Next Article in Journal / Special Issue
Solid State Nanostructured Metal Oxides as Photocatalysts and Their Application in Pollutant Degradation: A Review
Previous Article in Journal
Vibrational-Excitation-Induced and Spontaneous Conformational Changes in Solid Para-H2—Diminished Matrix Effects
Previous Article in Special Issue
Optical and Structural Characteristics of Rare Earth-Doped ZnO Nanocrystals Prepared in Colloidal Solution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress in the Photoreforming of Carboxylic Acids for Hydrogen Production

by
Anita Samage
1,
Pooja Gupta
2,
Mahaveer A. Halakarni
1,
Sanna Kotrappanavar Nataraj
1,* and
Apurba Sinhamahapatra
2,*
1
Centre for Nano and Material Sciences, Jain Global Campus, Jain University, Bengaluru 562112, Karnataka, India
2
Department of Chemical Engineering, Indian Institute of Technology (Indian School of Mines), Dhanbad 826004, Jharkhand, India
*
Authors to whom correspondence should be addressed.
Photochem 2022, 2(3), 580-608; https://doi.org/10.3390/photochem2030040
Submission received: 22 June 2022 / Revised: 17 July 2022 / Accepted: 22 July 2022 / Published: 29 July 2022

Abstract

:
Photoreforming is a process that connects the redox capability of photocatalysts upon light illumination to simultaneously drive the reduction of protons into hydrogen and the oxidation of organic substrates. Over the past few decades, researchers have devoted substantial efforts to enhancing the photocatalytic activity of the catalyst in hydrogen production. Currently, the realization of the potential of photocatalysts for simultaneous hydrogen production with value-added organics has motivated the research field to use the photo-oxidation path. As a distinct benefit, the less energetically demanding organic reforming is highly favorable compared to the slow kinetics of oxygen evolution, negating the need for expensive and/or harmful hole scavengers. Photocatalyst modifications, such as secondary component deposition, doping, defect, phase and morphology engineering, have been the main strategies adopted to tune the photo-oxidation pathways and oxidation products. The effect of the reaction parameters, including temperature, pH, reactant concentration and promising reactor strategies, can further enhance selectivity toward desired outcomes. This review provides a critical overview of photocatalysts in hydrogen production, including chemical reactions occurring with semiconductors and co-catalysts. The use of various oxygenates as sacrificial agents for hydrogen production is outlined in view of the transition of fossil fuels to clean energy. This review mainly focuses on recent development in the photoreforming of carboxylic acids, produced from the primary source, lignocellulose, through pyrolysis. The photo-oxidation of different carboxylic acids, e.g., formic acid, acetic acid and lactic acid, over different photocatalysts for hydrogen production is reviewed.

1. Introduction

One of the major technological challenges in today’s world is obtaining a sustainable clean energy source. Over the past few decades, the world’s energy consumption has increased exponentially due to overpopulation, overconsumption and aged energy storage systems, leading to the high consumption of non-renewable fossil fuels [1]. The more extensive use of fossil fuels results in elevated concentrations of atmospheric greenhouse gases; for example, carbon dioxide affects the environment of the total globe, which causes deviation and threatens biodiversity [2]. To overcome these major problems of the present earth, breakthrough solutions and innovative energy production technologies depending on renewable resources are required to compensate for carbon debt. As the world is traversing toward carbonless and renewable energy sources, hydrogen (H2) is becoming an increasingly powerful medium as a versatile energy carrier, enabling a globe with a zero-emission economy, and it has higher energy density than the conventional fossil-fuel-based sources [3]. H2 is a lightweight energy carrier, which can be easily converted into electricity and drinkable water as a byproduct via fuel cell technology. Therefore, H2 can prevent emissions from the transport and energy sectors, making the system emission-free. Moreover, it has been used in the production of chemicals such as ammonia (through Haber’s process), urea and in the manufacturing of steel. Here, in these processes, used H2 was derived from the gasification of coal [4], water gas shift reactions [5] and methane reforming [6]. These traditional hydrogen production methods have carbon dioxide footprints and are not renewable, environment-friendly and sustainable procedures. Most of the methods mentioned above require high temperatures, pressure conditions and the use of an external bias or non-renewable fossil fuels. Therefore, there is a need to produce hydrogen from renewable, sustainable and environmentally benign sources of green origin, which should be cost-effective [7]. Therefore, industries and transport sections can use hydrogen fuel efficiently and meet our energy demands for energy storage resolutions. As a result, moving from non-renewable fossil fuels to the hydrogen economy requires effective methodologies to allow clean, sustainable and cost-effective production and H2 storage systems. Various types of technologies have been projected so far, such as chemical, thermal, electrochemical, photo electrolytic and photocatalytic processes. These methods make use of water, biomass waste and fossil fuels as feed stocks for the generation of hydrogen [8]. In recent trends, solar-light-assisted methods utilizing solar radiations and water/renewable precursors have advantages and have gained noteworthy attraction over the past few decades. Recent investigation has predicted that H2 production using solar-light-supported methods are more advantageous than the traditional technologies, which are based non-renewable energy sources. The photocatalytic water-splitting reactions takes place via two half reactions (i.e., proton reduction and the transferring of four electrons for the oxidation of water) [9].
2H+ + 2e → H2 (hydrogen evolution)
2H2O + 4H+ → O2 + 4H+ (oxygen evolution)
2H2O → 2O2 + 2H2 (overall water splitting)
Subsequently, in 1972 Honda and Fusishima described the effective use of visible light radiation in an electrochemical system to decompose water into O2 and H2 using TiO2 as a photocatalyst [10]. Various types of catalysts are used in the water’s electrocatalytic and photocatalytic H2 production. The most demanding and exciting process is the photocatalytic production of H2 from water at ambient temperature and pressure, which converts and stores solar energy in the form of chemical energy [11]. Even though photocatalytic H2 production has neutral carbon footprints and renewable energy sources (secondary light sources or direct sunlight), it has encountered some challenges. Initially, it lacked advanced photocatalysts, which have more thermodynamic barriers for overall water splitting, which results in higher energy requirements with respect to intense light sources and requires a suitable sacrificial agent [12,13]. Various methods have been proposed to enhance the photocatalytic efficiency of catalysts in the form of structural modification, tuned morphology, the addition of cocatalysts, etc., which are mainly implemented for better charge generation and transfer efficiency. However, practical efforts are still required for the further enhancement of quantum efficiency and high yield production [14]. In addition, geographic segregation and the scarcity of clean water for conventional water splitting has led to challenging interests in using portable water sources. Therefore, to overcome these aforementioned challenges, photocatalytic H2 production with simultaneous organic reforming (photoreforming) is an efficient alternative method. The photoreforming reaction occurs via the oxidation of organic molecules [15]. This organic support acts as a proton source. Therefore, it is a coupling reaction of substrate oxidation and proton reduction. The primary reaction is as shown in the following equation: [16,17].
CxHyOz + (2x − z) H2O → (2x − z + y) H2 + xCO2
The advantages of the photoreforming process over traditional photocatalytic water splitting include: the backwards reaction of O2 and H2 during photocatalytic water splitting can be prevented; organic precursors or biomass are employed, which is a less energy-consuming process to replace kinetically sluggish oxygen evolution reactions by allowing higher efficiency in H2 evolution; and waste/biomass can be converted into valuable products or organic compounds. The use of organics in photocatalytic H2 production gives an alternative catalytic path; heat-absorbing organic molecules replace the vast energy required for water oxidation with organic molecule oxidation, which has lesser Gibbs free energy and leads to a lower thermodynamic energy barrier [18]. In this process, chemical feedstocks can be used from industries (food and beverages), agricultural waste and organics to simplify the reaction. Researchers have mainly focused on the photocatalytic oxidation of organics and biomass degradation during the last few decades, although various mechanistic depth investigations have been undertaken [19]. Most of the works concentrate exclusively on enhancing H2 production efficiency or increasing the selectivity in the oxidation process. The coupling of H2 production and selective organic oxidation into valuable products remains incoherent. Most of the studies have demonstrated the mechanism of the formation of H2 with increased efficiency without considering the oxidation mechanism. Understanding the oxidation and reduction mechanism in the whole photoreforming process is vital to realizing the potential of selective photoreforming in producing the H2 and individual organic molecules. In this context, the implementation of photoreforming process on an industrial scale and the selectivity of organic oxidation play an essential role in the formation of specific valuable organic products. For example, the oxidation of glycerol can produce glyceraldehyde, glyceric acid and glycolic acid [20] while preventing the generation of waste products results in photoreforming, which is the only process producing H2 with waste management [21]. The reaction selectivity and efficiency can be obtained by rationalizing catalyst material and reaction [22] media. This method allows a great opportunity to synthesize a particular organic molecule with simultaneous reductions in dependence on non-renewable fuels.
The present review describes the importance of photocatalysis in hydrogen formation with the support of literature reports, including review articles. The photoreforming of carboxylic acids into hydrogen with byproducts is expressed systematically, and the effect of catalytic materials, cocatalysts and sacrificial agents is reviewed thoroughly. A brief introduction to elementary steps in the photoreforming of carboxylic acids has been given. The impact of using organic acids on the production of H2 is discussed critically. Considerations of recent developments in novel synthetic strategies and activity toward the photoreforming of acids are discussed with organic/inorganic photocatalysts.

2. Details of the Photoreforming Process

The photocatalytic reactions occur through a chain of events in the presence of semiconducting materials. The photocatalytic reactions in different conditions are summarized in Figure 1 for water splitting and photoreforming reactions. In a photoreforming reaction, the initial crucial step is light absorption and the formation of charge carriers, such as electrons and hole pairs (e/h+). This charge carrier pair formation should occur upon irradiating the solar radiation on semiconducting materials using visible region energy, and to cause the reaction, the catalytic materials should be exposed to higher-energy radiation than its band gap energy. The electrons (e) from the valence band (VB) are excited to the conduction band (CB) by leaving holes (h+) behind [23]. The photogenerated charge carriers follow various pathways, such as the recombination of charge carriers at bulk/surface of catalyst leads to decrease in photocatalytic activity. Once these charge carriers are separated and transferred to the catalyst surface, they can reduce and oxidize the reactants. The formation and realization of a specific reaction are related to the band structures of semiconducting materials. The capacity of e and h+ towards reduction and oxidation reactions can be defined using the edge potentials of CB and VB [15].
In the photocatalytic production of H2, the bottom of the VB should be more negative than the H+/H2 redox couple (0 V vs. NHE) at a neutral pH. In a water-splitting reaction for the oxidation of water, the top of the VB should be more positive than the O2/H2O oxidation potential of 1.23 vs. NHE at a neutral pH [24]. Therefore, theoretically, the minimal value for the oxidation of water is 1.23 V vs. NHE at a neutral pH. In practice, this value increases to 2–2.4 eV because of the kinetic over-potential and energy loss during a reaction. With respect to thermodynamical parameters, water splitting requires higher positive Gibbs free energy (ΔG0 = +237.2 kJ/mol) [25]. This oxidation reaction is an endothermic half-reaction; however, the overall reaction provides H+ and e for the reduction reaction to produce H2. The mechanism involved in photoreforming, photocatalytic water splitting and photo-organic oxidation is displayed in Figure 1. In recent years, the demand for oxygen evolution reaction has been a bottleneck for the water-splitting reaction. Subsequently, there are narrow band gap (Eg) materials, which can catalyze both the oxidation and reduction reactions. In particular, working at a neutral pH is more challenging with respect to thermodynamics due to the lower availability of protons. Usually, in order to assess the capacity of photocatalysts, sacrificial agents are used to control the productivity towards one of the two half-reactions. However, the catalytic activity of the material towards one of the two half-reactions of water splitting is not necessary to be active for overall water splitting. In the photoreforming process, the oxidation potential is 0.08 V vs. NHE using organic substrates for scavenging h+ more productively, which consumes h+ rapidly by preventing charge recombination. Moreover, it can serve as a proton source [26]. With respect to thermodynamic studies, the photoreforming process is less challenging than water splitting, and it also reduces the reversible combination of H2 and O2. Therefore, the photoreforming process can be considered realistic; when the scavenging agents are abundant and easily accessible, sustainable and cost effective [15]. Various types of semiconducting materials have been established and utilized in photo-assisted H2 production processes. Despite noteworthy efforts, efficiency and promising results are far from commercialization. Among all the reported photocatalytic materials, TiO2 is most widely used because of its non-toxicity, chemical stability and low cost. In addition, TiO2 displays a high rate of recombination, and it has wide band gap of around 3.2 eV for anatase, which absorbs the UV light range. This results in low efficient materials towards photocatalysts because solar energy has only 3 to 4% of UV radiation [27]. Generally, an ideal photocatalyst should have the following properties: it should be efficient towards harvesting sun light and generating the redox charge carriers, it should have stability under working conditions and it should be more economical.
Along with this, some technical challenges are designing the appropriate semiconductors with effective charge carrier separation, a high rate of diffusion of the produced charge carriers and their movement towards the active centers of the surface, the appropriate arrangement of edge potentials in the CB and VB for the desired reactions, exposed reaction sites and easy preparation of catalytic materials without using infrequent materials [28]. This results in the enhancement of the efficiency of the catalyst, mainly depending on engineering materials, which should lead towards improving the significant parameters. The band gap, band positions and redox potentials of various types of photocatalytic materials are described in Figure 2.

