and
Fluorinated
Mg-Al LDH
  • Investigation of fluorination of LDH surfaces and its effect on CO2 photoreduction in aqueous solution.
  • UV–visible light obtained from 400 W Hg high pressure lamp
  • CO2 bubbled in water with suspended photocatalysts
  • Fluorinated Ni-Al LDH, with 7.5 ratio of Na2AlF6 in total Al species (Na2AlF6 + AlCl3) with 0.5 g suspended photocatalysts in 350 mL of 0.1 M NaCl solution
  • The maximum CO2 conversion rate producing CO: 93.3 µmolafter 20 h irradiation and selectivity of 80%
  • Fluorination incorporates the (AlF6)−3 units in LDH sheets, an important parameter leading to better CO2 adsorption and Cl ion in aqueous solution suppressing H2 formation and acting as a hole scavenger
[49]
Mg-Al LDH
  • Investigating the effect of various Mg+2 precursor (acetates and nitrates) and microwave irradiation time for conversion of CO2 by Mg-Al LDH
  • Solar simulator using 150 W Xe lamp
  • Reaction is carried out in both liquid and gas phase with CO2 and H2O as reactants
  • Mg-Al LDH synthesized using acetate precursor and 40 min of microwave irradiation yielded higher solar products in both liquid and gas phase
  • Liquid Phase maximum CH3OH: 9 µmol g−1 h−1
  • Gas Phase CO: 1.3 µmol g−1 h−1 CH4: 0.4 µmol g−1 h−1
  • An optimal cut off various parameters enhanced the photocatalytic performance which includes the following:
  • LDH Crystallinity
  • Flat band potential
  • CO2 adsorption
  • Charge transfer resistance
[50]
Co-Al LDH
  • The nanosheets of Co-Al LDH were synthesized and applied for atmospheric CO2 conversion to useful products CO and CH4
  • Simulated solar light obtained using 500 W Xe lamp
  • CO2 and H2O vapors were employed as reactants under gaseous closed system
  • Co-Al LDH nanosheets yielded CH4: 4.3 µmol g−1 h−1 after 5 h of irradiation and the yield reaches to 90% selectivity in 55 h of reaction
  • Better light adsorption as band gap is 2.1 eV
  • Surface alkaline OH groups for better CO2 adsorption
  • Divalent Co for efficient water splitting producing H2 for hydrogenation reactions
[51]
Nano crystals of Ni-Al LDH
  • Development of nanocrystals of Ni-Al LDH for enhanced CO2 photoreduction
  • Light is obtained from Hg high pressure Lamp of 400W
  • CO2 gas was bubbled through aqueous solution of 0.1 M NaCl with suspended photocatalyst (1.0 g)
  • Ni-Al LDH synthesized from nanocrystals (around 20 nm) produces CO: 50 µmol h−1, which is 7 times greater than standard crystalline Ni-Al LDH
  • The high surface affinity of the nano Ni-Al LDH to CO2 is attributed to the enhanced performance
  • Such specific surface property is related to the quenching of the metastable surface by rapid hydroxide formation form the respective molten salts
[52]
Ultrathin magnetic
Mg-Al LDH represented as Fe3O4/Mg-Al LDH
  • Synthesis of 2-D ultrathin magnetic Mg-Al LDH by coupling with Fe3O4 for enhanced CO2 photoreduction and recyclability of the photocatalyst
  • Ultraviolet light is obtained from 8 W lamp
  • CO2 gas was passed through solution of 0.1 M NaOH (70 mL) and acetonitrile (5 mL) with photocatalyst (0.05 g) suspended in it
  • Fe3O4/Mg-Al LDH with 10% of Fe3O4 content exhibited the maximum yield with production rate of, CO: 442.3 µmol g−1 h−1 CH4: 223.9 µmol g−1 h−1
  • Fe3O4 induces synergetic effect of photogenerated charges separation
  • The 2D ultra-thin Mg-Al LDH also reduces the transmission resistance of charge carriers resulting in increased reactions sites
[53]
u-MAl LDH, where M = Mg2+, Co2+, Ni2+, and Zn2+