3. Types of Organics (Oxygenates) in the Photoreforming Process for H2 Production

The photo-formation of H2 from different organic molecules in the existence of semiconducting materials was first introduced by Kawai and Sakata in the initial stage of the 1980s. Various organic molecules are used in photocatalytic reactions, i.e., sugars and carboxylic acids over RuO2/TiO2 [29] and Pt/TiO2 [30], respectively. In the last several decades, various sacrificial agents have been used in the production of hydrogen, for example, glycerol, glucose, ethanol [31], methanol [32], carboxylic acids [33], amino acids, biomass [34], fossil fuels, etc. Earlier reports suggest that the photo formation of H2 from water using hydrocarbons as a sacrificial mediator has recently attracted considerable attention. Research mainly focuses on the improvement of the efficiency of catalyst materials and also on the technological aspects and behavior of sacrificial agents towards efficiency. Various organic molecules have been demonstrated previously. Sacrificial agent/hole scavengers play a crucial role in the photoproduction of H2. The need for refilling sacrificial reagents in order to sustain high rates of reaction precludes their effective use for large-scale and continuous production of H2. Alternatively, sacrificial reagents can be used to assess the photocatalytic efficiency of H2 evolution [35]. Mainly, the fundamental mechanism associated with hydrogen-containing sacrificial reagents is whether these reagents simply serve as hole scavengers, which means that the species rapidly react with holes diminish their amount and thus stop their recombination with photogenerated electrons, leading towards higher efficiency. In addition, sacrificial agents may contribute their hydrogen atoms to produce H2 [36]. The most commonly used sacrificial agent is methanol due to its simplicity, its high amount of hydrogen, its understandable reaction mechanism and its capability of capturing h+. Earlier, methanol was used as an ideal feed stock in photoreforming [37]. The photochemical reaction of methanol and its degradation products are described as follows:
H2O (1) + h+OH + H+
CH3OH (1) + OH → CH2OH + H2O
CH2OH → HCHO + H+ + e
2H+ + 2e → H2 (g)
HCHO (1) + H2O (1) → HCOOH (1) + H2 (g)
HCOOH (1) → CO2 (g) + H2 (g)
Overall reaction:
CH3OH (l) + H2O → CO2 (g) + 3H2 (g)
For the first time, H2 production was performed using methanol by Kawai and Sakata in the 1980s. The proposed plausible mechanism predicted the formation of formaldehyde then into formic acid and further decomposing into H2 and carbon dioxide [29]. Li et al. reported the efficiency of hydrogen production in the presence of various sacrificial agents, thus showing improvement in photocatalytic activity by reducing the C/OH ratio. In addition, the scavengers react with hydroxyl radicals, enhancing the activity with the simplest molecule structure [38]. The systematic investigation was conducted using functionalized TiO2 with various noble and earth-abundant elements [39]. Recent research reports that copper-impregnated TiO2 [40], niobium pentoxides [41] MoS2/TiO2 [42] and H-Doped TiO2-x (H:TiO2-x) [43] are used in the photoreforming of methanol into hydrogen. Paramasivan et al. reported in situ simultaneous copper-deposition-enhanced photocatalytic hydrogen production from aqueous methanol solutions using ZnO as a photocatalyst. Hydrogen production was increased to be 130 times better than bare ZnO, and Cu-deposited Zn responded to visible light for hydrogen formation. The photogenerated holes attack methanol to form formaldehyde, which is further oxidized with OH radicals, and the holes produce formic acid. The photogenerated electrons reduce copper ions into a metallic copper cluster, which stores electrons. The electron sink role of Cu in ZnO enhances the separation of photogenerated electron–hole pairs and thus prevents the recombination of electron–hole pairs [44]. The schematic mechanism is shown in Figure 3a. Various transition metals as photocatalytic materials are used in hydrogen generation [45]. Furthermore, ethanol was investigated as a sacrificial agent in photoreforming H2 formation, because it can be obtained from biomass (cellulose and lignocellulose). Ethanol photoreforming includes the production of acetic acid and acetaldehyde, which is associated with the reduction of H+. There are other reaction pathways, including the direct reaction of h+ with ethanol that produces acetaldehyde by forming ethoxide ions on the catalyst surface and the oxidation of ethanol through OH· formation [46]. The produced byproducts are mainly acetaldehyde and CO, CO2, C2H4 and C2H6 in different amounts, as described in the following reaction [47,48]:
CH3CH2OH + TiO2(s) CH3CH2-Ti+4 + (s) OH
TiO2 + UV light → 2e (p) + 2h+
(s) CH3CH2O-Ti+4 + 2h+(s) CH3OH + Ti+4
(s) 2OH + e (p) → H2+ (S) 2O2
Here, (s) represents the surface of the photocatalyst, and (p) represents photoexcited electrons.
The ethanol oxidation produced H2 with 50% quantum efficiency using Pt/TiO2 catalyst. Recently, ethanol oxidation was carried out on Au and TiO2 (Figure 3b) [49] and in-situ photo-deposited Ni cocatalysts carbon nitride with cyanamide functionalities to enhance the activity and selectivity. The cyanamide functionalities mainly promote hole scavenging for ethanol oxidation [50]. Glycerol is a byproduct of the biodiesel and soap industry, which produces a large amount of 10 wt%. As it has less demand in the market, it is considered waste and is used in H2 production in photoreforming. As glycerol is a hydrogen source and an e releaser, various catalysts, such as engineered TiO2-based photocatalysts, are used. The photoreforming of glycerol produces acids, ketones and alcohols, as described in the following Equations (16) and (17) [51].
C3H8O3 + 3H2O + 14 h+ (VB) → intermediates (C2H4O2, C2H2O3, C2H4O3, C3H6O3, etc.) → 3CO2 + 14 H+
14 H+ + 14 e (CB) → 7 H2 (g)
Further research continued to use various catalysts, such as 2D Au/TiO2 [52], Pt/TiO2-Nb2O5 [53] and Au-NPs embedded in WO3/TiO2 (Figure 3c) [54] in recent studies for water and glycerol mixture reforming into hydrogen. Furthermore, polysaccharides are gaining more attention in the photoreforming process due to their thermodynamically acceptable nature with negative Gibbs free energy. Glucose can mainly be extracted from cellulose hydrolysis, and a higher amount of sugar is found in wastewater from food processing industries. Therefore, various photocatalysts are checked to obtain H2. The glucose-reforming mechanism was initially investigated in 1983 by Michael R. St. Jhon et al. on a platinized TiO2 surface [55]. It was thoroughly studied using d-glucose, and they proposed a detailed reaction mechanism in glucose reforming, as summarized in the following reaction Equations (18)–(23) [56,57].
C6H12O6 + H2O (ANA) → C5H10O5 + HCOOH + H2 (g)
C5H10O5 + H2O → C4H8O4 + HCOOH + H2 (g)
C4H8O4 + H2O + HCOOH + H2 (g) (A) → HCOOH + H2 (g) + CO2 (g)
C 6 H 12 O 6   TiO 2 ,   hv ,   H 2 O ,   O 2   C 6 H 12 O 7
C 6 H 12 O 7   TiO 2 ,   hv ,   H 2 O ,   O 2   C 6 H 10 O 8  
C 6 H 10 O 8   TiO 2 ,   hv ,   H 2 O ,   O 2   HCOOH + H 2 ( g ) + CO 2 ( g )
Here, ANA represents the anaerobic condition, and A represents the aerobic condition.
In recent studies, graphene dots functionalized with Pt cocatalysts were used in the photo reforming of cellulose in an alkaline medium, which forms glucose in the first step, and continuous H2 production for 6 days was achieved, as displayed in Figure 3d [58]. The photoreforming of protein molecules, amines and amino acids was studied using glycine and glutamic acid, which are the most common in living organisms [59]. The irradiation of these amino acids and proteins in the presence of Pt/TiO2 results in the production of H2 and CO2 in neutral water and H2 and NH3 in alkaline. Both NH3 in neutral solutions and CO2 in basic solutions were absorbed in water and not released in the gas phase [60]; moreover, CO2 can be reduced in valuable fuels [61]. As discussed above, along with photo conversion of biomass-derived compounds, the employment of raw biomass materials for photoreforming was also explored, in which photo generated holes oxidizes biomass substrates and electrons drove photoexcitation. However, the selective photo-transformation of unprocessed biomass into valuable products is a still difficult task. Very few efforts have been made to obtain the selective chemical transformation of unprocessed biomass into useful products. Initially, Greenbaum et al. used ZnO as a catalyst material for directly converting biomass into hydrocarbons and O2 under UV-visible light radiation [62]. In addition, Wu et al. reported an efficient approach toward producing functionalized aromatic compounds from lignin with a lignocellulosic biomass over CdS quantum dots. The mechanistic investigation indicated that both photoinduced electrons and holes are involved in the cleavage of the β-O-4 linkage of lignin via the EHCO mechanism. Most of the efforts are dedicated to sustainable hydrogen production by combining the simultaneous photo-oxidation of an aqueous biomass and the photocatalytic evolution of hydrogen at room temperature and atmospheric pressure [63]. Recently, an enhancement in the rate of H2 evolution was achieved in the photoreforming of lignocellulose with the irradiation of visible light. The high rate of H2 production was evidenced by the photoreforming of purified and raw lignocellulose over nanostructured carbon nitride under benign conditions without any toxic chemicals. Andrea Speltini et al. used oxidized g-C3N4 for the evolution of hydrogen assisted by an aqueous biomass in the presence of simulated solar radiation and compared the performance with that of P25-TiO2 [64]. Furthermore, Qiong Liu et al. evolved hydrogen by the photoreforming of biomass by using a functionalized terminal amino group in carbon nitride by in situ C–N coupling [65]. The photoreforming of the biomass into hydrogen and valuable products is carried out on an earth-abundant metal oxide, such as a Co/CoO hybrid structure with an excellent rate of 12 mmol/g/h, which is higher than the commercial catalyst [66]. The recent trend has been increasing catalyst activity with a lower cost and environmentally benign materials due to less energy demand for biomass reformation than photocatalytic water splitting.

4. Importance of Carboxylic Acids

As discussed in the previous section, along with alcohols, biomass, polysaccharides and carboxylic acids have been used as promising sacrificial agents in the photoproduction of H2. Alcoholic photoreforming proceeds through the abstraction of a hydrogen atom by a OH· radical from the methylene group and aldehydes. Therefore, alcoholic organic molecules decompose through the formation of carboxylic acid [67]. Furthermore, researchers have proved that carboxylic acid photoreforming into H2 production starts with a hole transfer to the carboxylic group instead of H abstraction from OH·. The carboxylic functionality is one of the oxygen-containing groups with excellent absorption and better reactivity in the photoreforming process [68]. Therefore, today, carboxylic acids are used as sacrificial agents for H2 production with hydrocarbons. Carboxylic acids are the intermediates in various reactions, such as hemicellulose derived from biomass, and the hydrolysis/pyrolysis of lignocellulose produces carboxylic acids, i.e., formic acid (FA) and acetic acid [69]. Moreover, the fermentation of saccharides produces carboxy functionality (lactic acid). In the photoreforming of acids, researchers have used direct or indirect acids (direct acids as a source and indirect acids as intermediates produced from biomass). In this process, the reaction occurs by binding the carboxylates to the photocatalyst surface, which processes decarboxylation, resulting in the formation of carbon dioxide followed by proton reduction into H2.
The CO2 evolution occurs via photo-Kolbe-type decarboxylation through carboxyl radicals. The efficiency of this process mainly depends on the design of the photocatalyst, reaction system, irradiation time and pH of the solution [70]. Various photocatalysts have been used for the degradation of carboxylic acids; most commonly, TiO2 has been introduced effectively in photoreforming due to its cost-effective, chemical and photochemical stabilities. TiO2 has been modified with various noble metals to obtain a significant hydrogen evolution rate. Yuexiang Li et al. investigated the evolution of hydrogen using oxalic acid on a Pt-decorated TiO2 catalyst, and the efficiency of Pt/TiO2 was carried out using oxalic acid, formic acid and formaldehyde [71,72]. Furthermore, the modification of TiO2 was performed using alkaline earth metals such as calcium, strontium and barium to enhance photocatalytic hydrogen evolution using formic acid, acetic acid and formaldehyde [73]. Various metal sulfides and oxides such as cadmium sulfides, vanadium pentoxides, iron oxides, zinc oxides and zinc sulfides are reviewed systematically in further individual sections on acid photoreforming.

4.1. Formic Acid

Formic acid (FA) is colorless, less toxic and completely biodegradable [74], and it can be stored/used at room temperature. Therefore, FA is the easiest root for hydrogen generation, with a higher volumetric and gravimetric H2 evolution capacity. This acid can be produced from photo-electrocatalytic CO2 reduction and the decomposition of biomass [75]. H2 generation during FA photoreforming and selectivity is an important aspect on which H2 evolution quality depends, along with reaction conditions, catalyst materials and temperature. Since the early 20th century, various studies using FA have developed several homogeneous and heterogeneous photocatalysts. TiO2 semiconducting materials have been used for H2 production by modifying them using noble metals as co-catalysts to enhance the efficiency of the photocatalysts. The combination of semiconductors with noble metals has been studied for photocatalytic FA degradation in H2 production with a higher quantum yield and a high production rate. Yuexiang Li et al. used Pt/TiO2 for H2 production through FA photoreforming. The overall H2 evolution rate was consistent with that of the decomposition of FA [72]. Chen Tao et al. used Pt/TiO2 for the anaerobic photocatalytic degradation of FA. In this reaction, adsorbed molecules of FA converted into formate species and transformed into carbonate species with the further formation of H2 while adding water vapor into the reaction media [76]. Alexia Patsoura et al. used Pt/TiO2 as a photocatalyst for the photoreforming of alcohol and organic acid [17]. Furthermore, Chiarello et al. investigated the formation of H2 using noble metals Au-, Ag- and Au–Ag-modified TiO2 with methanol steam with the formation of FA as an intermediate for H2 production. There are various byproducts that are formed, such as carbon monoxide, methane, acetaldehyde and methyl formate, with constant rates of H2 evolution and CO2 formation [39]. Furthermore, Valeriano Lanese et al. developed copper-doped TiO2 for aqueous FA photoreforming. The oxidation states of Cu affect the efficiency of hydrogen production. Zerovalent Cu does not affect hydrogen evolution, and cupric ions enhance the rate and quantity of hydrogen production while maintaining the experimental parameters in both the reactions [77]. Hainer et al. reported various metal-doped (Cu, Ru, Pt and Au) TiO2 for hydrogen evolution using methanol and FA. In this study, Pt/TiO2 showed higher H2 production using methanol. Cu/TiO2 and Ru/TiO2 showed higher yields towards hydrogen production using FA. Au/TiO2 was responsible for the high rate of hydrogen production using FA [78].
Jamal et al. selectively produced hydrogen photo-catalytically from FA efficiently using non-noble metals, such as cobalt and nickel, as redox mediators on CdS nanorods, as shown in the following. The effect of co-catalysts was also studied on the decomposition of FA. Here, the high hydrogen production of 32.6 mmol h−1g−1 from FA was obtained by sustaining the production rate for 12 h [79]. The mechanism involved a substantial increase in FA-to-H2 activity for Ni/CdS-NR, the quenching of the PL signal and the reduction of Ni+2 to Ni0 in the reaction system by photogenerated electrons at the CB of CdS. This band offset between CdS and Ni0 provides a conduit for electrons to pass from CdS to Ni0. The Co is not deposited on the CdS, but it is present in the reaction media. The observed enhanced activity presumably occurs through transitory contact with the CdS, forming a bound Co-hole formate species. Therefore, the improved catalysis is due to the cooperative redox mediation of Ni0 and Co3+, where both the reductive and oxidative half-reactions are enhanced, as shown in Figure 4. Further various co-catalysts are developed to increase hydrogen production with great sustainability, which are discussed in Table 1.

4.2. Acetic Acid

Most of the studies have employed organic acids as electron donors in hydrogen production in the photoreforming process as acetic acid, since it is a common organic molecule in industrial waste [99]. The use of acetic acid to produce H2 allows the environment-friendly pathway by wastewater treatment with desired valuable products. The degradation of acetic acid through the photoreforming process into valuable products is revolutionary, as it does not require a high capital cost. The use of acetic acid as a sacrificial agent in photocatalysis leads to the formation of H2 and various hydrocarbons, such as CH4 and C2H6, which can be further used as renewable fuels [100]. Various studies have been reported by Krauetler and Bard, which discuss the photocatalytic conversion of acetic acid into H2 and byproducts as hydrocarbons over Pt/TiO2 in acidic conditions. Initially, the decomposition of acetic acid using visible light results in the production of CH4 and CO2 due to a more favorable thermodynamic nature, as compared to the reforming of acetic acid [101]. The following scheme shows a plausible reaction pathway for aqueous acetic acid photocatalytic transformation on the photocatalyst surface.
PC   +   h ν   PC ( h + +   e )
PC ( e ) +   H + Pt   PC +   H ·
PC   ( h + ) +   CH 3 COOH   PC +   CH 3 COO · +   H +
PC   ( h + ) +   C 2 H 5 COOH   PC +   C 2 H 5 COO · +   H +
PC   ( h + ) +   H 2 O   PC + · OH +   H +
PC   ( h + ) + OH   PC + · OH
Formation of · CH 3 and · C 2 H 5
CH 3 COO ·     · CH 3 +   CO 2
C 2 H 5 COO ·     · C 2 H 5 +   CO 2
Formation of molecular hydrogen and alkanes (Kolbe products)
H · +   H ·   Pt   H 2
· CH 3 +   H ·     CH 4
· CH 3 + · CH 3     C 2 H 6
· C 2 H 5 +   H ·     C 2 H 6
· C 2 H 5 + · CH 3     C 3 H 8
Formation of C2H5COOH
CH 3 COOH + · OH     · CH 2 COOH +   H 2 O
· CH 2 COOH + · CH 3     C 2 H 5 COOH
Formation of alcohols (Hofer–Moest products)
· OH + · CH 3     CH 3 OH  
· OH + · C 2 H 5     C 2 H 5 OH
In these studies, the amount of hydrogen production is seen to be lower, therefore increasing the efficiency of the reaction towards the high quantity production of H2. Hlroshl et al. demonstrated the photoreforming of acetic acid into valuable products (i.e., CH4 H2) using Pt/TiO2 suspensions at various pH. The main reaction product observed was methane but increased the H2/CH4 ratio by increasing the pH [102]. Tadayoshi Sakata et al. studied the heterogeneous photocatalyst for the reaction of acetic, propionic and butyric acid on different semiconducting materials. The type of semiconductor and pH of the reaction system was studied. The production of hydrocarbons had completely stopped, and H2 evolution was increased by increasing the pH from acidic to neutral and alkaline. This study predicted that the position of the valence band of the semiconducting material depends on the pH and that the reaction path is controlled by the amount of OH- ions [100]. Xian-Jun Zheng et al. reported the photocatalysis of acetic acid over Pt/TiO2 in UV radiation. The effects of various parameters were checked with an optimal amount of photo-deposited Pt on TiO2 [103]. In addition, CuO/SnO2 as the photocatalyst for acetic acid reformation into H2 has been used [104]. Recently, Mikel Imizcoz et al. reported an earth-abundant co-catalyst on TiO2 for the selective production of H2 from acetic acid derived from biomass feedstocks. They reported higher decarboxylation activity due to copper nanoparticle insertion in semiconducting materials and also that the selectivity of H2 production increases with photo-deposited Cu in aqueous acetic suspensions as compared to Pt/TiO2 [105]. The simple mechanisms are depicted in Figure 5. Upon solar irradiation, photogenerated electrons from the CB of TiO2 are capable of moving to CuO nanoparticles and reducing them. H2 evolution increased marginally as copper mass loading increased due to the maximized surface reduction of copper. Various co-catalysts with semiconducting materials for acetic acid reforming are listed below in Table 2.

4.3. Lactic Acid

Lactic acid (LA) is an industrial product, and the polymers derived from acetic acid are biodegradable and can be synthesized from the fermentation of carbohydrates and biomass. LA has been used in several medical, textile and pharmaceutical industries [116]. LA is employed as a supplement in the synthesis of pharmaceuticals as well as in the production of hygiene and cosmetic goods in the cosmetics sector [117]. Due to this industrial application, LA is absorbed into water bodies. The purification of water, as well as the removal of LA, is needed. To overcome these issues, there must be reformation of LA into valuable products and hydrogen fuel. There are various strategies to produce H2, such as photon-assisted water splitting and the photoreforming process. LA is used as a sacrificial agent in the photo-production of H2 [118]. As discussed in the previous section, co-catalyst-assisted semiconducting materials are used as photocatalysts in the photoreforming process. Initially, Harada et al. found the decomposition of LA over platinized CdS and TiO2. The obtained quantum efficiencies for the Pt/TiO2 and Pt/CdS were 71% (360 nm) and 38% (440 nm). The products obtained during the photocatalytic decomposition of LA on Pt/TiO2 were acetaldehyde, carbon dioxide and hydrogen, and by using Pt/CdS, they were hydrogen and pyruvic acid [119,120].
Furthermore, LA photoreforming was carried out using a non-noble cost-effective metal as a cocatalyst with cadmium sulfides as the photocatalysts under visible light. Wei Zhang et al. reported the quantum efficiency for H2 production of 51.3% under the visible light range (420 nm) [120]. Qin et al. investigated photoreforming using LA as a sacrificial agent on a CdS cluster-decorated graphene nanosheet as a photocatalyst and Pt as a co-catalyst. The photoreforming was carried out under visible radiation with an increase in the rate and quantity of H2 by increasing the content of graphene in the catalyst compared to bare CdS due to an increase in crystallinity and surface area. They reported that the use of CdS as a photo-corrosive agent and of a Pt co-catalyst reduces overpotential in H2 production and opposes the backward combination reactions with a high rate of H2 (1.12 m mol h−1) and 22.5% of quantum efficiency at 420 nm [121]. Xu Zong et al. demonstrated that MoS2-loaded CdS catalyst showed a higher rate and quantity of H2 evolution from LA under visible light. MoS2/CdS was increased 36 times more than the bare CdS and had higher activity than the Pt/CdS [122]. Zhuofeng Hu et al. produced an innovative photocatalyst as a platinum cobalt alloy on CdS and TiO2 by a simple polyol reduction method. The introduction of cobalt as a cocatalyst with Pt increased composite conductivity, resulting in enhanced H2 evolution from acetic acid decomposition on Pt3Co/CdS and Pt3Co/TiO2 under visible light radiation with 15.86 and 13.01 m mol g−1 h−1, respectively [123]. Recently, Chunhe Li loaded NiS nanoparticles on CdS through solvothermal synthesis and used as a catalyst for the evolution of hydrogen from lignin and LA as a hole scavenger. The activity of NiS/CdS is 5041 time more than the bare CdS and has an apparent quantum efficiency of 44.9% for hydrogen evolution with five repetitive cycles [124]. This report, as shown in Figure 6c, generated charge carriers in CdS when it was excited with light. Even though the CB band of CdS is more negative than the reduction potential of H+/H2, the rate of hydrogen evolution is low on CdS because of the fast recombination of photoinduced charge carriers. When the NiS nanoparticles are loaded on the surface of CdS, because of the less negative CB of NiS than that of CdS, photogenerated charge carriers transfer to the NiS nanoparticles. Mainly, the intimate interfacial junction between NiS and CdS plays an essential role in facilitating the transmitting of electrons from CdS to NiS. NiS acts as an active site because the unsaturated sulfide ions of NiS possess more affinity towards H+ and thus enhance hydrogen evolution. The details of the photoreforming of LA in the coexistence of lignin under visible light radiation are described in Figure 6. Various reports on photoreformation of LA are tabulated in Table 3.

4.4. Other Carboxylic Acids

Including formic acid, acetic acid and lactic acid, other acids such as oxalic acid, butyric acid, pyruvic acid and gluconic acid have been used in the photo-assisted production of H2 through the decarboxylation and photoreforming process. Acidic functionalities undergo decarboxylation into alkanes (for example, ethane, propane and butane). The photoreforming H2 evolution increases with increases in the alkane chain length. H2 fuel, also produced from fatty acids through steric acid as an intermediate, has been carried out using platinized TiO2 [141]. The oxidation of saccharides processed using biomass feedstocks results in forming acidic derivatives. Gluconic and threonine salts act as the substrate producing H2 and CO2 in the presence of Pt/TiO2 in the presence of UV-rich light [60]. Yuexiang Li used oxalic acid as a sacrificial agent on Pt/TiO2 in photoreforming, which produced H2 and CO2 in 1:2 molar ratio [72]. There are many reports on the use of different acids as sacrificial agents in photoreforming for the production of H2, and other organic carboxylic acids are described in Table 4.

5. Future Aspects

To gain a high rate and quantity of hydrogen evolution from various biomasses and organic molecules through photoreforming, the evolution of hydrogen (half-reaction) must be carefully enhanced by controlling the selectivity of the photoreforming of organic compounds into hydrogen and preventing the production of harmful organic waste products and also the development of value-added products. Subsequently, most researchers have focused on enhancing the rate and efficiency of selective hydrogen production through photoreforming. Various approaches are described in Figure 7, regarding catalyst design, the selection of co-catalysts, reaction media, advanced characterization and computational studies. The designing of photocatalysts is the most challenging field. Most researchers are moving towards using environment-friendly and cost-effective semiconducting material by synthesizing composites with various co-catalysts. The selection of cocatalysts must be made using earth-abundant, non-toxic and environment-friendly materials instead of precious and toxic metals (Pt, Ru, Ag and Pd). Most of the above-mentioned photocatalysts exhibit reasonable efficiency of organic acid conversion into H2 and valuable products, but depth of the understanding of the mechanism, advanced characterization and computational studies are required to make use of photoreforming on an industrial scale and for commercialization. The photoreforming process of the conversion of organics into hydrogen fuel must be benign with a cost-effective photocatalyst, selectivity and efficiency.

6. Conclusions

In summary, the photocatalytic method has become the most challenging technique for renewable, zero-emission energy harvesting in the last few decades. The scientific community has shown excessive interest in employing photoactive materials and various reaction parameters for clean H2 production using solar energy through water splitting and the photoreforming of organic molecules. When employing the right catalyst, sacrificial agent and solar energy to produce H2, the organic process has been shown to be more advantageous and to have more industrial uses than the traditional water splitting method. The main challenge in photoreforming is developing an efficient, stable photocatalyst that can absorb sunlight. Along with this, important issues must be overcome, such as the development of co-catalysts without using noble, precious metals. Cost-effective, environmentally benign and abundant metals must be used as cocatalysts for commercialization and practical application. Recently, researchers have developed novel nanostructures with controlled morphology at the nanoscale and with multiphase composition. In addition, more attention should be given to addressing specific issues such as the reversible recombination of charges, and catalytic materials should be beneficial for H2 photoproduction, therefore contributing to the appearance of great promise in this field. The combination of two or more materials can be performed towards optimizing photocatalytic activity. The choice of the proper material and understanding the mechanism of the catalyst material is needed, but it is still challenging in some reactions. There are various semiconductors developed with optimized band energy gaps and co-catalysts with photoactivity to produce hydrogen from oxygenated organics. Various organics are reformed into renewable sources and valuable hydrocarbons using UV-vis radiation and semiconducting materials. The catalyst is developed from bare TiO2, a noble metal (Pt, Ru, Pd and Au) and TiO2 to a transition metal and less toxic and cost-effective cocatalysts loaded with TiO2 for carboxylic acid reforming. Furthermore, CdS and ZnS are used as semiconducting materials loaded with MoS2 and Cu2O for the formation of hydrogen from various carboxylic acids. The photocatalytic conversion of carboxylic acids occurs by decarboxylation and reforming into H2, carbon monoxide, carbon dioxide and hydrocarbons. Overall, the critical factors for the efficiency of photoreforming are selectivity towards the high rate and quantity of hydrogen evolution, efficiency, sustainability and long-term chemical and photocatalytic stability, and reaction conditions with cost-effective, environmentally benign and easily available photocatalysts should be developed for carboxylic acid photoreforming.

Author Contributions

A.S. (Apurba Sinhamahapatra) and S.K.N. conceived the idea and finalized the manuscript. A.S. (Anita Samage), P.G. and M.A.H. performed the literature search and drafting. All authors have read and agreed to the published version of the manuscript.

Funding

S.K.N. would like to thank the Department of Science and Technology (Government of India) for the NANO MISSION Project (SR/NM/NT-1073/2016) and DST INSPIRE (IFA12-CH-84) and the DST-Technology Mission Division (DST/TMD/HFC/2K18/124G) (Government of India) for its economic support. S.K.N. acknowledges the “Talent Attraction Programme funded by the Community of Madrid Spain/Spanish (2017-T1/AMB5610)”. A.S. (Apurba Sinhamahapatra) acknowledges the research funds from the Department of Science and Technology (Government of India) for the DST INSPIRE Fellowship (IFA17-MS107).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Balali, Y.; Stegen, S. Review of energy storage systems for vehicles based on technology, environmental impacts, and costs. Renew. Sustain. Energy Rev. 2021, 135, 110185. [Google Scholar] [CrossRef]
  2. Bhuiyan, M.A.; Khan, H.U.R.; Zaman, K.; Hishan, S.S. Measuring the impact of global tropospheric ozone, carbon dioxide and sulfur dioxide concentrations on biodiversity loss. Environ. Res. 2018, 160, 398–411. [Google Scholar] [CrossRef] [PubMed]
  3. Blanco, H.; Nijs, W.; Ruf, J.; Faaij, A. Potential for hydrogen and Power-to-Liquid in a low-carbon EU energy system using cost optimization. Appl. Energy 2018, 232, 617–639. [Google Scholar] [CrossRef]
  4. Chen, J.; Wang, Q.; Xu, Z.; Jiaqiang, E.; Leng, E.; Zhang, F.; Liao, G. Process in supercritical water gasification of coal: A review of fundamentals, mechanisms, catalysts and element transformation. Energ. Convers. Manag. 2021, 237, 114122. [Google Scholar] [CrossRef]
  5. Chen, C.; Wood, P. A turbulence closure model for dilute gas-particle flows. Can. J. Chem. Eng. 1985, 63, 349–360. [Google Scholar] [CrossRef]
  6. Nguyen, H.M.; Sunarso, J.; Li, C.; Pham, G.H.; Phan, C.; Liu, S. Microwave-assisted catalytic methane reforming: A review. Appl. Catal. A. Gen. 2020, 599, 117620. [Google Scholar] [CrossRef]
  7. Ahmadi, M.H.; Ghazvini, M.; Alhuyi Nazari, M.; Ahmadi, M.A.; Pourfayaz, F.; Lorenzini, G.; Ming, T. Renewable energy harvesting with the application of nanotechnology: A review. Int. J. Energy Res. 2019, 43, 1387–1410. [Google Scholar] [CrossRef]
  8. Luo, P. perspectives in photo-and electrochemical-oxidation of biomass for sustainable chemicals and hydrogen production. Adv. Energy Mater 2021, 11, 2101180. [Google Scholar] [CrossRef]
  9. Luo, H.; Zeng, Z.; Zeng, G.; Zhang, C.; Xiao, R.; Huang, D.; Lai, C.; Cheng, M.; Wang, W.; Xiong, W. Recent progress on metal-organic frameworks based-and derived-photocatalysts for water splitting. Chem. Eng. J. 2020, 383, 123196. [Google Scholar] [CrossRef]
  10. O’regan, B.; Grätzel, M. A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 1991, 353, 737–740. [Google Scholar] [CrossRef]
  11. Navarro, R.M.; Sanchez-Sanchez, M.C.; Alvarez-Galvan, M.C.; Valle, F.; Fierro, J.L. Hydrogen production from renewable sources: Biomass and photocatalytic opportunities. Energy Environ. Science. 2009, 2, 35–54. [Google Scholar] [CrossRef]
  12. Colmenares, J.C.; Magdziarz, A.; Aramendia, M.A.; Marinas, A.; Marinas, J.M.; Urbano, F.J.; Navio, J.A. Influence of the strong metal support interaction effect (SMSI) of Pt/TiO2 and Pd/TiO2 systems in the photocatalytic biohydrogen production from glucose solution. Catal. Commun. 2011, 16, 1–6. [Google Scholar] [CrossRef]
  13. Bowker, M. Sustainable hydrogen production by the application of ambient temperature photocatalysis. Green Chem. 2011, 13, 2235–2246. [Google Scholar] [CrossRef]
  14. Ran, J.; Gao, G.; Li, F.-T.; Ma, T.-Y.; Du, A.; Qiao, S.-Z. Ti3C2 MXene co-catalyst on metal sulfide photo-absorbers for enhanced visible-light photocatalytic hydrogen production. Nat. Commun. 2017, 8, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Lakhera, S.K.; Rajan, A.; Rugma, T.; Bernaurdshaw, N. A review on particulate photocatalytic hydrogen production system: Progress made in achieving high energy conversion efficiency and key challenges ahead. Renew. Sustain. Energy Rev. 2021, 152, 111694. [Google Scholar] [CrossRef]
  16. Arkhipova, N.; Kuznetsova, A. Photocatalytic activity and physicochemical characteristics of modified potassium polytitanates in the reaction of decomposition of aqueous-alcoholic solutions. Russ. J. Appl. Chem. 2017, 90, 186–192. [Google Scholar] [CrossRef]
  17. Patsoura, A.; Kondarides, D.I.; Verykios, X.E. Photocatalytic degradation of organic pollutants with simultaneous production of hydrogen. Catal. Today. 2007, 124, 94–102. [Google Scholar] [CrossRef]
  18. Christoforidis, K.; Fornasiero, P. Photocatalytic hydrogen production: A rift into the future energy supply. ChemCatChem 2017, 9, 1523–1544. [Google Scholar] [CrossRef] [Green Version]
  19. Colmenares, J.C.; Luque, R. Heterogeneous photocatalytic nanomaterials: Prospects and challenges in selective transformations of biomass-derived compounds. Chem. Soc. Rev. 2014, 43, 765–778. [Google Scholar] [CrossRef]
  20. Quispe, C.A.; Coronado, C.J.; Carvalho, J.A., Jr. Glycerol: Production, consumption, prices, characterization and new trends in combustion. Renew. Sustain. Energy Rev. 2013, 27, 475–493. [Google Scholar] [CrossRef]
  21. Pagliaro, M.; Ciriminna, R.; Kimura, H.; Rossi, M.; Della Pina, C. From glycerol to value-added products. Angew. Chem. Int. Edition 2007, 46, 4434–4440. [Google Scholar] [CrossRef] [PubMed]
  22. Toe, C.Y.; Tsounis, C.; Zhang, J.; Masood, H.; Gunawan, D.; Scott, J.; Amal, R. Advancing photoreforming of organics: Highlights on photocatalyst and system designs for selective oxidation reactions. Energy Environ. Sci. 2021, 14, 1140–1175. [Google Scholar] [CrossRef]
  23. Sinhamahapatra, A.; Jeon, J.-P.; Yu, J.-S. A new approach to prepare highly active and stable black titania for visible light-assisted hydrogen production. Energy Environ. Sci. 2015, 8, 3539–3544. [Google Scholar] [CrossRef] [Green Version]
  24. McKone, J.R.; Pieterick, A.P.; Gray, H.B.; Lewis, N.S. Hydrogen evolution from Pt/Ru-coated p-type WSe2 photocathodes. J. Am. Chem. Soc. 2013, 135, 223–231. [Google Scholar] [CrossRef] [PubMed]
  25. Walte, M.; Warren, E.; McKone, J.; Boettcher, S.; Qixi, M.; Santori, E.; Lewis, N. Solar water splitting cell. Chem. Rev. 2010, 110, 6446–6473. [Google Scholar] [CrossRef] [PubMed]
  26. Rossetti, I. Hydrogen production by photoreforming of renewable substrates. Int. Sch. Res. Notices 2012, 2012. [Google Scholar] [CrossRef] [Green Version]
  27. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light photocatalysis in nitrogen-doped titanium oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef]
  28. Khan, K.; Tareen, A.K.; Aslam, M.; Sagar, R.U.R.; Zhang, B.; Huang, W.; Mahmood, A.; Mahmood, N.; Khan, K.; Zhang, H. Recent progress, challenges, and prospects in two-dimensional photocatalyst materials and environmental remediation. Micro Nano Lett. 2020, 12, 1–77. [Google Scholar] [CrossRef]
  29. Kawai, T.; Sakata, T. Conversion of carbohydrate into hydrogen fuel by a photocatalytic process. Nature 1980, 286, 474–476. [Google Scholar] [CrossRef]
  30. Sakata, T.; Hashimoto, K.; Kawai, T. Catalytic properties of ruthenium oxide on n-type semiconductors under illumination. J. Phys. Chem. A 1984, 88, 5214–5221. [Google Scholar] [CrossRef]
  31. Pajares, A.; Wang, Y.; Kronenberg, M.; de la Piscina, P.R.; Homs, N. Photocatalytic H2 production from ethanol aqueous solution using TiO2 with tungsten carbide nanoparticles as co-catalyst. Int. J. Hydrogen Energy 2020, 45, 20558–20567. [Google Scholar] [CrossRef]
  32. Alshehri, A.; Narasimharao, K. PtOx-TiO2 anatase nanomaterials for photocatalytic reformation of methanol to hydrogen: Effect of TiO2 morphology. J. Mater. Res. Technol. 2020, 9, 14907–14921. [Google Scholar] [CrossRef]
  33. Hippargi, G.; Anjankar, S.; Krupadam, R.J.; Rayalu, S.S. Simultaneous wastewater treatment and generation of blended fuel methane and hydrogen using Au-Pt/TiO2 photo-reforming catalytic material. Fuel 2021, 291, 120113. [Google Scholar] [CrossRef]
  34. Huang, C.-W.; Nguyen, B.-S.; Wu, J.C.-S.; Nguyen, V.-H. A current perspective for photocatalysis towards the hydrogen production from biomass-derived organic substances and water. Int. J. Hydrogen Energy 2020, 45, 18144–18159. [Google Scholar] [CrossRef]
  35. Kumaravel, V.; Imam, M.D.; Badreldin, A.; Chava, R.K.; Do, J.Y.; Kang, M.; Abdel-Wahab, A. Photocatalytic hydrogen production: Role of sacrificial reagents on the activity of oxide, carbon, and sulfide catalysts. Catalysts 2019, 9, 276. [Google Scholar] [CrossRef] [Green Version]
  36. Guzman, F.; Chuang, S.S.; Yang, C. Role of methanol sacrificing reagent in the photocatalytic evolution of hydrogen. Ind. Eng. Chem. Res. 2013, 52, 61–65. [Google Scholar] [CrossRef]
  37. Ismael, M. Latest progress on the key operating parameters affecting the photocatalytic activity of TiO2-based photocatalysts for hydrogen fuel production: A comprehensive review. Fuel 2021, 303, 121207. [Google Scholar] [CrossRef]
  38. Li, Y.; Wang, B.; Liu, S.; Duan, X.; Hu, Z. Synthesis and characterization of Cu2O/TiO2 photocatalysts for H2 evolution from aqueous solution with different scavengers. Appl. Surf. Sci. 2015, 324, 736–744. [Google Scholar] [CrossRef]
  39. Chiarello, G.L.; Forni, L.; Selli, E. Photocatalytic hydrogen production by liquid-andgas-phasereforming of CH3OHoverflame-made TiO2 andAu1 TiO2. Catal. Today 2009, 144, 69–74. [Google Scholar] [CrossRef]
  40. Vitiello, G.; Clarizia, L.; Abdelraheem, W.; Esposito, S.; Bonelli, B.; Ditaranto, N.; Vergara, A.; Nadagouda, M.; Dionysiou, D.D.; Andreozzi, R. Near UV-Irradiation of CuOx-Impregnated TiO2 Providing Active Species for H2 Production Through Methanol Photoreforming. ChemCatChem 2019, 11, 4314–4326. [Google Scholar] [CrossRef]
  41. Asencios, Y.J.; Machado, V.A. Photodegradation of Organic Pollutants in Seawater and Hydrogen Production via Methanol Photoreforming with Hydrated Niobium Pentoxide Catalysts. Sustain. Chem. 2022, 3, 172–191. [Google Scholar] [CrossRef]
  42. Saha, A.; Sinhamahapatra, A.; Kang, T.-H.; Ghosh, S.C.; Yu, J.-S.; Panda, A.B. Hydrogenated MoS2 QD-TiO2 heterojunction mediated efficient solar hydrogen production. Nanoscale 2017, 9, 17029–17036. [Google Scholar] [CrossRef]
  43. Sinhamahapatra, A.; Lee, H.-Y.; Shen, S.; Mao, S.S.; Yu, J.-S. H-doped TiO2-x prepared with MgH2 for highly efficient solar-driven hydrogen production. Appl. Catal. B 2018, 237, 613–621. [Google Scholar] [CrossRef]
  44. Gomathisankar, P.; Hachisuka, K.; Katsumata, H.; Suzuki, T.; Funasaka, K.; Kaneco, S. Enhanced photocatalytic hydrogen production from aqueous methanol solution using ZnO with simultaneous photodeposition of Cu. Int. J. Hydrogen Energy 2013, 38, 11840–11846. [Google Scholar] [CrossRef]
  45. Sinhamahapatra, A.; Jeon, J.-P.; Kang, J.; Han, B.; Yu, J.-S. Oxygen-deficient zirconia (ZrO2−x): A new material for solar light absorption. Sci. Rep. 2016, 6, 1–8. [Google Scholar] [CrossRef] [PubMed]
  46. Nguyen-Phan, T.-D.; Baber, A.E.; Rodriguez, J.A.; Senanayake, S.D. Au and Pt nanoparticle supported catalysts tailored for H2 production: From models to powder catalysts. Appl. Catal. A Gen. 2016, 518, 18–47. [Google Scholar] [CrossRef] [Green Version]
  47. Sampaio, M.J.; Oliveira, J.W.; Sombrio, C.I.; Baptista, D.L.; Teixeira, S.R.; Carabineiro, S.A.; Silva, C.G.; Faria, J.L. Photocatalytic performance of Au/ZnO nanocatalysts for hydrogen production from ethanol. Appl. Catal. A Gen. 2016, 518, 198–205. [Google Scholar] [CrossRef]
  48. Montini, T.; Gombac, V.; Sordelli, L.; Delgado, J.J.; Chen, X.; Adami, G.; Fornasiero, P. Nanostructured Cu/TiO2 Photocatalysts for H2 Production from Ethanol and Glycerol Aqueous Solutions. ChemCatChem 2011, 3, 574–577. [Google Scholar] [CrossRef]
  49. Zhang, X.; Luo, L.; Yun, R.; Pu, M.; Zhang, B.; Xiang, X. Increasing the activity and selectivity of TiO2-supported Au catalysts for renewable hydrogen generation from ethanol photoreforming by engineering Ti3+ defects. ACS Sustain. Chem. Eng. 2019, 7, 13856–13864. [Google Scholar] [CrossRef]
  50. Gunawan, D.; Toe, C.Y.; Kumar, P.; Scott, J.; Amal, R. Synergistic Cyanamide Functionalization and Charge-Induced Activation of Nickel/Carbon Nitride for Enhanced Selective Photoreforming of Ethanol. ACS Appl. Mater. Interfaces 2021, 13, 49916–49926. [Google Scholar] [CrossRef]
  51. Tran, N.H.; Kannangara, G.K. Conversion of glycerol to hydrogen rich gas. Chem. Soc. Rev. 2013, 42, 9454–9479. [Google Scholar] [CrossRef] [PubMed]
  52. Zhong, W.; Wang, C.; Zhao, H.; Peng, S.; Tian, Z.; Shu, R.; Chen, Y. Synergistic Effect of Photo-thermal Catalytic Glycerol Reforming Hydrogen Production over 2D Au/TiO2 Nanoflakes. Chem. Eng. J. 2022, 137063. [Google Scholar] [CrossRef]
  53. Iervolino, G.; Vaiano, V.; Murcia, J.; Lara, A.; Hernández, J.; Rojas, H.; Navío, J.; Hidalgo, M. Photocatalytic production of hydrogen and methane from glycerol reforming over Pt/TiO2–Nb2O5. Int. J. Hydrogen Energy 2021, 46, 38678–38691. [Google Scholar] [CrossRef]
  54. Tahir, M.; Siraj, M.; Tahir, B.; Umer, M.; Alias, H.; Othman, N. Au-NPs embedded Z–scheme WO3/TiO2 nanocomposite for plasmon-assisted photocatalytic glycerol-water reforming towards enhanced H2 evolution. Appl. Surf. Sci. 2020, 503, 144344. [Google Scholar] [CrossRef]
  55. St. John, M.R.; Furgala, A.J.; Sammells, A.F. Hydrogen generation by photocatalytic oxidation of glucose by platinized n-titania powder. J. Phys. Chem. A 1983, 87, 801–805. [Google Scholar] [CrossRef]
  56. Fu, X.; Long, J.; Wang, X.; Leung, D.Y.; Ding, Z.; Wu, L.; Zhang, Z.; Li, Z.; Fu, X. Photocatalytic reforming of biomass: A systematic study of hydrogen evolution from glucose solution. Int. J. Hydrogen Energy 2008, 33, 6484–6491. [Google Scholar] [CrossRef]
  57. Iervolino, G.; Vaiano, V.; Sannino, D.; Rizzo, L.; Ciambelli, P. Production of hydrogen from glucose by LaFeO3 based photocatalytic process during water treatment. Int. J. Hydrogen Energy 2016, 41, 959–966. [Google Scholar] [CrossRef]
  58. Nguyen, V.-C.; Nimbalkar, D.B.; Nam, L.D.; Lee, Y.-L.; Teng, H. Photocatalytic cellulose reforming for H2 and formate production by using graphene oxide-dot catalysts. ACS Catal. 2021, 11, 4955–4967. [Google Scholar] [CrossRef]
  59. Mondal, A.; Biswas, S.; Kumar, A.; Yu, J.S.; Sinhamahapatra, A. Sub 10 nm CoO nanoparticle-decorated graphitic carbon nitride for solar hydrogen generation via efficient charge separation. Nanoscale Adv. 2020, 2, 4473–4481. [Google Scholar] [CrossRef]
  60. Kawai, T.; Sakata, T. Photocatalytic hydrogen production from water by the decomposition of poly-vinylchloride, protein, algae, dead insects, and excrement. Chem. Lett. 1981, 10, 81–84. [Google Scholar] [CrossRef] [Green Version]
  61. Razzaq, A.; Sinhamahapatra, A.; Kang, T.-H.; Grimes, C.A.; Yu, J.-S.; In, S.-I. Efficient solar light photoreduction of CO2 to hydrocarbon fuels via magnesiothermally reduced TiO2 photocatalyst. Appl. Catal. B 2017, 215, 28–35. [Google Scholar] [CrossRef]
  62. Greenbaum, E.; Tevault, C.V.; Ma, C. New Photosynthesis: Direct photoconversion of biomass to molecular oxygen and volatile hydrocarbons. Energy Fuels 1995, 9, 163–167. [Google Scholar] [CrossRef]
  63. Wu, X.; Fan, X.; Xie, S.; Lin, J.; Cheng, J.; Zhang, Q.; Chen, L.; Wang, Y. Solar energy-driven lignin-first approach to full utilization of lignocellulosic biomass under mild conditions. Nat. Catal. 2018, 1, 772–780. [Google Scholar] [CrossRef]
  64. Speltini, A.; Gualco, F.; Maraschi, F.; Sturini, M.; Dondi, D.; Malavasi, L.; Profumo, A. Photocatalytic hydrogen evolution assisted by aqueous (waste) biomass under simulated solar light: Oxidized g-C3N4 vs. P25 titanium dioxide. Int. J. Hydrogen Energy 2019, 44, 4072–4078. [Google Scholar] [CrossRef]
  65. Liu, Q.; Wei, L.; Xi, Q.; Lei, Y.; Wang, F. Edge functionalization of terminal amino group in carbon nitride by in-situ C–N coupling for photoreforming of biomass into H2. Chem. Eng. J. 2020, 383, 123792. [Google Scholar] [CrossRef]
  66. Guo, P.; Xiong, Z.; Yuan, S.; Xie, K.; Wang, H.; Gao, Y. The synergistic effect of Co/CoO hybrid structure combined with biomass materials promotes photocatalytic hydrogen evolution. Chem. Eng. J. 2021, 420, 130372. [Google Scholar] [CrossRef]
  67. Yasuda, M.; Matsumoto, T.; Yamashita, T. Sacrificial hydrogen production over TiO2-based photocatalysts: Polyols, carboxylic acids, and saccharides. Renew. Sustain. Energy Rev. 2018, 81, 1627–1635. [Google Scholar] [CrossRef]
  68. Tan, H.; Kong, P.; Liu, M.; Gu, X.; Zheng, Z. Enhanced photocatalytic hydrogen production from aqueous-phase methanol reforming over cyano-carboxylic bifunctionally-modified carbon nitride. ChemComm. 2019, 55, 12503–12506. [Google Scholar] [CrossRef]
  69. Liu, X.; Duan, X.; Wei, W.; Wang, S.; Ni, B.-J. Photocatalytic conversion of lignocellulosic biomass to valuable products. Green Chem. 2019, 21, 4266–4289. [Google Scholar] [CrossRef]
  70. AlSalka, Y.; Al-Madanat, O.; Curti, M.; Hakki, A.; Bahnemann, D.W. Photocatalytic H2 evolution from oxalic acid: Effect of cocatalysts and carbon dioxide radical anion on the surface charge transfer mechanisms. ACS Appl. Energy Mater. 2020, 3, 6678–6691. [Google Scholar] [CrossRef]
  71. Li, Y.; Lu, G.; Li, S. Photocatalytic hydrogen generation and decomposition of oxalic acid over platinized TiO2. Appl. Catal. A Gen. 2001, 214, 179–185. [Google Scholar] [CrossRef]
  72. Li, Y.; Lu, G.; Li, S. Photocatalytic production of hydrogen in single component and mixture systems of electron donors and monitoring adsorption of donors by in situ infrared spectroscopy. Chemosphere 2003, 52, 843–850. [Google Scholar] [CrossRef]
  73. Zieliñska, B.; Borowiak-Palen, E.; Kalenczuk, R.J. Photocatalytic hydrogen generation over alkaline-earth titanates in the presence of electron donors. Int. J. Hydrogen Energy 2008, 33, 1797–1802. [Google Scholar] [CrossRef]
  74. Reuss, G.; Disteldorf, W.; Grundler, O.; Hilt, A. Ullmann’s Encycl; Society of Chemical Industry: Chichester, UK, 1985. [Google Scholar]
  75. Klankermayer, J.; Wesselbaum, S.; Beydoun, K.; Leitner, W. Cover Picture: Selective Catalytic Synthesis Using the Combination of Carbon Dioxide and Hydrogen: Catalytic Chess at the Interface of Energy and Chemistry (Angew. Chem. Int. Ed. 26/2016). Angew. Chem. Int. Ed. 2016, 55, 7267. [Google Scholar] [CrossRef]
  76. Tao, C.; Guopeng, W.; Zhaochi, F.; Gengshen, H.; Weiguang, S.; Pinliang, Y.; Can, L. In situ FT-IR study of photocatalytic decomposition of formic acid to hydrogen on Pt/TiO2 catalyst. Chin. J. Catal. 2008, 29, 105–107. [Google Scholar]
  77. Lanese, V.; Spasiano, D.; Marotta, R.; Di Somma, I.; Lisi, L.; Cimino, S.; Andreozzi, R. Hydrogen production by photoreforming of formic acid in aqueous copper/TiO2 suspensions under UV-simulated solar radiation at room temperature. Int. J. Hydrogen Energy 2013, 38, 9644–9654. [Google Scholar]
  78. Hainer, A.S.; Hodgins, J.S.; Sandre, V.; Vallieres, M.; Lanterna, A.E.; Scaiano, J.C. Photocatalytic hydrogen generation using metal-decorated TiO2: Sacrificial donors vs. true water splitting. ACS Energy Lett. 2018, 3, 542–545. [Google Scholar]
  79. Nasir, J.A.; Hafeez, M.; Arshad, M.; Ali, N.Z.; Teixeira, I.F.; McPherson, I.; Khan, M.A. Photocatalytic dehydrogenation of formic acid on CdS nanorods through Ni and Co redox mediation under mild conditions. ChemSusChem 2018, 11, 2587–2592. [Google Scholar] [CrossRef]
  80. Chiarello, G.L.; Aguirre, M.H.; Selli, E. Hydrogen production by photocatalytic steam reforming of methanol on noble metal-modified TiO2. J. Catal. 2010, 273, 182–190. [Google Scholar] [CrossRef]
  81. Song, R.; Luo, B.; Liu, M.; Geng, J.; Jing, D.; Liu, H. Synergetic coupling of photo and thermal energy for efficient hydrogen production by formic acid reforming. AIChE J. 2017, 63, 2916–2925. [Google Scholar] [CrossRef]
  82. Willner, I.; Goren, Z. Photodecomposition of formic acid by cadmium sulphide semiconductor particles. J. Chem. Soc. Chem. Commun. 1986, 2, 172–173. [Google Scholar] [CrossRef]
  83. Li, Y.; Du, J.; Peng, S.; Xie, D.; Lu, G.; Li, S. Enhancement of photocatalytic activity of cadmium sulfide for hydrogen evolution by photoetching. Int. J. Hydrogen Energy. 2008, 33, 2007–2013. [Google Scholar] [CrossRef]
  84. Li, Y.; Hu, Y.; Peng, S.; Lu, G.; Li, S. Synthesis of CdS nanorods by an ethylenediamine assisted hydrothermal method for photocatalytic hydrogen evolution. J. Phys. Chem. A C 2009, 113, 9352–9358. [Google Scholar] [CrossRef]
  85. Matsumura, M.; Hiramoto, M.; Iehara, T.; Tsubomura, H. Photocatalytic and photoelectrochemical reactions of aqueous solutions of formic acid, formaldehyde, and methanol on platinized cadmium sulfide powder and at a cadmium sulfide electrode. J. Phys. Chem. A 1984, 88, 248–250. [Google Scholar] [CrossRef]
  86. Zhang, Z.; Cao, S.-W.; Liao, Y.; Xue, C. Selective photocatalytic decomposition of formic acid over AuPd nanoparticle-decorated TiO2 nanofibers toward high-yield hydrogen production. Appl. Catal. B 2015, 162, 204–209. [Google Scholar] [CrossRef]
  87. Puangpetch, T.; Chavadej, S.; Sreethawong, T. Hydrogen production over Au-loaded mesoporous-assembled SrTiO3 nanocrystal photocatalyst: Effects of molecular structure and chemical properties of hole scavengers. Energy Convers. Manag. 2011, 52, 2256–2261. [Google Scholar] [CrossRef]
  88. Zhu, R.; Tian, F.; Yang, R.; He, J.; Zhong, J.; Chen, B. Z scheme system ZnIn2S4/RGO/BiVO4 for hydrogen generation from water splitting and simultaneous degradation of organic pollutants under visible light. Renew. Energy 2019, 139, 22–27. [Google Scholar] [CrossRef]
  89. Wu, G.; Chen, T.; Su, W.; Zhou, G.; Zong, X.; Lei, Z.; Li, C. H2 production with ultra-low CO selectivity via photocatalytic reforming of methanol on Au/TiO2 catalyst. Int. J. Hydrogen Energy 2008, 33, 1243–1251. [Google Scholar] [CrossRef]
  90. Zhang, Y.J.; Zhang, L. Photocatalytic degradation of formic acid with simultaneous production of hydrogen over Pt and Ru-loaded CdS/Al-HMS photocatalysts. Desalination 2009, 249, 1017–1021. [Google Scholar] [CrossRef]
  91. Zhang, Y.J.; Zhang, L.; Li, S. Synthesis of Al-substituted mesoporous silica coupled with CdS nanoparticles for photocatalytic generation of hydrogen. Int. J. Hydrogen Energy 2010, 35, 438–444. [Google Scholar] [CrossRef]
  92. Kuehnel, M.F.; Wakerley, D.W.; Orchard, K.L.; Reisner, E. Photocatalytic formic acid conversion on CdS nanocrystals with controllable selectivity for H2 or CO. Angew. Chem. Int. Ed. 2015, 54, 9627–9631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Yeh, H.; Lo, S.; Chen, M.; Chen, H. Hydrogen production from formic acid solution by modified TiO2 and titanate nanotubes in a two-step system under visible light irradiation. Water Sci. Technol. 2014, 69, 1676–1681. [Google Scholar] [CrossRef] [PubMed]
  94. Chen, H.-Y.; Lo, S.-L.; Lai, Y.-C.; Liou, Y.-H. Titanate nanotubes coupled with Pt nanoparticles for the inhibition of CdS photocorrosion during visible-light-driven hydrogen production from formic acid. Mater. Res. Express. 2018, 5, 095507. [Google Scholar] [CrossRef]
  95. Li, Y.; Tang, L.; Peng, S.; Li, Z.; Lu, G. Phosphate-assisted hydrothermal synthesis of hexagonal CdS for efficient photocatalytic hydrogen evolution. CrystEngComm 2012, 14, 6974–6982. [Google Scholar]
  96. Cai, Y.; Li, X.; Zhang, Y.; Wei, X.; Wang, K.; Chen, J. Highly Efficient Dehydrogenation of Formic Acid over a Palladium-Nanoparticle-Based Mott-Schottky Photocatalyst. Angew. Chem. 2013, 125, 12038–12041. [Google Scholar] [CrossRef]
  97. Kakuta, S.; Abe, T. A novel example of molecular hydrogen generation from formic acid at visible-light-responsive photocatalyst. ACS Appl. Mater. Interfaces 2009, 1, 2707–2710. [Google Scholar] [CrossRef] [PubMed]
  98. Loges, B.; Boddien, A.; Gärtner, F.; Junge, H.; Beller, M. Catalytic generation of hydrogen from formic acid and its derivatives: Useful hydrogen storage materials. Top. Catal. 2010, 53, 902–914. [Google Scholar]
  99. Park, J.Y.; Lee, I.H. Decomposition of acetic acid by advanced oxidation processes. Korean J. Chem Eng. 2009, 26, 387–391. [Google Scholar]
  100. Hisatomi, T.; Kubota, J.; Domen, K. Recent advances in semiconductors for photocatalytic and photoelectrochemical water splitting. Chem. Soc. Rev. 2014, 43, 7520–7535. [Google Scholar] [CrossRef]
  101. Kraeutler, B.; Bard, A.J. Heterogeneous photocatalytic decomposition of saturated carboxylic acids on titanium dioxide powder. Decarboxylative Route Alkanes. J. Am. Chem. Soc. 1978, 100, 5985–5992. [Google Scholar]
  102. Yoneyama, H.; Takao, Y.; Tamura, H.; Bard, A.J. Factors influencing product distribution in photocatalytic decomposition of aqueous acetic acid on platinized titania. J. Phys. Chem. A 1983, 87, 1417–1422. [Google Scholar] [CrossRef]
  103. Zheng, X.-J.; Wei, L.-F.; Zhang, Z.-H.; Jiang, Q.-J.; Wei, Y.-J.; Xie, B.; Wei, M.-B. Research on photocatalytic H2 production from acetic acid solution by Pt/TiO2 nanoparticles under UV irradiation. Int. J. Hydrogen Energy 2009, 34, 9033–9041. [Google Scholar] [CrossRef]
  104. Zheng, X.-J.; Wei, Y.-J.; Wei, L.-F.; Xie, B.; Wei, M.-B. Photocatalytic H2 production from acetic acid solution over CuO/SnO2 nanocomposites under UV irradiation. Int. J. Hydrogen Energy 2010, 35, 11709–11718. [Google Scholar] [CrossRef]
  105. Imizcoz, M.; Puga, A.V. Optimising hydrogen production via solar acetic acid photoreforming on Cu/TiO2. Catal. Sci. Technol. 2019, 9, 1098–1102. [Google Scholar] [CrossRef] [Green Version]
  106. Asal, S.; Saif, M.; Hafez, H.; Mozia, S.; Heciak, A.; Moszyński, D.; Abdel-Mottaleb, M. Photocatalytic generation of useful hydrocarbons and hydrogen from acetic acid in the presence of lanthanide modified TiO2. Int. J. Hydrogen Energy 2011, 36, 6529–6537. [Google Scholar] [CrossRef]
  107. Sun, W.; Zhang, S.; Liu, Z.; Wang, C.; Mao, Z. Studies on the enhanced photocatalytic hydrogen evolution over Pt/PEG-modified TiO2 photocatalysts. Int. J. Hydrogen Energy 2008, 33, 1112–1117. [Google Scholar] [CrossRef]
  108. Sreethawong, T.; Puangpetch, T.; Chavadej, S.; Yoshikawa, S. Quantifying influence of operational parameters on photocatalytic H2 evolution over Pt-loaded nanocrystalline mesoporous TiO2 prepared by single-step sol–gel process with surfactant template. J. Power Sources 2007, 165, 861–869. [Google Scholar] [CrossRef]
  109. Hamid, S.; Ivanova, I.; Jeon, T.H.; Dillert, R.; Choi, W.; Bahnemann, D.W. Photocatalytic conversion of acetate into molecular hydrogen and hydrocarbons over Pt/TiO2: pH dependent formation of Kolbe and Hofer-Moest products. J. Catal. 2017, 349, 128–135. [Google Scholar] [CrossRef]
  110. Sato, S. Photo-Kolbe reaction at gas-solid interfaces. J. Phys. Chem. A 1983, 87, 3531–3537. [Google Scholar] [CrossRef]
  111. Sakata, T.; Kawai, T.; Hashimoto, K. Heterogeneous photocatalytic reactions of organic acids and water. New React. Paths Besides Photo-Kolbe React. J. Phys. Chem. A 1984, 88, 2344–2350. [Google Scholar] [CrossRef]
  112. Heciak, A.; Morawski, A.W.; Grzmil, B.; Mozia, S. Cu-modified TiO2 photocatalysts for decomposition of acetic acid with simultaneous formation of C1–C3 hydrocarbons and hydrogen. Appl. Catal. B 2013, 140, 108–114. [Google Scholar] [CrossRef]
  113. Mozia, S.; Heciak, A.; Morawski, A.W. Preparation of Fe-modified photocatalysts and their application for generation of useful hydrocarbons during photocatalytic decomposition of acetic acid. J. Photochem. Photobiol. 2010, 216, 275–282. [Google Scholar] [CrossRef]
  114. Mozia, S.; Heciak, A.; Morawski, A.W. Photocatalytic acetic acid decomposition leading to the production of hydrocarbons and hydrogen on Fe-modified TiO2. Catal. Today 2011, 161, 189–195. [Google Scholar] [CrossRef]
  115. Sclafani, A.; Palmisano, L.; Schiavello, M.; Augugliaro, V.; Coluccia, S.; Marchese, L. The photodecomposition of ethanoic acid adsorbed over semiconductor and insulator oxides. I. Pure Oxides. New J. Chem. (1987) 1988, 12, 129–135. [Google Scholar]
  116. Abdel-Rahman, M.A.; Tashiro, Y.; Sonomoto, K. Lactic acid production from lignocellulose-derived sugars using lactic acid bacteria: Overview and limits. J. Biotechnol. 2011, 156, 286–301. [Google Scholar] [CrossRef] [PubMed]
  117. Joshi, D.; Singhvi, M.; Khire, J.; Gokhale, D. Strain improvement of Lactobacillus lactis for D-lactic acid production. Biotechnol. Lett. 2010, 32, 517–520. [Google Scholar] [CrossRef]
  118. Zhang, W.; Wang, Y.; Wang, Z.; Zhong, Z.; Xu, R. Highly efficient and noble metal-free NiS/CdS photocatalysts for H 2 evolution from lactic acid sacrificial solution under visible light. ChemComm 2010, 46, 7631–7633. [Google Scholar] [CrossRef]
  119. Harada, H.; Sakata, T.; Ueda, T. Effect of semiconductor on photocatalytic decomposition of lactic acid. J. Am. Chem. Soc. 1985, 107, 1773–1774. [Google Scholar] [CrossRef]
  120. Harada, H.; Sakata, T.; Ueda, T. Semiconductor effect on the selective photocatalytic reaction of. alpha.-hydroxycarboxylic acids. J. Phys. Chem. A 1989, 93, 1542–1548. [Google Scholar]
  121. Li, Q.; Guo, B.; Yu, J.; Ran, J.; Zhang, B.; Yan, H.; Gong, J.R. Highly efficient visible-light-driven photocatalytic hydrogen production of CdS-cluster-decorated graphene nanosheets. J. Am. Chem. Soc. 2011, 133, 10878–10884. [Google Scholar] [CrossRef]
  122. Zong, X.; Han, J.; Ma, G.; Yan, H.; Wu, G.; Li, C. Photocatalytic H2 evolution on CdS loaded with WS2 as cocatalyst under visible light irradiation. J. Phys. Chem. A C 2011, 115, 12202–12208. [Google Scholar] [CrossRef]
  123. Hu, Z.; Jimmy, C.Y. Pt 3 Co-loaded CdS and TiO2 for photocatalytic hydrogen evolution from water. J. Mater. Chem. 2013, 1, 12221–12228. [Google Scholar] [CrossRef]
  124. Li, C.; Wang, H.; Naghadeh, S.B.; Zhang, J.Z.; Fang, P. Visible light driven hydrogen evolution by photocatalytic reforming of lignin and lactic acid using one-dimensional NiS/CdS nanostructures. Appl. Catal. B 2018, 227, 229–239. [Google Scholar] [CrossRef]
  125. Zong, X.; Yan, H.; Wu, G.; Ma, G.; Wen, F.; Wang, L.; Li, C. Enhancement of photocatalytic H2 evolution on CdS by loading MoS2 as cocatalyst under visible light irradiation. J. Am. Chem. Soc. 2008, 130, 7176–7177. [Google Scholar] [CrossRef]
  126. Chen, G.; Li, D.; Li, F.; Fan, Y.; Zhao, H.; Luo, Y.; Yu, R.; Meng, Q. Ball-milling combined calcination synthesis of MoS2/CdS photocatalysts for high photocatalytic H2 evolution activity under visible light irradiation. Appl. Catal. A Gen. 2012, 443, 138–144. [Google Scholar] [CrossRef]
  127. Lang, D.; Shen, T.; Xiang, Q. Roles of MoS2 and graphene as cocatalysts in the enhanced visible-light photocatalytic H2 production activity of multiarmed CdS nanorods. ChemCatChem 2015, 7, 943–951. [Google Scholar] [CrossRef]
  128. Li, Y.; Wang, H.; Peng, S. Tunable photodeposition of MoS2 onto a composite of reduced graphene oxide and CdS for synergic photocatalytic hydrogen generation. J. Phys. Chem. A C 2014, 118, 19842–19848. [Google Scholar] [CrossRef]
  129. Chang, K.; Mei, Z.; Wang, T.; Kang, Q.; Ouyang, S.; Ye, J. MoS2/graphene cocatalyst for efficient photocatalytic H2 evolution under visible light irradiation. ACS Nano 2014, 8, 7078–7087. [Google Scholar] [CrossRef]
  130. Shaislamov, U.; Yang, B.L. CdS-sensitized single-crystalline TiO2 nanorods and polycrystalline nanotubes for solar hydrogen generation. J. Mater. Res. 2013, 28, 418–423. [Google Scholar] [CrossRef]
  131. Hou, Y.; Laursen, A.B.; Zhang, J.; Zhang, G.; Zhu, Y.; Wang, X.; Dahl, S.; Chorkendorff, I. Layered nanojunctions for hydrogen-evolution catalysis. Angew. Chem. 2013, 125, 3709–3713. [Google Scholar] [CrossRef]
  132. Pham, M.-H.; Dinh, C.-T.; Vuong, G.-T.; Ta, N.-D.; Do, T.-O. Visible light induced hydrogen generation using a hollow photocatalyst with two cocatalysts separated on two surface sides. Phys. Chem. Chem. Phys. 2014, 16, 5937–5941. [Google Scholar] [CrossRef] [PubMed]
  133. Tu, H.; Xu, L.; Mou, F.; Guan, J. Single crystalline tantalum oxychloride microcubes: Controllable synthesis, formation mechanism and enhanced photocatalytic hydrogen production activity. ChemComm 2015, 51, 12455–12458. [Google Scholar] [CrossRef]
  134. Cao, S.; Chen, Y.; Wang, C.-J.; Lv, X.-J.; Fu, W.-F. Spectacular photocatalytic hydrogen evolution using metal-phosphide/CdS hybrid catalysts under sunlight irradiation. ChemComm 2015, 51, 8708–8711. [Google Scholar] [CrossRef] [Green Version]
  135. Cao, S.; Chen, Y.; Kang, L.; Lin, Z.; Fu, W.-F. Enhanced photocatalytic H2-evolution by immobilizing CdS nanocrystals on ultrathin Co0.85 Se/RGO–PEI nanosheets. J. Mater. Chem. A 2015, 3, 18711–18717. [Google Scholar] [CrossRef]
  136. Pan, Y.X.; Zhuang, H.; Hong, J.; Fang, Z.; Liu, H.; Liu, B.; Huang, Y.; Xu, R. Cadmium Sulfide Quantum Dots Supported on Gallium and Indium Oxide for Visible-Light-Driven Hydrogen Evolution from Water. ChemSusChem 2014, 7, 2537–2544. [Google Scholar] [CrossRef]
  137. Wang, X.; Yu, H.; Yang, L.; Shao, L.; Xu, L. A highly efficient and noble metal-free photocatalytic system using NixB/CdS as photocatalyst for visible light H2 production from aqueous solution. Catal. Commun. 2015, 67, 45–48. [Google Scholar] [CrossRef]
  138. Hu, J.; Liu, A.; Jin, H.; Ma, D.; Yin, D.; Ling, P.; Wang, S.; Lin, Z.; Wang, J. A versatile strategy for shish-kebab-like multi-heterostructured chalcogenides and enhanced photocatalytic hydrogen evolution. J. Am. Chem. Soc. 2015, 137, 11004–11010. [Google Scholar] [CrossRef]
  139. Shen, L.; Luo, M.; Liu, Y.; Liang, R.; Jing, F.; Wu, L. Noble-metal-free MoS2 co-catalyst decorated UiO-66/CdS hybrids for efficient photocatalytic H2 production. Appl. Catal. B Environ. 2015, 166, 445–453. [Google Scholar] [CrossRef]
  140. Liu, K.; Litke, A.; Su, Y.; Van Campenhout, B.G.; Pidko, E.A.; Hensen, E.J. Photocatalytic decarboxylation of lactic acid by Pt/TiO2. ChemComm 2016, 52, 11634–11637. [Google Scholar] [CrossRef]
  141. Shiragami, T.; Tomo, T.; Tsumagari, H.; Yuki, R.; Yamashita, T.; Yasuda, M. Pentose acting as a sacrificial multielectron source in photocatalytic hydrogen evolution from water by Pt-doped TiO2. Chem. Lett. 2012, 41, 29–31. [Google Scholar] [CrossRef]
  142. Li, Y.; Lu, G.; Li, S. Photocatalytic transformation of rhodamine B and its effect on hydrogen evolution over Pt/TiO2 in the presence of electron donors. J. Photochem. Photobiol. 2002, 152, 219–228. [Google Scholar] [CrossRef]
  143. Wei, L.F.; Zheng, X.J.; Zhang, Z.H.; Wei, Y.J.; Xie, B.; Wei, M.B.; Sun, X.L. A systematic study of photocatalytic H2 production from propionic acid solution over Pt/TiO2 photocatalyst. Int. J. Energy Res. 2012, 36, 75–86. [Google Scholar] [CrossRef]
  144. Scandura, G.; Rodríguez, J.; Palmisano, G. Hydrogen and propane production from butyric acid photoreforming over Pt-TiO2. Front. Chem. 2019, 7, 563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Mechanism involved in photoreforming, photocatalytic water splitting and organic photo-oxidation [15].
Figure 1. Mechanism involved in photoreforming, photocatalytic water splitting and organic photo-oxidation [15].
Photochem 02 00040 g001
Figure 2. Presentation of band gaps and positions of various photocatalysts as well as selected redox potentials [27].
Figure 2. Presentation of band gaps and positions of various photocatalysts as well as selected redox potentials [27].
Photochem 02 00040 g002
Figure 3. Illustration of mechanism involved in photoreforming of (a) methanol. Reprinted with permission from Elsevier Ref. [44] 2013, Paramasivan et al, (b) ethanol. Reprinted with permission from American chemical society Ref. [49], 2019, Xin Zhang et al, (c) glycerol. Reprinted with permission from Elsevier Ref. [54], 2020, Muhammad et al and (d) cellulose. Reprinted with permission from American chemical society Ref. [58], 2021, Van-Can et al.
Figure 3. Illustration of mechanism involved in photoreforming of (a) methanol. Reprinted with permission from Elsevier Ref. [44] 2013, Paramasivan et al, (b) ethanol. Reprinted with permission from American chemical society Ref. [49], 2019, Xin Zhang et al, (c) glycerol. Reprinted with permission from Elsevier Ref. [54], 2020, Muhammad et al and (d) cellulose. Reprinted with permission from American chemical society Ref. [58], 2021, Van-Can et al.
Photochem 02 00040 g003
Figure 4. (a) Photoreforming of FA on NiCo/CdS; (b) Effect of Ni and Co on the production of H2; (c) Long-term H2 production using NiCo/CdS. Reprinted with permission from Elsevier Ref. [79], 2018, M. Abdullah Khan et al.
Figure 4. (a) Photoreforming of FA on NiCo/CdS; (b) Effect of Ni and Co on the production of H2; (c) Long-term H2 production using NiCo/CdS. Reprinted with permission from Elsevier Ref. [79], 2018, M. Abdullah Khan et al.
Photochem 02 00040 g004
Figure 5. (a,b) TEM images of copper-loaded TiO2; (c,d) Mechanism involved in the production of hydrogen using acetic acid as a sacrificial agent; (e) H2 and CH4 production for 12 h on Cu/TiO2; (f) Overall mechanism that occurs in the photoreforming of acetic acid on Cu/TiO2 [105].
Figure 5. (a,b) TEM images of copper-loaded TiO2; (c,d) Mechanism involved in the production of hydrogen using acetic acid as a sacrificial agent; (e) H2 and CH4 production for 12 h on Cu/TiO2; (f) Overall mechanism that occurs in the photoreforming of acetic acid on Cu/TiO2 [105].
Photochem 02 00040 g005
Figure 6. (a) Comparison of the quantum efficiency of hydrogen production on NiS/CdS using lignin and LA as a sacrificial agent; (b) Long-term stability of the production of H2 on a NiS/CdS photocatalyst; (c) Mechanism involved in the lignin and LA reformation of NiS/CdS photocatalysts. Reprinted with permission from Elsevier Ref. [124], 2018, Chunhe Li et al.
Figure 6. (a) Comparison of the quantum efficiency of hydrogen production on NiS/CdS using lignin and LA as a sacrificial agent; (b) Long-term stability of the production of H2 on a NiS/CdS photocatalyst; (c) Mechanism involved in the lignin and LA reformation of NiS/CdS photocatalysts. Reprinted with permission from Elsevier Ref. [124], 2018, Chunhe Li et al.
Photochem 02 00040 g006
Figure 7. A proposed approach for the selective photoreforming of various acids into H2 and valuable organic compounds.
Figure 7. A proposed approach for the selective photoreforming of various acids into H2 and valuable organic compounds.
Photochem 02 00040 g007
Table 1. Illustration of various catalysts, light sources and cocatalysts used in the photoreforming of FA (HCOOH) and the rate of hydrogen production.
Table 1. Illustration of various catalysts, light sources and cocatalysts used in the photoreforming of FA (HCOOH) and the rate of hydrogen production.
ConcentrationLight Source, Wavelength (nm)PhotocatalystCo-CatalystTime Course (h)Rate of Production (μmolg−1 h−1)Reference
H2CO2CO
aq. 1.3 mMsolar0.5%-Pt/TiO2Pt201275--[17]
aq. 0.01 MHg0.5%-Pt/TiO2Pt51150--[72]
H2O/HCOOH (10:1), vHg, 350–450 FP-0.5%-Pt/TiO2Pt-5400410080[80]
10 vol%, (35 °C)UV light, 200–800 1%-Pt/TiO2Pt (photo)861.5--[81]
10 vol%, (90 °C)UV light1%-Pt/TiO2Pt (thermal)8119.3--[81]
10 vol%, (photo + 90 °C)UV light1%-Pt/TiO2Pt (photothermal)8499.8--[81]
0.5 M HCO2/H2OUV lightCdS-12420 (μL) -[82]
0.5 M DCO2/H2OUV lightCdS-12110 (μL) -[82]
aq, 5 mL of 88% HCOOHHg, >3001.5%-Pt/CdS (photoetching)Pt101128--[83]
aq, 5 mL of 88 wt% HCOOHHgCdS-1079--[83]
aq, 5 mL of 88 wt% HCOOHHg, 4200.05%-Pt/CdSPt104460--[84]
aq, 5 mL of 88 wt% HCOOHHgCdS-10219--[84]
aq, 2 MHg, 400Pt/CdSPt2038538577[85]
aq, 20 mLXe, >420Co-Ni/CdS-NRCo, Ni1832,600--[79]
aq, 20 mLXeCo/CdS-NRCo1214,200--[79]
aq, 20 mLXeNi/CdS-NRNi1222,800--[79]
aq, 20 mLXeCdS-NR-1213,400--[79]
aq, 2.7 MSolar, 420–7800.75%-Au, 0.25%-Pd/TiO2Au, Pd1017,700--[86]
aq, 200 mL of 2.5 vol% CHOOHHg1%-Au-loaded mesoporous assembled SrTiO3Au5647--[87]
aq, 150 mL H2O + HCOOHXeCdS-ZnSCd:Zn (0.8:0.2)41263--[88]
aq, 150 mL H2O + HCOOHXe5%-Ru/CdS-ZnSRu, Cd:Zn (0.8:0.2)45800--[88]
aq, 200 mL, 0.05 MXe, <355Au/TiO2Au4452--[89]
aq, (4:1 in volume) H2O:HCOOHXe, <4200.34%-Pt/2.5%-CdS/Al-HMSPt61705--[90]
aq, (9:1 in volume) H2O:HCOOHXe, <4200.99%Ru/21%-CdS/Al-HMSRu62753--[91]
aq, 1 MHg, 399SrTiO3:
TiO2
SrTiO35280--[73]
aq, 2.5 MUV light, >420QD-MPA-16852,100- [92]
aq, 2.5 MUV lightQD-MPA/CoCl2CoCl2168116,000- [92]
aq, 20 vol% HCOOHUV light0.01%-Pt/CdS/TNTPt342,600--[93]
aq, 10 vol% HCOOHUV light, 2541%-Pt(P)/CdS/TNTPt33300--[94]
5 mL HCOOH in 100 mL H2O + 180 °CUV light, ≥4200.0.25%-Pt/CdS-QDPt3012,200--[95]
aq, 10 mL, 1 M HCOOHUV light, ≥420Pd/C3N4Pd653,400--[96]
aq, 1 M HCOOHUV lightCu/TiO2 (anatase)Cu55000--[77]
aq, 2.5 M HCOOH, pH 5Halogen, 4200.5%-Pt/Cu2OPt40155158-[97]
aq, 2.5 M HCOOH, pH 5HalogenCu2O-406564-[97]
1 mL DMF, 5 mL 5HCO2H ·2NEt3Xe[Fe3 (CO)12] + PPh3,2,2′:6′,2″-terpyridine 32700-trace[98]
Table 2. Illustration of various catalysts, light sources and cocatalysts used in the photoreforming of acetic acid (CH3COOH) and the rate of hydrogen production.
Table 2. Illustration of various catalysts, light sources and cocatalysts used in the photoreforming of acetic acid (CH3COOH) and the rate of hydrogen production.
ConcentrationLight Source, Wavelength (nm)PhotocatalystCo-CatalystTime Course (h)Rate of Production (μmolg−1 h−1)Reference
H2CO2CH4C2H6
aq, 1 M CH3COOHUV lightLn3+(0.02%-Eu)/TiO2Ln3+ (Eu3+)279881084[106]
aq, 1 M CH3COOHUV lightLn3+(0.05%-Sm)/TiO2Ln3+ (Sm3+)2731311243[106]
aq, 0.87 mM CH3COOHXe0.5%-Pt/P25TiO2Pt20278---[17]
aq, H2O/CH3COOH (5:1)Hg0.5%Pt/PEG-TiO2Pt61000---[107]
aq, H2O/CH3COOH (10:1)Hg0.6%-Pt/TiO2 (mesoporous)Pt5390---[108]
aq, 50 mL (0.5 M) CH3COOH and H2O, pH 2Xe1%-Pt/TiO2Pt1522 (μmol/h)65 (μmol/h)35 (μmol/h)2 (μmol/h)[109]
aq, 450 mL CH3COOH (4.35 g/L), pH 1.0Hg1%Pt-TiO2Pt428,478---[103]
l, 15 mL solution CH3COOH/Na HACXe-Hg1–5%-Pt/TiO2(anatase)Pt -1600--[101]
l, H2O/CH3COOH (1:9)Xe-Hg1–5%-Pt/TiO2(anatase)Pt -4060--[101]
l, H2O/CH3COOH (1:1)Xe-Hg1–5%-Pt/TiO2(anatase)Pt -15631322120[101]
aq, AcOH/Na [AcO] (4:0.6 M), pH 3.9Hg-conc.3%-Pt/TiO2 (anatase)Pt 2465434[102]
aq, AcOH/Na [AcO] (4:0.6 M), pH 3.9Hg-conc.3%-Pt/TiO2 (rutile)Pt 911101[102]
v, CH3COOH (11 torr)Hg2%-Pt/TiO2(anatase)Pt3461325443[110]
v, H2O/CH3COOH (24:11 torr)Hg2%-Pt/TiO2(anatase)Pt3180453104180[110]
l, CH3COOH (1 mL)Hg2%-Pt/TiO2(anatase)Pt34629016312[110]
l, H2O/CH3COOH (6:1), pH 2.1Hg, 3667%-Pt/TiO2 (rutile)Pt 77-397-[111]
l, H2O/CH3COOH (6:1), pH 8.8Hg7%-Pt/TiO2 (rutile)Pt 367-2-[111]
aq, Na (AcO) (1.7% w/v), pH 7.4Hg7%-Pt/TiO2 (anatase)Pt 165270.24-[111]
aq, 1 M CH3COOH, pH 2.6UV light10%-Cu/TiO2Cu514464059066[112]
aq, 1 M CH3COOHUV light, 366P25TiO2 272141504[113]
aq, 1 M CH3COOHUV light, 36610%-Fe/TiO2Fe277102934[113]
aq, 1 M CH3COOHHg20-%Fe/TiO2Fe5725726015[114]
v, CH3COOH (665 Pa)Hg-Xe, >420TiO2- -0.3700.018-[115]
v, CH3COOH (665 Pa)Hg-XeZnO2- -0.0980.034-[115]
v, CH3COOH (665 Pa)Hg-XeMgO- -0.9230.179-[115]
v, CH3COOH (665 Pa)Hg-XeSiO2- -0.3360.168-[115]
v, CH3COOH (665 Pa)Hg-XeWO3- -0.0260.03-[115]
v, CH3COOH (665 Pa)Hg-Xeγ-Al2O3- -0.3360.086-[115]
Table 3. Illustration of various catalysts, light sources and co-catalysts used in photoreforming of LA (CH3CH(OH)COOH) and the rate of hydrogen production.
Table 3. Illustration of various catalysts, light sources and co-catalysts used in photoreforming of LA (CH3CH(OH)COOH) and the rate of hydrogen production.
ConcentrationLight Source, Wavelength (nm)PhotocatalystCo-CatalystTime Course (h)Rate of Production (μmolg−1 h−1)Reference
H2CO2
l, CH3CH(OH)COOH/H2O (1:10) pH 2Xe, 360–5205%-Pt/TiO2 (rutile)Pt410081192[120]
l, CH3CH(OH)COOH/H2O (1:10) pH 2Xe5%-Pt/CdSPt4100013[120]
l, CH3CH(OH)COOH/H2O (1:9)Xe, 4201%-Pt/CdSPt 8170-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-(Pt3Co)/CdSPt, Co 15,860-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-(Pt2.3-Co)/CdSPt, Co 13,010-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-(PtCo)/CdSPt, Co 7050-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-Co/CdSCo 1070-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-(Pt3Au)/CdSPt, Au 14,900-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-(Pt3Ni)/CdSPt, Ni 12,810-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-(Pt3Cu)/CdSPt, Cu 3890-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-Pt/P25TiO2Pt 690-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-(Pt3Co)/P25TiO2Pt, Co 1040-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-(Pt3Au)/P25TiO2Pt, Au 890-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-(Pt3Ni)/P25TiO2Pt, Ni 820-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe, 4201%-(Pt3Cu)/P25TiO2Pt, Cu 460-[123]
l, CH3CH(OH)COOH/H2O (1:9)Xe0.2%-MoS2/CdSMoS2 5400-[125]
l, CH3CH(OH)COOH/H2O (1:9)Xe0.2%-Pt/CdSPt54400-[125]
l, CH3CH(OH)COOH/H2O (1:9)Xe0.2%-Ru/CdSRu53650-[125]
l, CH3CH(OH)COOH/H2O (1:9)Xe0.2%-Rh/CdSRh52500-[125]
l, CH3CH(OH)COOH/H2O (1:9)Xe0.2%-Pd/CdSPd51800-[125]
l, CH3CH(OH)COOH/H2O (1:9)Xe0.2%-Au/CdSAu5400-[125]
l, CH3CH(OH)COOH/H2O (1:9)Xe, 420CdS-5150-[125]
l, CH3CH(OH)COOH/H2O (1:9)Xe, 4201%-WS2/CdSWS254000-[122]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-Pt/CdSPt53550-[122]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-Ru/CdSRu52930-[122]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-Rh/CdSRh52070-[122]
l, CH3CH(OH)COOH/H2O (1:9)Xe1%-Au/CdSAu5455-[122]
aq, H2O/CH3CH(OH)COOH (10:1)Xe0.9%-MoS2/CdSMoS2513,151-[126]
aq, H2O/CH3CH(OH)COOH (10:1)Xe, 4200.2%-Pt/CdSPt54880-[126]
aq, H2O/CH3CH(OH)COOH (9:1)Xe, 4202.5%-MoS2-RGO/CdSMoS2-RGO 621.3-[127]
aq, H2O/CH3CH(OH)COOH (9:1)Xe2.5%-MoS2/CdSMoS2 551.3 [127]
aq, H2O/CH3CH(OH)COOH (9:1)Xe0.25%-Pt/CdSPt 450-[127]
aq, H2O/CH3CH(OH)COOH (9:1)Xe, 4201.5%-RGO/CdS/1.5%-MoS2RGO, MoS251980-[128]
aq, H2O/CH3CH(OH)COOH (4:1)Xe, 4202%-(MoS2/RGO)/CdSMoS2, RGO59000-[129]
aq, H2O/CH3CH(OH)COOH (7:3)Xe, 4201.2%-NiS/CdSNiS 7267-[118]
aq, H2O/CH3CH(OH)COOH (7:3)Xe, 420CdS- 210 [118]
aq, H2O/CH3CH(OH)COOH (7:3)Xe1%-Pt/CdSPt 1333-[118]
aq, H2O/CH3CH(OH)COOH (7:3)Xe1.2%-CoS/CdSCoS 1000-[118]
aq, H2O/CH3CH(OH)COOH (9:1)Xe, 420TiO2-1.2%-Pt/CdSPt, TiO2 NPs 14,750-[130]
aq, H2O/CH3CH(OH)COOH (9:1)XeP25TiO2-2%Pt/CdsPt, TiO2 NPs 13,750-[130]
aq, H2O/CH3CH(OH)COOH (9:1)Xe, 4202%-Pt/P25TiO2 3875-[130]
aq, H2O/CH3CH(OH)COOH (9:1)Xe, 4200.5%-Pt/0.5%-RGO/CdSPt, RGO 19,000-[121]
aq, H2O/CH3CH(OH)COOH (9:1)Xe0.5%-Pt/1%-RGO/CdSPt, RGO 56,000-[121]
aq, H2O/CH3CH(OH)COOH (9:1)Xe0.5%-Pt/2.5%-RGO/CdSPt, RGO 27,500-[121]
aq, H2O/CH3CH(OH)COOH (9:1)Xe, 4200.2%-MoS2/g-C3N4(mesoporous)MoS241375-[131]
aq, H2O/CH3CH(OH)COOH (9:1)Xe2%-Pt/g-C3N4(mesoporous)Pt41000-[131]
aq, H2O/CH3CH(OH)COOH (9:1)Xe0.5%-WS2/g-C3N4(mesoporous)WS24340-[131]
aq, H2O/CH3CH(OH)COOH (9:1)Xe, 420Fe2O3-TiO2-PtOxFe2O3, PtOx51100-[132]
aq, H2O/CH3CH(OH)COOH (7:3)Xe, 2805%-Pt/TaO2.18Cl0.64Pt81500-[133]
aq, H2O/CH3CH(OH)COOH (19:1), pH 3LED, 420-780CoP/CdSCoP10251,500-[134]
aq, H2O/CH3CH(OH)COOH (19:1), pH 3LEDNi2P/CdSNi2P10143,600-[134]
aq, H2O/CH3CH(OH)COOH (19:1), pH 3LEDCu3P/CdSCu3P1077,600-[134]
aq, H2O/CH3CH(OH)COOH (19:1), pH 3LEDPt/CdSPt1077,300-[134]
aq, H2O/CH3CH(OH)COOH (20:3)LED, 420Co0.85Se/RGO–PEI nanosheets/CdSCo0.85Se/RGO–PEI nanosheets1017,600-[135]
aq, H2O/CH3CH(OH)COOH (20:3)LED0.1%-Pt/CdSPt1018,600-[135]
aq, H2O/CH3CH(OH)COOH (4:1)Xe, 4209%-CdS/1%-Pt/In2O3Pt>109384-[136]
aq, H2O/CH3CH(OH)COOH (4:1)Xe9%-CdS/1%-Pt/Ga2O3Pt>109053-[136]
aq, H2O/CH3CH(OH)COOH (4:1)Xe, 4209%-CdS/1%-Pt/P25TiO2Pt>105482-[136]
aq, H2O/CH3CH(OH)COOH (9:1)Xe, 4200.8%-NiBx/CdSNiBx104800-[137]
aq, H2O/CH3CH(OH)COOH (9:1)Xe, 4209%-NiS-37%-CdS/TeNiS, Te12317-[138]
aq, H2O/CH3CH(OH)COOH (9:1)Xe, 42011%-Pt/11%-Pd/31%-CdS/TePt, Pd, Te12236-[138]
aq, H2O/CH3CH(OH)COOH (9:1)Xe1.5%-MoS2/UiO-66//CdSMoS2432,500-[139]
aq, CH3CH(OH)COOH (0.1 M)Hg-Xe, 420TiO2 30693[140]
aq, CH3CH(OH)COOH (0.1 M)Hg-Xe1%-Pt/TiO2Pt313,89013,500[140]
aq, CH3CH(OH)COOH (0.1 M)Hg-Xe1%-Au/TiO2Au366676600[140]
aq, CH3CH(OH)COOH (0.1 M)Hg-Xe1%-Pd/TiO2Pd383008700[140]
aq, CH3CH(OH)COOH (0.1 M)Hg-Xe1%-Ru/TiO2Ru311,50011,120[140]
Table 4. Various catalysts, light sources and co-catalyst used in photoreforming of other carboxylic acids and the rate of hydrogen production.
Table 4. Various catalysts, light sources and co-catalyst used in photoreforming of other carboxylic acids and the rate of hydrogen production.
Acid ConcentrationLight Source, Wavelength (nm)PhotocatalystCo-CatalystTime Course (h)Rate of Production (μmolg−1 h−1)Reference
H2CO2C2H6
aq, 0.01 M HOOCCOOH (Oxalic acid), pH 2Hg0.5%-Pt/P25TiO2Pt52850--[72]
aq, 0.049 M
HOOCCOOH (Oxalic acid)
Hg0.5%-Pt/P25TiO2Pt51160--[71]
aq, 0.001 M
HOOCCOOH (Oxalic acid)
Hg0.3%-Pt/P25TiO2Pt5.53750--[142]
aq, 1.3 g/L HOCH2(CHOH)4COOH (Gluconic acid) Hg2%-Pt/TiO2Pt30675296-[141]
aq, 0.51 g/L H2O/HOCH2COOH (Glycolic acid) Hg2%-Pt/TiO2Pt46284109-[141]
H2O/HOCH2COOH (Glycolic acid) (1:10 v/v)Xe, 360–5205%-Pt/TiO2(rutile)Pt510576-[120]
H2O/HOCH2COOH (Glycolic acid) (1:10)Xe5%-Pt/CdSPt53923-[120]
aq, C18H36O2 or CH3(CH2)16COOH (Stearic acid) (1.7%w/v)Xe5%-Pt/TiO2Pt24290-[60]
aq, H2O/CH3CH2COOH (Propionic acid) (6:1)Hg, 3207%-Pt/TiO2(rutile)Pt 51-490[111]
aq, H2O/CH3CH2CH2COOH (L-Propionic acid) (6:1)Hg, 3207%-Pt/TiO2(rutile)Pt 111-332[111]
aq, H2O/CH3(CH2)3COOH (6:1)Hg7%-Pt/TiO2(rutile)Pt 174-339[111]
l, 6.66 g/L CH3CH2COOH (Butyric acid)Hg, 4201%-Pt/TiO2Pt>166800 (μmol h−1)--[143]
aq, 5 mM CH3CH2CH2COOH (Valeric acid) (400 °C)LED, 360–370Pt/TiO2-NTsPt8387.5362.5-[144]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Samage, A.; Gupta, P.; Halakarni, M.A.; Nataraj, S.K.; Sinhamahapatra, A. Progress in the Photoreforming of Carboxylic Acids for Hydrogen Production. Photochem 2022, 2, 580-608. https://doi.org/10.3390/photochem2030040

AMA Style

Samage A, Gupta P, Halakarni MA, Nataraj SK, Sinhamahapatra A. Progress in the Photoreforming of Carboxylic Acids for Hydrogen Production. Photochem. 2022; 2(3):580-608. https://doi.org/10.3390/photochem2030040

Chicago/Turabian Style

Samage, Anita, Pooja Gupta, Mahaveer A. Halakarni, Sanna Kotrappanavar Nataraj, and Apurba Sinhamahapatra. 2022. "Progress in the Photoreforming of Carboxylic Acids for Hydrogen Production" Photochem 2, no. 3: 580-608. https://doi.org/10.3390/photochem2030040

Article Metrics

Back to TopTop