  • Development of ultrathin LDH photocatalysts under 600 nm irradiation for high performance CO2 reduction
  • Light is obtained (400–800 nm) from Xenon Lamp of 300 W
  • CO2 was bubbled through the solution containing TEOA/H2O/CH3CN with dispersed photocatalysts and Ru based photosensitizer
  • u-CoAl LDH yielded the maximum products containing:
  • With 10 mg sample
  • CO: 2.52 mol g−1 h−1
  • With 0.05 mg sample
  • CO: 218.13 mmol g−1 h−1
  • With 0.05 mg sample under irradiation of 600 nm wavelength lightCO: 43.73 mmol g−1 h−1
  • Abundant availability of vacancies in terms of V o and V M enhances the photocatalytic performance due to enhanced charge transport and separation
[54]
Zn based LHD, represented by ZnM-LDH, where M = Ti, Fe, Co, Ga, and Al
  • To investigate the role of trivalent and tetravalent metal cations in the LDH towards CO2 photoreduction and product selectivity
  • UV light is obtained from Xenon lamp of 300 W with wavelength of 400–800 nm
  • The water and photocatalyst are spread on the base of the reaction chamber and CO2 gas in introduced to the reactor
  • Zn-Ti LDH produces CH4: 1.2µmol g−1 h−1 H2: 0.7 µmol g−1 h−1 with selectivity of 68% to CH4
  • Zn-Al LDH yielded CO: 1.3 µmol g−1 h−1 with selectivity of 90% to CO
  • Zn-Ga LDH produces CO: 1.6 µmol g−1 h−1 with selectivity of 78% to CO
  • Zn-Fe LDH yielded H2: 1.8 µmol g−1 h−1 with selectivity of 100% to H2
  • Zn-Co LDH yielded H2: 1.1 µmol g−1 h−1 with selectivity of 100% to H2
  • Trivalent and tetravalent metal cations improved alters the selectivity of the products
  • The metal ions with d-band centers relatively close to Fermi level strongly adsorb the CO2 and produced CH4 or CO dominantly
  • Whereas the metal ions with d-band centers far from Fermi level poorly adsorb the CO2 and yielded H2 as a major product
[55]
NiO Nano sheets synthesized form oxidation of NiAl-LDH
  • Synthesis approach for thins NiO sheets having Ni and O vacancies via oxidation of NiAl-LDH
  • Light irradiation was obtained from 300 W Xe lamp (300–800 nm)
  • The reaction occurred in liquid phase with CO2 bubbled through water/acetonitrile mixture, Ru based photosensitizer and TEOA as a sacrificial agent
  • NiO nanosheets produced at a temperature of 275 °C, represented by NiAl-275 showed the maximum product yield, i.e., CH4: 95 µmol g−1 h−1
  • At the optimum temperature, i.e., 275 °C, a special increase in the concentration’s special structure with optimized defects of Ni and oxygen is achieved, which tunes the selectivity and performance of the photocatalyst
[56]
Table 2. Summary of various metal loaded/embedded LDH photocatalysts with research goal, reaction conditions, value added chemicals production by photocatalytic CO2 conversion, and key parameters for improving performance.
Table 2. Summary of various metal loaded/embedded LDH photocatalysts with research goal, reaction conditions, value added chemicals production by photocatalytic CO2 conversion, and key parameters for improving performance.
LDH Photo Catalysts + CompositeResearch ObjectiveLight Source and Reactants EmployedPhoto Catalytic Activity, CO2 Conversion to Value Added Chemicals Parameters Contributing towards Improved Performance Ref.
Ti-embedded MgAl LDH represented as MgAlTi-LDH
  • Development of a ternary LDH
  • Exploration of three different synthesis approaches and investigation of LDH photocatalysts for CO2 photoreduction
  • UV–Vis light Obtained from 400 W Xe lamp providing light in the range of 200–1000 nm.
  • CO2 gas was bubbled through H2O towards the reactor for PCC
  • Only CO was detected as a main product for all samples with minor CH4
  • The sample which was synthesized by co-precipitation approach and followed by hydrothermal treatment at 150 °C showed the best CO yield of
  • CO: 50 µmol for 5 h of irradiation normalized per g of TiO2
  • Band gap, specific surface area, TiO2 crystallinity, and crystallite size; all these factors contributed to the improved photocatalytic activity
[57]
Pt-MgAL-LDO-TiO2
  • Development of an efficient LDO photocatalysts via introduction of MgAl LDO at interface of TiO2 and Pt nanoparticles
  • Light is obtained from 300 W Xenon lamp
  • CO2 with H2O vapor were used as reactants
  • The CO2 photoreduction to useful chemicals for all samples were as follows:
  • Bare TiO2 sample CO: 0.016 µmol g−1 h−1 CH4: 0.004 µmol g−1 h−1
  • Pt-MgAl-LDO/TiO2-n samples, where n = 1,2, 5 representing the volume of MgAl colloidal solution Pt-MgAl-LDO/TiO2-1 CO: 0.028 µmol g−1 h−1 CH4: 0.038 µmol g−1 h−1
  • Pt-MgAl-LDO/TiO2-2 CO: 0.038 µmol g−1 h−1 CH4: 0.046 µmol g−1 h−1
  • Pt-MgAl-LDO/TiO2-5 CO: 0.026 µmol g−1 h−1 CH4: 0.032 µmol g−1 h−1
  • Pt-MgAl-LDO/TiO-10 CO: 0.018 µmol g−1 h−1 CH4: 0.024 µmol g−1 h−1
  • Enhanced adsorption of CO2 on MgAl-LDH surface and thus easiness of its dissociation with H2O molecules
  • Efficient photogenerated charge separation by Pt leading to improved CO2 photoreduction
[59]
Pd/CoAl-LDH and Ru based photosensitizer
  • Development of an efficient heterostructure for photocatalytic CO2 conversion to syngas under light irradiation
  • CO2 reduction carried out under visible light irradiation with wavelength more than 400 nm and in gaseous phase
  • Pd/CoAl-LDH and Ru based photosensitizer increase the tunable syngas (CO/H2) molar ratio from 1:0.74 to 1:3 under visible light irradiation
  • The presence of Pd NPs lead to well adsorb the water molecules which lead to formation of H2 during light irradiation thus providing a pathway to tune the CO to H2 ratio
[60]
Table 3. Summary of various LDH based composite photocatalysts with research goal, reaction conditions, value added chemicals production by photocatalytic CO2 conversion, and key parameters for improved performance.
Table 3. Summary of various LDH based composite photocatalysts with research goal, reaction conditions, value added chemicals production by photocatalytic CO2 conversion, and key parameters for improved performance.
LDH
Photo-Catalyst
Research ObjectiveLight Source and Reactants EmployedPhoto-Catalytic Activity, CO2 Conversion to Value Added ChemicalsParameters Contributing towards Improved PerformanceRef
MgAl-LDO/ TiO2
  • Investigation of MgAl-LDO/ TiO2 composite photocatalysts for PCC by varying the ration of Mg + Al to Ti
  • 50 W Xe lamp with a UV cut-off filter was used as a light irradiation
  • CO2 bubbled through water was used as a reactant
  • Sample with 10 wt.% of MgAl-LDO TiO2 yields the best performance with production rate of CO: 4.3 µmol g−1 h−1 (under UV light) CO: 1.0 µmol g−1 h−1 (under Visible light)
  • Ratio of Mg + Al to Ti in the composite LDH enhances the CO2 adsorption
  • Under light irradiation, the photogenerated charges form TiO2 can travel to adsorbed CO2 and photoreduce to respective product
[61]
P25@CoAl- LDH Nano Composite
  • A facile and cost-effective synthesis strategy for P25 encapsulated CoAl-LDH with improved CO2 photoreduction
  • UV–Vis and visible light is obtained from 300 W Xe lamp
  • CO2 gas was bubbled through water containing suspended photocatalyst
  • 20 wt.% P25@CoAl- LDH yielded the best production rate of CO: 2.21 µmol g−1 h−1 (UV–Vis light irradiation) CO: 0.714 µmol g−1 h−1 (with visible light irradiation)
  • The type II junctions formed between P25 and CoAl LDH lead to the efficient separation of photogenerated electro-hole pairs leading to enhanced CO yield
  • The bandgap positions well aligned to produce selectively CO with a value of greater than 90%
[62]
BiOCl-ZnCr- LDH
  • The development of homo–hetero junctioned BiOCl-ZnCr-LDH for efficient CO2 photo-reduction
  • Light is obtained from 500 W Xenon Lamp
  • CO2 bubbled through water was used as a reactant
  • Stacking 10% wt.% of ZnCr-LDH with BiOCl resulted in the best photocatalytic performance with yielding CH4: 0.86 µmol g−1 (for 6 h irradiation)
  • The composite formation of ZnCr-LDH with BiOCl, resulting in homo–hetero junctioned system which lead to improved holes separation and thus enhanced photocatalytic performance
[63]
In situ loaded RGO and g-C3N4 on CoZnAl-LDH, represented by LDH/RGO/CN
  • Synthesis of a Z-scheme hybrid Photocatalysts
  • Investigation of various components for improved CO2 photoreduction
  • Light is obtained from Xenon lamp of 300 W
  • CO2 gas and water vapors were employed as key reactants
  • Bare CoZnAl-LDH produces CO: 5.99 µmol g−1 h−1
  • Bare g-C3N produces CO: 10.69 µmol g−1 h−1
  • LDH/CN produces CO: 15.06 µmol g−1 h−1
  • LDH/RGO/CN-1 produces CO: 19.39 µmol g−1 h−1
  • LDH/RGO/CN-2 produces CO: 50.53 µmol g−1 h−1
  • LDH/RGO/CN-3 produces CO: 15 µmol g−1 h−1
  • The optimum content of g-C3N4 provides the best performance due to increased charge lifetime and enhanced separation via Z-scheme mechanism
  • The RGO acts as an electron mediator at the junction of CN and LDH
[64]
Y@S TiO2−x/LDH
  • Development of a novel 3-D nanoarchitecture for efficient CO2 photoreduction
  • Investigation of time dependent experiments for selective fuel/chemical production
  • Light is obtained from Xenon lamp of 300 W
  • CO2 was bubbled through water and used as key reactants
  • Y@S TiO2−x/LDH yielded CH3OH/CH4 with respect to time of irradiation
  • For 2 h of irradiation, the key product obtained was CH3OH: 216 µmol g−1 h−1
  • After 12 h of irradiation, the amounts of CH3OH and CH4 increased to 726 and 453 μmol g−1, respectively
  • Advanced architecture of Y@S TiO2 containing oxygenvacancies linked with LDH plates enhances CO2 absorption in defect sites and interlayerspaces
  • Moreover, TiO2-x provides a better light absorption with narrower band gap
  • The time dependent selectivity was attributed to the conversion of radicals adsorbed to CH4 and CH3OH
[65]
NiFe-LDH wrapped Cu2O nanocubes,
(NFC)
  • Inexpensive and facile synthesis of Z-scheme 2-D NiFe-LDH and Cu2O nanocubes for efficient CO2 photoreduction
  • Light is obtained from Xenon Bulb intensity of 300 W
  • CO2 gas was generated in situ by reaction of NaHCO3 and H2SO4
  • NFC-4 with 4 hr of aging at 65 °C yielded the best solar fuel, i.e., CH4 around 3.33 mmol g−1 for 4 h irradiation, which is about 6.9 times of Cu2O and 5.6 times of NiFe-LDH
  • Z-scheme mechanism for charge separation and improved light absorption were the key reason for enhanced CH4 yield
[67]
xCu2O@Zn2−2xCr LDH
  • Inexpensive facile strategy to synthesize the ternary LDH photocatalyst for efficient CO2 photoreduction
  • UV light is obtained from 200 W Xe lamp
  • CO2 was bubbled through DI water containing respective photocatalysts in suspended form
  • Cu2O@Zn2−2xCr LDH produces maximum yield of solar products (for 24 h), which contains, H2: 41.2 µmol and CO: 0.8 µmol with additive of Na2SO4
  • The loaded Cu2O nanoparticles acts as an efficient electron extractor from LDH photocatalyst
  • The addition of Na2SO3 suppresses the generation of CO and promotes the water splitting leading to formation of H2.
[68]
Mono-metallic Cerium LDH
(MCe-LDH)
  • Synthesis of MCe-LDH by a simple and facile approach
  • Investigation of the heat treatment effect at various temperatures for CO2 photoreduction
  • UV–Vis light obtained from 300 W Xenon lamp
  • CO2 and water vapors were used as reactant in gaseous phase
  • MCe-LDH heated upto 800 °C yielded the maximum CO yield of 13.4 µmol g−1 in 8 h
  • Heat treatment resulted in enhancement of concentration ratio of and increased surface area, both factors enhanced the photo reduction of CO2
[69]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Razzaq, A.; Ali, S.; Asif, M.; In, S.-I. Layered Double Hydroxide (LDH) Based Photocatalysts: An Outstanding Strategy for Efficient Photocatalytic CO2 Conversion. Catalysts 2020, 10, 1185. https://doi.org/10.3390/catal10101185

AMA Style

Razzaq A, Ali S, Asif M, In S-I. Layered Double Hydroxide (LDH) Based Photocatalysts: An Outstanding Strategy for Efficient Photocatalytic CO2 Conversion. Catalysts. 2020; 10(10):1185. https://doi.org/10.3390/catal10101185

Chicago/Turabian Style

Razzaq, Abdul, Shahzad Ali, Muhammad Asif, and Su-Il In. 2020. "Layered Double Hydroxide (LDH) Based Photocatalysts: An Outstanding Strategy for Efficient Photocatalytic CO2 Conversion" Catalysts 10, no. 10: 1185. https://doi.org/10.3390/catal10101185

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop