Next Article in Journal
Effects of Lanthanide Doping on the Catalytic Activity and Hydrothermal Stability of Cu-SAPO-18 for the Catalytic Removal of NOx (NH3-SCR) from Diesel Engines
Next Article in Special Issue
One-Step Synthesis of b-N-TiO2/C Nanocomposites with High Visible Light Photocatalytic Activity to Degrade Microcystis aeruginosa
Previous Article in Journal
Substitution of Co with Ni in Co/Al2O3 Catalysts for Fischer–Tropsch Synthesis
Previous Article in Special Issue
Pt Nanowire-Anchored Dodecahedral Ag3PO4{110} Constructed for Significant Enhancement of Photocatalytic Activity and Anti-Photocorrosion Properties: Spatial Separation of Charge Carriers and PhotogeneratedElectron Utilization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalytic Reforming for Hydrogen Evolution: A Review

Key Laboratory of Marine Chemistry Theory and Technology (Ministry of Education), College of Chemistry & Chemical Engineering, Ocean University of China, Qingdao 266100, Shandong, China
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(3), 335; https://doi.org/10.3390/catal10030335
Submission received: 27 February 2020 / Revised: 12 March 2020 / Accepted: 16 March 2020 / Published: 17 March 2020
(This article belongs to the Special Issue Progression in Photocatalytic Materials for Efficient Performance)

Abstract

:
Hydrogen is considered to be an ideal energy carrier to achieve low-carbon economy and sustainable energy supply. Production of hydrogen by catalytic reforming of organic compounds is one of the most important commercial processes. With the rapid development of photocatalysis in recent years, the applications of photocatalysis have been extended to the area of reforming hydrogen evolution. This research area has attracted extensive attention and exhibited potential for wide application in practice. Photocatalytic reforming for hydrogen evolution is a sustainable process to convert the solar energy stored in hydrogen into chemical energy. This review comprehensively summarized the reported works in relevant areas, categorized by the reforming precursor (organic compound) such as methanol, ethanol and biomass. Mechanisms and characteristics for each category were deeply discussed. In addition, recommendations for future work were suggested.

Graphical Abstract

1. Introduction

Approximately 3 × 1024 J of solar energy is irradiated on the surface of earth annually, which is over 104 -fold the global energy consumption annually [1]. Effective restoration or transfer of solar energy is of great significance to human society. The photocatalytic process, one of the most promising techniques in solar energy conversion, has increasingly attracted attention since 1972 [2]. Applications of this advanced technology have been implemented in various areas, such as water/air purification, water splitting, chemical conversions, disinfection, reduction of CO2 and N2 fixation, etc. [3,4,5,6]. Hydrogen can be evolved through either photocatalytic water splitting or reforming [7,8]. Through these processes, solar energy can be converted to chemical energy, and stored in hydrogen. Considering the depletion of fossil fuel reserves and increasing environmental problems, hydrogen is an attractive alternative energy source with the following merits [9,10,11,12]:
(i)
Hydrogen atom is the most abundant element in the universe;
(ii)
The heating value of hydrogen (141.8 × 106 J/kg at 298 K) is much higher than that of most fuels (e.g., gasoline: 44 × 106 J/kg at 298 K);
(iii)
Water (H2O) is the only exhaust product when hydrogen is converted into energy, which is more environmentally-friendly compared to fossil fuels.
Hydrogen is currently derived from non-renewable resources (such as natural gas and petroleum) through thermal catalytic reforming in industry. Compared with the industrial process for hydrogen evolution, photocatalysis can be realized in mild conditions, and induced by light energy instead of thermal energy. As a promising process, a large amount of work have been reported on photocatalytic reforming for hydrogen evolution. In this review, typical photocatalytic reforming processes were categorized based on the precursor, and comprehensively discussed. Future development of photocatalytic reforming for hydrogen production has been proposed at the end.

2. Industrial Hydrogen Production

In industry, hydrogen is typically derived from fossil fuels. Specifically, the distribution of fossil fuel feedstock for hydrogen production is shown in Figure 1 [13]. Hydrogen is extracted from natural gas, coal and oil, in the ratio of 48%, 18% and 30%, respectively. The rest (4%) of hydrogen is produced from electrocatalysis.
Most of the hydrogen is produced from natural gas through steam reforming. Typically, methane is primarily applied, as steam reforming of methane is the most efficient route for hydrogen evolution. This process is carried out at the temperature of 970–1100 K and pressure of around 3.5 Mpa, and nickel is widely applied as the catalyst [14]. The basic reactions for steam reforming of natural gas are shown as Reactions 1 and 2.
C n H m + n H 2 O n C O + ( n + m 2 ) H 2
C O + H 2 O C O 2 + H 2 Δ H 0 = 41   kJ / mol
The target for the second step is converting CO to CO2, and utilizes the thermal energy left from the first step [15]. Even though steam reforming of natural gas is currently the primary process for hydrogen evolution, the disadvantages of this process are typical, such as large requirement of energy and large emission of CO2.
Coal is also widely applied to produce hydrogen through a gasification process; the reactions are shown in Reaction 3. Alkali metal (such as potassium) salts are utilized as the catalysts for this process [16]. Disadvantages of this process include not only the emission of large quantities of CO2, but also critical reaction conditions (1673 K and 2–7 MPa) are required to obtain a high conversion rate [17].
C H 0.8 + 0.7 O 2 + 0.6 H 2 O C O 2 + H 2
Due to the low cost of heavy oil or petroleum refinery residual oil, they are also widely applied as the feedstock for hydrogen production. The basic reaction is shown in Reaction 4. This process can occur with or without catalyst. Similar to the other above-mentioned processes, this process also emits a massive amount of both CO2 and CO.
2 C n H m + 1.5 n O 2 + H 2 O n C O 2 + n C O + ( m + 1 ) H 2
Despite hydrogen being currently derived from fossil fuels is the primary industrial process, explorations of new technologies and renewable energy sources have become more and more promising, especially in the background of severe energy shortage and environmental pollution. Water (H2O) is regarded as a promising source for hydrogen production, as it is carbon-free [18,19,20]. Electrolysis is capable of dissociating water to evolve hydrogen in ambient conditions. Theoretically, the applied voltage for water electrolysis is 1.23 V. However, the hydrogen evolution reaction (HER) at the cathode requires high over-potential to yield enough current density. Hence, the energy consumption for water electrolysis in practice is about 1.5–2.2 times higher than that in principle. Currently, the energy-conversion efficiency in practical use is only 48%–70%. To improve the efficiency, noble metal-based catalysts were applied. Hindrances for electrolysis on a large scale include not only the high energy demand for electricity, but also the high cost of the catalysts [12,21].
Currently, global industrial hydrogen production is approximately 7.2 × 1018 J per year. Most of the hydrogen were derived from fossil fuels, resulting in approximately 5 × 1011 kg CO2 emitted yearly. Such a large amount of CO2 emissions is inconsistent with the IEA’s (International Energy Agency) forecast of reducing CO2 emissions by 60% of the world by 2050, aiming to slow down the global warming [22]. Therefore, the research on photocatalytic hydrogen production is of great significance.

3. Photocatalysis for Hydrogen Evolution

Since 1972, photocatalytic water splitting for hydrogen evolution has been widely reported and regarded as one of the most promising methods for solar energy storage [23,24,25]. Theoretically, semiconductor-based photocatalysis can be described based on the band structure. When the energy of the irradiated light photon is larger than the band-gap energy of the semiconductor, electrons on the valence band (VB) may be activated and jump to the conduction band (CB) with positive holes left behind. The electrons on CB may either return the VB to recombine with the positive holes, or react with adsorbed species on the surface of the semiconductor [1,3,5,26,27]. When the potential of electrons on CB is more negative than the redox potential of H2/H2O, the separated electrons are capable of reducing water to hydrogen (Figure 2). It has been comprehensively reviewed and discussed in reported works [28,29,30,31,32]. Photocatalytic production of hydrogen has been realized not only by water splitting, but also by reforming. As shown in Figure 2, the minimum requirements for photocatalytic production of H2 are 1) the electrons on the conduction band of the semiconductor were capable reduce protons to form hydrogen and 2) the holes on the valence band of the semiconductor were capable of oxidizing water or R–OH to release protons. As reported in works on photocatalytic reforming for hydrogen evolution, it can be divided into three primary categories on the basis of the reforming precursor, which will be discussed in the following pages.

3.1. Photocatalytic Reforming of Methanol

Methanol is the most widely used compound in photocatalytic reforming [33,34]. Unlike photocatalytic water splitting, the products of photocatalytic reforming of organic compound were H2 and CO2, instead of H2 and O2 (Reaction 5) [35].
C H 3 O H + H 2 O h v C O 2 + 3 H 2     Δ H 0 = 130.897   kJ / mol
Reaction pathways for photocatalytic reforming of methanol on Au-TiO2 involve the following processes (Figure 3) [36]:
  • Photogeneration of charge carriers (electrons and holes);
  • Adsorbed species on the surface of photocatalyst react with separated charge carriers: i.e., adsorbed H2O and CH3OH are oxidized by holes with the formation of OH and various fragments including CH3O, CH2O, CHO, HCOO and HCOOH;
  • As-formed HCOOH is further oxidized to CO2;
  • A large quantity of protons formed from the last steps, and are reduced by the electrons either on the CB of TiO2 or Au to form H2.
The presence of methanol in the system acts as a scavenger for photo-generated holes on the VB of TiO2, which was in favor of improving the separation of photo-generated charge carriers. To further improve the photocatalytic activity, modifications on the photocatalyst are still required. Nanocrystalline TiO2 has paid attention throughout the past 50 years. Due to its strong oxidizing capacity under ultraviolet irradiation, high chemical stability, low cost, and environment-friendly, TiO2 is the most widely studied photocatalyst, and has been applied for reforming of methanol [37,38]. However, as for the intrinsic properties of TiO2, especially the large band gap (>3.2 eV) restricts its utilization of the visible light and the high electron–hole recombination rate, its photocatalytic activity is significantly limited [39,40,41]. Michael et al. reported on metal-modified TiO2 applied in photocatalytic reforming of methanol [42]. Four different metals (Pt, Pd, Au and Ag) were selected to be combined with TiO2. The products including H2, CO2 and CO were detected for photocatalytic reforming of methanol on different photocatalysts, the result of which is depicted in Figure 4. For pure TiO2, a small amount of H2 was formed. With the metal modification, the production rate of H2 was greatly improved. And Pt- modified TiO2 exhibited the highest activity, with a H2 production rate of 0.24 mmol h−1 gcat−1. The loading of the metals on the surface not only facilitated the separation of photogenerated charge carriers, but also provided more reactive sites, resulting in the enhanced photocatalytic activity of TiO2 in reforming of methanol. E. Tálas et al. further explored the Sn and Pt co-modified TiO2 in photocatalytic reforming of methanol [43]. It was found that the formation rate of H2 was further improved to 2 mmol h−1 gcat−1. Carbon nanotubes (CNT) had been proven to be an excellent co-catalyst for improving photocatalytic activity. G. Silva et al. prepared CNT-TiO2 using a hydration-dehydration method, and further loaded Pt on CNT-TiO2 through impregnation process [44]. It was found that CNT provided more reactive sites and further improved the photocatalytic activity of TiO2 in reforming methanol.
In order to reduce the cost of photocatalytic materials, research on improving the photoactivity of TiO2 modified by non-noble metals (for example, Ni, Cu, Co) was also conducted [45]. Bernareggi et al. investigated Cu and Pt commodified TiO2 on photocatalytic reforming of methanol [46]. The ternary composite, Cu-Pt-TiO2 exhibited excellent production rate of H2, which was 23.6 mmol h−1 gcat−1. Takuya Miwa et al. deposited CuO and Al2O3 onto TiO2, to improve the photocatalytic activity towards hydrogen production from methanol reforming [47]. With the formation of ternary structure, heterojunctions were fabricated, which favored improving the lifetime of photogenerated charge carriers. The prolonged lifetime can further improve the photocatalytic activity of TiO2. The results suggested that with the increase in the lifetime of photogenerated charge carriers, the photocatalytic activity of TiO2 in methanol reforming was significantly improved, with H2 production rate of 1.2 mmol h−1 gcat−1 [48,49]. Graphene oxide (GO) has been widely applied to improve the photocatalytic activity [50,51,52,53]. Majrik et al. prepared GO modified TiO2 by a hetero-coagulation method and evaluated its photocatalytic activity of methanol reforming [54]. It was found that with 2% of GO in the composite (GO-TiO2), the photocatalytic production rate of H2 was more than 5 times higher than pure TiO2.

3.2. Photocatalytic Reforming of Ethanol

Ethanol is an attractive source of hydrogen, due to its non-toxic and capable of being stored, transported and handled safely [55,56,57,58,59]. There are two main processes for the production of H2 from ethanol, which are steam reforming and thermal reforming [60,61]. Both processes are required to occur under high temperatures with noble metals as catalysts. Comparatively, photocatalytic reforming of ethanol for H2 production can be conducted at ambient conditions under irradiation. Kawai et al. reported on photocatalytic reforming of ethanol on TiO2 in 1981 [62]. Theoretically, the overall reaction for photocatalytic reforming of ethanol is shown in Reaction 6.
C H 3 C H 2 O H + 3 H 2 O h v 2 C O 2 + 6 H 2     Δ H 0 = 348.122   kJ / mol
For the reaction pathways, ethanol may be first photocatalytically converted to acetaldehyde with H2 released (Reaction 7). As-formed acetaldehyde may either reacts with water to form CO2 and H2 (Reaction 8), or be further converted to acetic acid (Reaction 9), which further reacts with water to form CO2 and H2 (Reaction 10) [63,64].
C H 3 C H 2 O H h v C H 3 C H O + H 2     Δ H 0 = 111.27   kJ / mol
C H 3 C H O + 3 H 2 O h v 2 C O 2 + 5 H 2     Δ H 0 = 236.852   kJ / mol
C H 3 C H O + H 2 O h v C H 3 C O O H + H 2     Δ H 0 = 15.798   kJ / mol
C H 3 C O O H + 2 H 2 O h v 4 C O 2 + 4 H 2     Δ H 0 = 565.968   kJ / mol
Photocatalytic reforming of ethanol has been widely studied on TiO2 [65,66,67]. The photoactivity was pretty low for pure TiO2, and should be improved further [68,69]. Owing to multiphasic TiO2, materials demonstrate higher performance with respect to the correspondent monophasic systems in many photocatalytic processes [70,71]. The crystal structure of TiO2 plays an important role in catalytic activity. There are three types crystal: anatase, rutile, brookite. The brookite is unstable, and the anatase-type TiO2 lattice contains mare defects and vacancies, which results in more oxygen vacancies to capture electrons. Rutile TiO2 is the most stable crystalline structure, but has few defects, which makes electrons and holes easy to recombine. Montini et al. prepared anatase/brookite TiO2 nano-composites by hydrothermal treatment of Na2Ti2O5 precursor [72]. Compared with TiO2 with single phase, the composite exhibited higher photocatalytic activity in ethanol reforming. Further, the authors loaded Pt onto this composite, and the production rate of H2 in photocatalytic reforming of ethanol improved to 1.2 mmol h−1 gcat−1. Yin et al. prepared Au onto TiO2 to form a core-shell structure [73]. The preparation procedures are shown in Figure 5. TiO2 as the shell can absorb light energy, and Au as the core acts as the electron acceptor and narrow the band gap. They found that controllable morphology of TiO2 nanocrystals facilitates electron-hole separation and can expose more active surfaces. With the fabrication, the photocatalytic activity of TiO2 in ethanol reforming improved to 1.07 mol h−1 gcat−1.
Other than TiO2, Wang et al. investigated on ZnO applied for photocatalytic reforming of ethanol [74]. ZnO with various morphologies including needle-like (ZnO-n), flower-like (ZnO-f) and rods (ZnO-r), were prepared. It was found that ZnO-f exhibited the highest photocatalytic activity with the H2 production rate of 0.305 mmol h−1 gcat−1. That might be due to its flower-like morphology that may scatter the incoming light, enhancing light absorption efficiency and increasing the photocatalytic efficiency. With further surface modification with Au onto ZnO-f, the photocatalytic activity was further improved, and the H2 production rate was 0.427 mmol h−1 gcat−1. Fu et al. prepared ZnSn(OH)6 nanocubes (ZHS) by a solvothermal method [75]. The photocatalyst obtained in the autoclave with a longer reaction time has higher crystallinity and better photocatalytic efficiency. The H2 production rate was measured at as high as 16 mmol h−1 gcat−1. ZHS with moderate longer reaction time has higher crystallinity, because amorphous materials with low crystallinity usually comprise numerous defects. These defects may act as recombination centers for the electrons and holes and result in the low photocatalytic efficiency.

3.3. Photocatalytic Reforming of Biomass

Biomass, as a promising renewable energy source, is a possible alternative to fossil fuels. The reforming of biomass is an effective way to convert biofuels to chemical fuels [76]. Production of hydrogen from biomass is feasible and very promising for future energy development [77,78]. Conventional processes for H2 derived from biomass require critical reaction conditions [79,80,81,82,83,84]. Photocatalysis applied for reforming of biomass not only overcomes the problem of critical conditions required by thermal catalysis, but also combines the solar energy with biofuels [85].
The mechanism of photocatalytic reforming of biomass is similar to the photocatalytic reforming of other organics [35]. When the substrates are biomass, the substances consumed in this process are renewable [86,87]. Taking glucose as an example, the reaction pathways for photocatalytic reforming of glucose are described in Figure 6. It can be found that the photogenerated electrons and holes separately react with adsorbed species on the surface of photocatalyst, and trigger a series of redox reactions. Overall, 1 mol of glucose can be dissociated and produce 12 mol of H2 [85]. Typical examples of photocatalytic reforming of biomass are summarized in Table 1.

3.3.1. TiO2

In previous discussions, Pt/TiO2 exhibited photocatalytic activity in reforming. Xu et al. investigated the effect of crystals on photocatalytic hydrogen production. It was found that compared with Pt/P25, the finely adjusted anatase-rutile Pt/TiO2 photocatalyst improved the overall photocatalytic activity by 3 to 5 times, and greatly reduced the CO concentration in the hydrogen product, which can meet the fuel application requirements for battery technology. They believe that this result can be attributed to the adjustment of the crystal form in favor of light-induced charge separation, as well as the adjustment of the acidity and basicity of the interface surface. [89,97]. Bellardita et al. doped N and W into the lattice of TiO2 using a hydrothermal method and further surface modified by Pt [88]. It was found that hydrogen production rate on Pt/(N and W)-TiO2 in photocatalytic reforming of glucose under natural solar light was as high as 1.0 mmol h−1 gcat−1. This activity under sunlight proves the applicability of such materials under realistic conditions. Yasuda et al. further studied Pt/TiO2 in photocatalytic reforming of Italian ryegrass and Napier grass [98]. They used a combination of biological treatment and light reforming to first enzymatically saccharify the two alkali-treated grasses, and then used Pt/TiO2 to perform phototransformation under UV light irradiation. Through this process, they can convert two natural plants into ethanol and hydrogen with recovery rates of 82.7% and 77.2%, respectively. However, the photocatalysts they used are still poorly responsive to visible light, which limits the application of this technology in real-life conditions.
Wu et al. investigated Pt, Rh, Ru, Cu, Ni Au and Ir modified TiO2 in photocatalytic reforming of glucose for hydrogen production [90]. It was found that the supported metal exhibited the property of trapping electrons at the Schottky barrier of metal in contact with TiO2 surface, which greatly improved the hydrogen generation rate and also reduced the CO selectivity. The specific impact depends on the type of metal, the generation rate of H2 on these metal- modified TiO2 followed the order: Ir < Ru < Au < Ni ≈ Cu ≈ Pt < Rh. Fu et al. [85] tested the photocatalytic reforming activity of another group of metals/TiO2 photocatalysts, and the test results are shown in Figure 7. It can be found that Pd/TiO2 exhibited the highest activity in photocatalytic reforming of glucose. In the research, they also found that the rate of hydrogen production using glucose depends on the initial pH of the solution, and the hydrogen production rate was highest at pH 11 in their system.

3.3.2. Metal Sulfide

Even though TiO2 has been widely investigated in photocatalytic reforming of biomass, it is necessary to explore other visible light-responsive materials applied in this field. Metal sulfides such as cadmium sulfide (CdS) and molybdenum disulfide (MoS2) exhibited excellent activity in both photocatalytic water splitting and reforming to evolve H2 [92]. Li et al. successfully prepared CdS/MoS2 composites and tested their performance in photocatalytic reforming of glucose [92]. It was found that the composite containing 40 wt% of CdS exhibited the highest activity with a hydrogen production rate of 55.0 μmol h−1 gcat−1. They speculate that the heterojunction formed between MoS2 and CdS promotes the interface charge transfer, inhibits the recombination of photogenerated electron-hole pairs, and then promotes the hydrogen evolution reaction. At the same time, due to the good matching characteristics of the conduction band edges of the two materials, the electrons in the CdS conduction band can be transferred to MoS2 with higher efficiency, preventing the accumulation of electrons in the CdS conduction band, inhibiting the photocorrosion of the catalyst itself, and greatly improving stability. Then, Li et al. synthesized NiS-CdS composite through a two-pot solvothermal process [91]. This composite exhibited excellent visible light-induced activity. The apparent quantum efficiency (AQE) of H2 in photocatalytic reforming of lactic acid and lignin acid in the presence of different NiS-CdS composites was shown in Figure 8. It was found that the 0.2-NiS/CdS exhibited the highest activity with the AQE as high as 29.33% for lactic acid and 44.9% for lignin acid, and the high stability was also showed. However, the potential toxicity of this photocatalyst has limited its application to some extent.

3.3.3. Graphene Analog Photocatalyst

Photocatalysts commonly used in the past are usually metal and its derivatives. They all have more or less economic and environmental problems and are difficult to scale up. Graphitic carbon nitride (g-C3N4) has attracted wide attention as an advanced non-metal photocatalyst [26,99]. Xu et al. improved the photocatalytic activity of g-C3N4 in reforming of biomass with modification of ZnS. After examining the performance of carbon nitride derived from different raw materials as photocatalysts, they believed that the rate of reaction hydrogen evolution was affected by various factors such as the specific surface area of the catalyst, the band width, and the particle size of the doped material [94]. Nguyen et al. combined graphene with g-C3N4. Due to the vacancy defects of graphene oxide itself are not conducive to photocatalytic reactions, they synthesized sulfur and nitrogen co-doped graphene oxide dots (SNGODs). And found that the photocatalytic activity of SNGODs in reforming of sugar and glucose were both improved significantly [95]. Speltini et al. treated g-C3N4 in hot HNO3 aqueous solution to obtain oxidized g-C3N4 (O-g-C3N4) [93]. After treatment, the production rate of H2 in photocatalytic reforming of glucose was about 26 times higher than pure g-C3N4. Further, when the O-g-C3N4 were treated through ultrasonication, the production rate further improved to 1370 mmol h−1 gcat−1. These studies have confirmed the possibility of hydrogen production with non-metal photocatalysts. But g-C3N4 catalysts doped with no metal elements still exhibited low activity and did not exhibit sufficient stability. These problems hinder its application in actual production.

3.3.4. Oxide Solid Solution Photocatalyst

From the aspect of thermodynamics, Jing et al. proposed that oxide photocatalysts were generally more stable and exhibited more positive valence band positions than sulfides and nitrides, which are more suitable for the oxidation of biomass [96]. They found that doping a certain amount of Y can promote the transformation of the crystal phase of BiVO4 from a monoclinic phase to a tetragonal phase, which is beneficial to the improvement of catalytic activity. Therefore, an oxide solid solution, BixY1−xVO4 (BYV) was prepared and tested in photocatalytic reforming of glucose under visible light. The test results are shown in Figure 9. It was found that when the solution pH was at 3, the production rate of H2 was the highest at 50 μmol h−1 gcat−1. The results also showed that the CO2 gas produced by the hydrogen evolution may inhibit the efficiency of the hydrogen evolution reaction. How to transfer gas products in time is obviously an important issue in the actual production of photocatalytic reforming.

4. Future Work and Concluding Remarks

Photocatalytic reforming is very promising for hydrogen evolution at ambient conditions. Through this process, solar energy can be converted and stored in the form of chemical energy. To date, various chemical compounds, especially alcohols and biomass, have been widely applied as the feedstock for photocatalytic reforming. Typical works on this research subject have been comprehensively discussed in this review. Albeit photocatalysis exhibited superiorities in reforming for H2 production, the hindrances were also obvious and need to be overcome before its wide application. Possible approaches to overcome these hindrances and recommendations for future work are suggested below:
  • Low photocatalytic efficiency. Currently, the photocatalytic activity is too low to use on a large scale, which is also the most significant problem for not only photocatalytic reforming, but also all other photocatalytic applications. With the rapid development of material science, design and synthesis of novel materials are one of the most effective approaches. Meanwhile, exploration of effective modification methods to improve the photocatalytic activity is of the essence, as well.
  • Low visible light-responsivity. When we aim to utilize solar light as the energy source, it is essential to expand the utilization of the solar spectrum, especially the visible-light part. Feasible approaches include the modifications of existing photocatalyst to narrow its band gap, and preparation of new materials with narrow band gap. These effective modification methods include doping, noble metal nanoparticles loading and fabrication of heterojunction. Recently, with the rapid development of material science and preparation technologies, many advanced semiconductors exhibited excellent visible-light-induced photoactivity, including g-C3N4 and bismuth-based semiconductors (BiOBr, Bi2WO6, BiVO4 etc.). All those materials are of promise to be applied in photocatalytic reforming for H2 production, whereas were rarely been reported in this research field.
  • Ambiguous mechanism for photocatalytic processes. Most of the mechanism studies on photocatalysis were on the basis of band structures. However, some phenomena were difficult to explain, based on the electrical band structures. Instead, advanced material characterization techniques, as well as simulation methodologies should be adopted to acquire an insight into the photocatalytic process.
  • Rare studies on photoreactor design. In the process of scaling-up photocatalysis, photoreactor design plays significant role. Photocatalytic activity can also be improved with rational photoreactor design. However, most of studies on photocatalytic reforming were still in lab scale. More contributions on the exploration of photoreactor design should be made to upscale this process.
  • No standard platform. So far, there has been no standard platform that can be used to compare photoactivity for either different materials or the same material tested from a different lab. There is an urgent need to establish a standard to evaluate different materials, and guide the direction of future work.
Aiming to provide an alternative approach for thermal reforming of hydrocarbons for hydrogen production, photocatalytic process is of promise with superiorities such as renewable energy resource applied and mild conditions required. However, due to the low activity which is still far below the requirements if applied in practice. More contributions should be made to further develop this advanced technology via breaking the bottleneck.

Author Contributions

Conceptualization, Literature review, Writing–Original draft preparation (Y.Y.); Writing–Original draft preparation and editing (X.G.); Conceptualization, supervision (Z.L.); Supervision, Writing–Original draft preparation, review and editing, manuscript submission and correspondence (X.M.). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Taishan Scholar Foundation of Shandong province, grant number tsqn201909058.

Acknowledgments

X.M. acknowledges the financial supports from the Key Laboratory of Marine Chemistry Theory and Technology (Ministry of Education, Ocean University of China).

Conflicts of Interest

The authors declare no conflict of interest

References

  1. Meng, X.; Zhang, Z.; Li, X. Synergetic photoelectrocatalytic reactors for environmental remediation: A review. J. Photochem. Photobiol. C Photochem. Rev. 2015, 24, 83–101. [Google Scholar] [CrossRef]
  2. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  3. Li, Z.; Meng, X. Recent development on palladium enhanced photocatalytic activity: A review. J. Alloys Compd. 2020, 830, 154669. [Google Scholar] [CrossRef]
  4. Meng, X.; Yun, N.; Zhang, Z. Recent advances in computational photocatalysis: A review. Can. J. Chem. Eng. 2019, 97, 1982–1998. [Google Scholar] [CrossRef]
  5. Meng, X.; Zhang, Z. Bismuth-based Photocatalytic Semiconductors: Introduction, Challenges and Possible Approaches. J. Mol. Catal. A Chem. 2016, 423, 533–549. [Google Scholar] [CrossRef]
  6. Zhang, K.-L.; Liu, C.-M.; Huang, F.-Q.; Zheng, C.; Wang, W.-D. Study of the electronic structure and photocatalytic activity of the BiOCl photocatalyst. Appl. Catal. B Environ. 2006, 68, 125–129. [Google Scholar] [CrossRef]
  7. Cao, S.; Chan, T.-S.; Lu, Y.-R.; Shi, X.; Fu, B.; Wu, Z.; Li, H.; Liu, K.; Alzuabi, S.; Cheng, P.; et al. Photocatalytic pure water splitting with high efficiency and value by Pt/porous brookite TiO2 nanoflutes. Nano Energy 2020, 67. [Google Scholar] [CrossRef]
  8. Zhang, F.; Zhao, P.; Niu, M.; Maddy, J. The survey of key technologies in hydrogen energy storage. Int. J. Hydrog. Energy 2016, 41, 14535–14552. [Google Scholar] [CrossRef]
  9. Abe, J.O.; Popoola, A.P.I.; Ajenifuja, E.; Popoola, O.M. Hydrogen energy, economy and storage: Review and recommendation. Int. J. Hydrog. Energy 2019, 44, 15072–15086. [Google Scholar] [CrossRef]
  10. Dawood, F.; Anda, M.; Shafiullah, G.M. Hydrogen production for energy: An overview. Int. J. Hydrog. Energy 2020, 45, 3847–3869. [Google Scholar] [CrossRef]
  11. Ibrahim, A.B.A.; Akilli, H. Supercritical water gasification of wastewater sludge for hydrogen production. Int. J. Hydrog. Energy 2019, 44, 10328–10349. [Google Scholar] [CrossRef]
  12. Xu, Y.; Zhang, B. Recent Advances in Electrochemical Hydrogen Production from Water Assisted by Alternative Oxidation Reactions. ChemElectroChem 2019, 6, 3214–3226. [Google Scholar] [CrossRef]
  13. Kothari, R.; Buddhi, D.; Sawhney, R.L. Comparison of environmental and economic aspects of various hydrogen production methods. Renew. Sustain. Energy Rev. 2008, 12, 553–563. [Google Scholar] [CrossRef]
  14. Chisalita, D.-A.; Cormos, C.-C. Techno-economic assessment of hydrogen production processes based on various natural gas chemical looping systems with carbon capture. Energy 2019, 181, 331–344. [Google Scholar] [CrossRef]
  15. Yamamuro, K.; Tamura, S.; Watanabe, R.; Sekine, Y. Hydrogen Production by Water Gas Shift Reaction Over Pd–K Impregnated Co Oxide Catalyst. Catal. Lett. 2013, 143, 339–344. [Google Scholar] [CrossRef]
  16. Seçer, A.; Küçet, N.; Fakı, E.; Hasanoğlu, A. Comparison of co–gasification efficiencies of coal, lignocellulosic biomass and biomass hydrolysate for high yield hydrogen production. Int. J. Hydrog. Energy 2018, 43, 21269–21278. [Google Scholar] [CrossRef]
  17. Kim, J.; Choi, H.; Lim, J.; Rhim, Y.; Chun, D.; Kim, S.; Lee, S.; Yoo, J. Hydrogen production via steam gasification of ash free coals. Int. J. Hydrog. Energy 2013, 38, 6014–6020. [Google Scholar] [CrossRef] [Green Version]
  18. Yuzer, B.; Selcuk, H.; Chehade, G.; Demir, M.E.; Dincer, I. Evaluation of hydrogen production via electrolysis with ion exchange membranes. Energy 2019. [Google Scholar] [CrossRef]
  19. Bairachnyi, V.; Rudenko, N.; Zhelavska, Y.; Pilipenko, A. Using aluminum alloys in the electrochemical hydrogen production. Mater. Today Proc. 2019, 6, 299–304. [Google Scholar] [CrossRef]
  20. Raptis, D.; Seferlis, A.K.; Mylona, V.; Politis, C.; Lianos, P. Electrochemical hydrogen and electricity production by using anodes made of commercial aluminum. Int. J. Hydrog. Energy 2019, 44, 1359–1365. [Google Scholar] [CrossRef]
  21. Wang, M.; Chen, L.; Sun, L. Recent progress in electrochemical hydrogen production with earth-abundant metal complexes as catalysts. Energy Environ. Sci. 2012, 5. [Google Scholar] [CrossRef]
  22. Voldsund, M.; Jordal, K.; Anantharaman, R. Hydrogen production with CO2 capture. Int. J. Hydrog. Energy 2016, 41, 4969–4992. [Google Scholar] [CrossRef]
  23. Li, Z.; Meng, X.; Zhang, Z. Recent development on MoS2-based photocatalysis: A review. J. Photochem. Photobiol. C Photochem. Rev. 2018, 35, 39–55. [Google Scholar] [CrossRef]
  24. Kim, M.K.; Sim, W.H.; Choi, M.; Lim, H.; Kwon, Y.; Jeong, H.M. Electrochemically Li-intercalated TiO2 nanoparticles for High performance photocatalytic production of hydrogen. Catal. Today 2019. [Google Scholar] [CrossRef]
  25. Zhu, J.; Zäch, M. Nanostructured materials for photocatalytic hydrogen production. Curr. Opin. Colloid Interface Sci. 2009, 14, 260–269. [Google Scholar] [CrossRef]
  26. Meng, X.; Zisheng, Z. Two Dimensional Graphitic Materials for Photoelectrocatalysis: A Short Review. Catal. Today 2018. [Google Scholar] [CrossRef]
  27. Li, Z.; Meng, X.; Zhang, Z. An Effective Approach to Improve the Photocatalytic Activity of Graphitic Carbon Nitride via Hydroxyl Surface Modification. Catalysts 2018, 9, 17. [Google Scholar] [CrossRef] [Green Version]
  28. Ni, M.; Leung, M.K.H.; Leung, D.Y.C.; Sumathy, K. A review and recent developments in photocatalytic water-splitting using TiO2 for hydrogen production. Renew. Sustain. Energy Rev. 2007, 11, 401–425. [Google Scholar] [CrossRef]
  29. Ismail, A.A.; Bahnemann, D.W. Photochemical splitting of water for hydrogen production by photocatalysis: A review. Sol. Energy Mat. Sol. Cells 2014, 128, 85–101. [Google Scholar] [CrossRef]
  30. Acar, C.; Dincer, I.; Naterer, G.F. Review of photocatalytic water-splitting methods for sustainable hydrogen production. Int. J. Energy Res. 2016, 40, 1449–1473. [Google Scholar] [CrossRef]
  31. Acar, C.; Dincer, I.; Zamfirescu, C. A review on selected heterogeneous photocatalysts for hydrogen production. Int. J. Energy Res. 2014, 38, 1903–1920. [Google Scholar] [CrossRef]
  32. Shimura, K.; Yoshida, H. Heterogeneous photocatalytic hydrogen production from water and biomass derivatives. Energy Environ. Sci. 2011, 4, 2467–2481. [Google Scholar] [CrossRef]
  33. Dickinson, A.; James, D.; Perkins, N.; Cassidy, T.; Bowker, M. The photocatalytic reforming of methanol. J. Mol. Catal. A Chem. 1999, 146, 211–221. [Google Scholar] [CrossRef]
  34. Li, Z.; Ivanenko, A.; Meng, X.; Zhang, Z. Photocatalytic oxidation of methanol to formaldehyde on bismuth-based semiconductors. J. Hazard. Mater. 2019, 380, 120822. [Google Scholar] [CrossRef] [PubMed]
  35. Chiarello, G.L.; Selli, E. Photocatalytic production of hydrogen. In Advances in Hydrogen Production, Storage and Distribution; Woodhead Publishing: Cambridge, UK, 2014; pp. 216–247. [Google Scholar] [CrossRef]
  36. 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. Hydrog. Energy 2008, 33, 1243–1251. [Google Scholar] [CrossRef]
  37. Lee, C.; Park, J.; Jeon, Y.; Park, J.-I.; Einaga, H.; Truong, Y.B.; Kyratzis, I.L.; Mochida, I.; Choi, J.; Shul, Y.-G. Phosphate-Modified TiO2/ZrO2 Nanofibrous Web Composite Membrane for Enhanced Performance and Durability of High-Temperature Proton Exchange Membrane Fuel Cells. Energy Fuels 2017, 31, 7645–7652. [Google Scholar] [CrossRef]
  38. Ola, O.; Maroto-Valer, M.M. Review of material design and reactor engineering on TiO2 photocatalysis for CO2 reduction. J. Photochem. Photobiol. C Photochem. Rev. 2015, 24, 16–42. [Google Scholar] [CrossRef] [Green Version]
  39. Kim, D.S.; Kwak, S.-Y. The hydrothermal synthesis of mesoporous TiO2 with high crystallinity, thermal stability, large surface area, and enhanced photocatalytic activity. Appl. Catal. A Gen. 2007, 323, 110–118. [Google Scholar] [CrossRef]
  40. Sreethawong, T.; Yamada, Y.; Kobayashi, T.; Yoshikawa, S. Catalysis of nanocrystalline mesoporous TiO2 on cyclohexene epoxidation with H2O2: Effects of mesoporosity and metal oxide additives. J. Mol. Catal. A Chem. 2005, 241, 23–32. [Google Scholar] [CrossRef]
  41. Tschirch, J.; Bahnemann, D.; Wark, M.; Rathouský, J. A comparative study into the photocatalytic properties of thin mesoporous layers of TiO2 with controlled mesoporosity. J. Photochem. Photobiol. A Chem. 2008, 194, 181–188. [Google Scholar] [CrossRef]
  42. Bingham, M.; Mills, A. Photonic efficiency and selectivity study of M (M=Pt, Pd, Au and Ag)/TiO2 photocatalysts for methanol reforming in the gas phase. J. Photochem. Photobiol. A Chem. 2019. [Google Scholar] [CrossRef]
  43. Tálas, E.; Pászti, Z.; Korecz, L.; Domján, A.; Németh, P.; Szíjjártó, G.P.; Mihály, J.; Tompos, A. PtOx-SnOx-TiO2 catalyst system for methanol photocatalytic reforming: Influence of cocatalysts on the hydrogen production. Catal. Today 2018, 306, 71–80. [Google Scholar] [CrossRef] [Green Version]
  44. Silva, C.G.; Sampaio, M.J.; Marques, R.R.N.; Ferreira, L.A.; Tavares, P.B.; Silva, A.M.T.; Faria, J.L. Photocatalytic production of hydrogen from methanol and saccharides using carbon nanotube-TiO2 catalysts. Appl. Catal. B Environ. 2015, 178, 82–90. [Google Scholar] [CrossRef] [Green Version]
  45. Irie, H.; Kamiya, K.; Shibanuma, T.; Miura, S.; Tryk, D.A.; Yokoyama, T.; Hashimoto, K. Visible Light-Sensitive Cu(II)-Grafted TiO2 Photocatalysts: Activities and X-ray Absorption Fine Structure Analyses. J. Phys. Chem. C 2009, 113, 10761–10766. [Google Scholar] [CrossRef]
  46. Bernareggi, M.; Dozzi, M.; Bettini, L.; Ferretti, A.; Chiarello, G.; Selli, E. Flame-Made Cu/TiO2 and Cu-Pt/TiO2 Photocatalysts for Hydrogen Production. Catalysts 2017, 7, 301. [Google Scholar] [CrossRef]
  47. Miwa, T.; Kaneco, S.; Katsumata, H.; Suzuki, T.; Ohta, K.; Chand Verma, S.; Sugihara, K. Photocatalytic hydrogen production from aqueous methanol solution with CuO/Al2O3/TiO2 nanocomposite. Int. J. Hydrog. Energy 2010, 35, 6554–6560. [Google Scholar] [CrossRef]
  48. Tawalare, P.K.; Belsare, P.D.; Moharil, S.V. Semiconductor host for designing phosphors for modification of solar spectrum. Opt. Mater. 2020, 100, 109668. [Google Scholar] [CrossRef]
  49. Liang, Z.; Bai, X.; Hao, P.; Guo, Y.; Xue, Y.; Tian, J.; Cui, H. Full solar spectrum photocatalytic oxygen evolution by carbon-coated TiO2 hierarchical nanotubes. Appl. Catal. B Environ. 2019, 243, 711–720. [Google Scholar] [CrossRef]
  50. Meng, X.; Zhang, Z. Plasmonic ternary Ag–rGO–Bi2MoO6 composites with enhanced visible light-driven photocatalytic activity. J. Catal. 2016, 344, 616–630. [Google Scholar] [CrossRef]
  51. Meng, X.; Li, Z.; Zhang, Z. Palladium nanoparticles and rGO co-modified BiVO4 with greatly improved visible light-induced photocatalytic activity. Chemosphere 2018, 198, 1–12. [Google Scholar] [CrossRef]
  52. Meng, X.; Zhang, Z. Bi2MoO6 co-modified by reduced graphene oxide and palladium (Pd2+ and Pd0) with enhanced photocatalytic decomposition of phenol. Appl. Catal. B Environ. 2017, 209, 383–393. [Google Scholar] [CrossRef]
  53. Krissanasaeranee, M.; Wongkasemjit, S.; Cheetham, A.K.; Eder, D. Complex carbon nanotube-inorganic hybrid materials as next-generation photocatalysts. Chem. Phys. Lett. 2010, 496, 133–138. [Google Scholar] [CrossRef]
  54. Majrik, K.; Turcsányi, Á.; Pászti, Z.; Szabó, T.; Domján, A.; Mihály, J.; Tompos, A.; Dékány, I.; Tálas, E. Graphite Oxide-TiO2 Nanocomposite Type Photocatalyst for Methanol Photocatalytic Reforming Reaction. Top. Catal. 2018, 61, 1323–1334. [Google Scholar] [CrossRef] [Green Version]
  55. Ashok, J.; Dewangan, N.; Das, S.; Hongmanorom, P.; Wai, M.H.; Tomishige, K.; Kawi, S. Recent progress in the development of catalysts for steam reforming of biomass tar model reaction. Fuel Process. Technol. 2020, 199. [Google Scholar] [CrossRef]
  56. Chen, D.; Wang, W.; Liu, C. Ni-encapsulated graphene chainmail catalyst for ethanol steam reforming. Int. J. Hydrog. Energy 2019, 44, 6560–6572. [Google Scholar] [CrossRef]
  57. Cifuentes, A.; Torres, R.; Llorca, J. Modelling of the ethanol steam reforming over Rh-Pd/CeO2 catalytic wall reactors. Int. J. Hydrog. Energy 2019. [Google Scholar] [CrossRef] [Green Version]
  58. Wang, F.; Cai, W.; Tana; Provendier, H.; Schuurman, Y.; Descorme, C.; Mirodatos, C.; Shen, W. Ageing analysis of a model Ir/CeO2 catalyst in ethanol steam reforming. Appl. Catal. B Environ. 2012, 125, 546–555. [Google Scholar] [CrossRef]
  59. Wu, R.-C.; Tang, C.-W.; Huang, H.-H.; Wang, C.-C.; Chang, M.-B.; Wang, C.-B. Effect of boron doping and preparation method of Ni/Ce0.5Zr0.5O2 catalysts on the performance for steam reforming of ethanol. Int. J. Hydrog. Energy 2019, 44, 14279–14289. [Google Scholar] [CrossRef]
  60. Freni, S.; Cavallaro, S.; Mondello, N.; Spadaro, L.; Frusteri, F. Production of hydrogen for MC fuel cell by steam reforming of ethanol over MgO supported Ni and Co catalysts. Catal. Commun. 2003, 4, 259–268. [Google Scholar] [CrossRef]
  61. Cavallaro, S.; Chiodo, V.; Vita, A.; Freni, S. Hydrogen production by auto-thermal reforming of ethanol on Rh/Al2O3 catalyst. J. Power Sources 2003, 123, 10–16. [Google Scholar] [CrossRef]
  62. Sakata, T.; Kawai, T. Heterogeneous photocatalytic production of hydrogen and methane from ethanol and water. Chem. Phys. Lett. 1981, 80, 341–344. [Google Scholar] [CrossRef]
  63. Sola, A.C.; Homs, N.; Ramírez de la Piscina, P. Photocatalytic H2 production from ethanol(aq) solutions: The effect of intermediate products. Int. J. Hydrog. Energy 2016, 41, 19629–19636. [Google Scholar] [CrossRef]
  64. Sola, A.C.; Ramírez de la Piscina, P.; Homs, N. Behaviour of Pt/TiO2 catalysts with different morphological and structural characteristics in the photocatalytic conversion of ethanol aqueous solutions. Catal. Today 2020, 341, 13–20. [Google Scholar] [CrossRef]
  65. Bamwenda, G.R.; Tsubota, S.; Nakamura, T.; Haruta, M. Photoassisted hydrogen production from a water-ethanol solution: A comparison of activities of Au TiO2 and Pt TiO2. J. Photochem. Photobiol. A Chem. 1995, 89, 177–189. [Google Scholar] [CrossRef]
  66. Strataki, N.; Bekiari, V.; Kondarides, D.I.; Lianos, P. Hydrogen production by photocatalytic alcohol reforming employing highly efficient nanocrystalline titania films. Appl. Catal. B Environ. 2007, 77, 184–189. [Google Scholar] [CrossRef]
  67. Zou, J.-J.; He, H.; Cui, L.; Du, H.-Y. Highly efficient Pt/TiO2 photocatalyst for hydrogen generation prepared by a cold plasma method. Int. J. Hydrog. Energy 2007, 32, 1762–1770. [Google Scholar] [CrossRef]
  68. Luo, B.; Song, R.; Jing, D. Particle aggregation behavior during photocatalytic ethanol reforming reaction and its correlation with the activity of H2 production. Colloids Surf. A Physicochem. Eng. Asp. 2017, 535, 114–120. [Google Scholar] [CrossRef]
  69. Meng, X.; Li, Z.; Zhang, Z. Pd-nanoparticle-decorated peanut-shaped BiVO4 with improved visible light-driven photocatalytic activity comparable to that of TiO2 under UV light. J. Catal. 2017, 356, 53–64. [Google Scholar] [CrossRef]
  70. Liu, Y.; Wang, Z.; Wang, W.; Huang, W. Engineering highly active TiO2 photocatalysts via the surface-phase junction strategy employing a titanate nanotube precursor. J. Catal. 2014, 310, 16–23. [Google Scholar] [CrossRef]
  71. Zhao, B.; Chen, F.; Jiao, Y.; Yang, H.; Zhang, J. Ag0-loaded brookite/anatase composite with enhanced photocatalytic performance towards the degradation of methyl orange. J. Mol. Catal. A Chem. 2011, 348, 114–119. [Google Scholar] [CrossRef]
  72. Romero Ocaña, I.; Beltram, A.; Delgado Jaén, J.J.; Adami, G.; Montini, T.; Fornasiero, P. Photocatalytic H2 production by ethanol photodehydrogenation: Effect of anatase/brookite nanocomposites composition. Inorg. Chim. Acta 2015, 431, 197–205. [Google Scholar] [CrossRef]
  73. Goebl, J.; Joo, J.B.; Dahl, M.; Yin, Y. Synthesis of tailored Au@TiO2 core–shell nanoparticles for photocatalytic reforming of ethanol. Catal. Today 2014, 225, 90–95. [Google Scholar] [CrossRef]
  74. Sampaio, M.J.; Oliveira, J.W.L.; Sombrio, C.I.L.; Baptista, D.L.; Teixeira, S.R.; Carabineiro, S.A.C.; 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]
  75. Fu, X.; Leung, D.Y.C.; Wang, X.; Xue, W.; Fu, X. Photocatalytic reforming of ethanol to H2 and CH4 over ZnSn(OH)6 nanocubes. Int. J. Hydrog. Energy 2011, 36, 1524–1530. [Google Scholar] [CrossRef]
  76. Puga, A.V. Photocatalytic production of hydrogen from biomass-derived feedstocks. Coord. Chem. Rev. 2016, 315, 1–66. [Google Scholar] [CrossRef]
  77. Guo, F.; Zhao, X.; Peng, K.; Liang, S.; Jia, X.; Qian, L. Catalytic reforming of biomass primary tar from pyrolysis over waste steel slag based catalysts. Int. J. Hydrog. Energy 2019, 44, 16224–16233. [Google Scholar] [CrossRef]
  78. Iwasaki, W. A consideration of the economic efficiency of hydrogen production from biomass. Int. J. Hydrog. Energy 2003, 28, 939–944. [Google Scholar] [CrossRef]
  79. Balat, H.; Kırtay, E. Hydrogen from biomass—Present scenario and future prospects. Int. J. Hydrog. Energy 2010, 35, 7416–7426. [Google Scholar] [CrossRef]
  80. Ni, M.; Leung, D.Y.C.; Leung, M.K.H.; Sumathy, K. An overview of hydrogen production from biomass. Fuel Process. Technol. 2006, 87, 461–472. [Google Scholar] [CrossRef]
  81. Setiabudi, H.D.; Aziz, M.A.A.; Abdullah, S.; Teh, L.P.; Jusoh, R. Hydrogen production from catalytic steam reforming of biomass pyrolysis oil or bio-oil derivatives: A review. Int. J. Hydrog. Energy 2019. [Google Scholar] [CrossRef]
  82. Jeong, Y.-S.; Choi, Y.-K.; Kang, B.-S.; Ryu, J.-H.; Kim, H.-S.; Kang, M.-S.; Ryu, L.-H.; Kim, J.-S. Lab-scale and pilot-scale two-stage gasification of biomass using active carbon for production of hydrogen-rich and low-tar producer gas. Fuel Process. Technol. 2020, 198, 106240. [Google Scholar] [CrossRef]
  83. Zhang, S.-p.; Li, X.-j.; Li, Q.-y.; Xu, Q.-l.; Yan, Y.-j. Hydrogen production from the aqueous phase derived from fast pyrolysis of biomass. J. Anal. Appl. Pyrolysis 2011, 92, 158–163. [Google Scholar] [CrossRef]
  84. Watanabe, M.; Inomata, H.; Arai, K. Catalytic hydrogen generation from biomass (glucose and cellulose) with ZrO2 in supercritical water. Biomass Bioenergy 2002, 22, 405–410. [Google Scholar] [CrossRef]
  85. Fu, X.; Long, J.; Wang, X.; Leung, D.; 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. Hydrog. Energy 2008, 33, 6484–6491. [Google Scholar] [CrossRef]
  86. Navarro, R.M.; Sánchez-Sánchez, M.C.; Alvarez-Galvan, M.C.; Valle, F.d.; Fierro, J.L.G. Hydrogen production from renewable sources: Biomass and photocatalytic opportunities. Energy Environ. Sci. 2009, 2, 35–54. [Google Scholar] [CrossRef]
  87. Melo, M.d.O.; Silva, L.A. Photocatalytic production of hydrogen: An innovative use for biomass derivatives. J. Braz. Chem. Soc. 2011, 22, 1399–1406. [Google Scholar] [CrossRef] [Green Version]
  88. Bellardita, M.; García-López, E.I.; Marcì, G.; Nasillo, G.; Palmisano, L. Photocatalytic Solar Light H2 Production by Aqueous Glucose Reforming. Eur. J. Inorg. Chem. 2018, 2018, 4522–4532. [Google Scholar] [CrossRef] [Green Version]
  89. Xu, Q.; Ma, Y.; Zhang, J.; Wang, X.; Feng, Z.; Li, C. Enhancing hydrogen production activity and suppressing CO formation from photocatalytic biomass reforming on Pt/TiO2 by optimizing anatase–rutile phase structure. J. Catal. 2011, 278, 329–335. [Google Scholar] [CrossRef]
  90. Wu, G.; Chen, T.; Zhou, G.; Zong, X.; Li, C. H2 production with low CO selectivity from photocatalytic reforming of glucose on metal/TiO2 catalysts. Sci. China Ser. B Chem. 2008, 51, 97–100. [Google Scholar] [CrossRef]
  91. 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 Environ. 2018, 227, 229–239. [Google Scholar] [CrossRef]
  92. Li, C.; Wang, H.; Ming, J.; Liu, M.; Fang, P. Hydrogen generation by photocatalytic reforming of glucose with heterostructured CdS/MoS2 composites under visible light irradiation. Int. J. Hydrog. Energy 2017, 42, 16968–16978. [Google Scholar] [CrossRef]
  93. Speltini, A.; Scalabrini, A.; Maraschi, F.; Sturini, M.; Pisanu, A.; Malavasi, L.; Profumo, A. Improved photocatalytic H2 production assisted by aqueous glucose biomass by oxidized g-C3N4. Int. J. Hydrog. Energy 2018, 43, 14925–14933. [Google Scholar] [CrossRef]
  94. Xu, X.; Zhang, J.; Wang, S.; Yao, Z.; Wu, H.; Shi, L.; Yin, Y.; Wang, S.; Sun, H. Photocatalytic reforming of biomass for hydrogen production over ZnS nanoparticles modified carbon nitride nanosheets. J. Colloid Interface Sci. 2019, 555, 22–30. [Google Scholar] [CrossRef] [PubMed]
  95. Nguyen, V.-C.; Ke, N.-J.; Nam, L.D.; Nguyen, B.-S.; Xiao, Y.-K.; Lee, Y.-L.; Teng, H. Photocatalytic reforming of sugar and glucose into H2 over functionalized graphene dots. J. Mater. Chem. A 2019, 7, 8384–8393. [Google Scholar] [CrossRef]
  96. Jing, D.; Liu, M.; Shi, J.; Tang, W.; Guo, L. Hydrogen production under visible light by photocatalytic reforming of glucose over an oxide solid solution photocatalyst. Catal. Commun. 2010, 12, 264–267. [Google Scholar] [CrossRef]
  97. Zhang, J.; Xu, Q.; Feng, Z.; Li, M.; Li, C. Importance of the Relationship between Surface Phases and Photocatalytic Activity of TiO2. Angew. Chem. Int. Ed. 2008, 47, 1766–1769. [Google Scholar] [CrossRef]
  98. Yasuda, M.; Misriyani; Takenouchi, Y.; Kurogi, R.; Uehara, S.; Shiragami, T. Fuelization of Italian Ryegrass and Napier Grass through a Biological Treatment and Photocatalytic Reforming. J. Sustain. Bioenergy Syst. 2015, 5, 1–9. [Google Scholar] [CrossRef] [Green Version]
  99. Li, Z.; Meng, X.; Zhang, Z. Fabrication of surface hydroxyl modified g-C3N4 with enhanced photocatalytic oxidation activity. Catal. Sci. Technol. 2019, 9, 3979–3993. [Google Scholar] [CrossRef]
Figure 1. Distribution of feedstock for hydrogen production (Adapted from ref. [13]).
Figure 1. Distribution of feedstock for hydrogen production (Adapted from ref. [13]).
Catalysts 10 00335 g001
Figure 2. Photocatalytic production of H2.
Figure 2. Photocatalytic production of H2.
Catalysts 10 00335 g002
Figure 3. Reaction pathways of photocatalytic reforming of methanol on Au/TiO2 catalyst (Adapted from Ref. [36]).
Figure 3. Reaction pathways of photocatalytic reforming of methanol on Au/TiO2 catalyst (Adapted from Ref. [36]).
Catalysts 10 00335 g003
Figure 4. Products of photocatalytic reforming of methanol on different photocatalyst (no shading: H2, diagonal shading: CO2 and vertical shading: CO, adapted from Ref. [42]).
Figure 4. Products of photocatalytic reforming of methanol on different photocatalyst (no shading: H2, diagonal shading: CO2 and vertical shading: CO, adapted from Ref. [42]).
Catalysts 10 00335 g004
Figure 5. Scheme for preparation of Au @ TiO2 core-shell composite. (Adapted from Ref. [73]).
Figure 5. Scheme for preparation of Au @ TiO2 core-shell composite. (Adapted from Ref. [73]).
Catalysts 10 00335 g005
Figure 6. Pathways of glucose in photocatalytic reforming (Adapted from Ref. [85]).
Figure 6. Pathways of glucose in photocatalytic reforming (Adapted from Ref. [85]).
Catalysts 10 00335 g006
Figure 7. Effect of the kinds of noble-metal-loaded TiO2 on hydrogen evolution. (Adapted from Ref. [85]).
Figure 7. Effect of the kinds of noble-metal-loaded TiO2 on hydrogen evolution. (Adapted from Ref. [85]).
Catalysts 10 00335 g007
Figure 8. AQE for photocatalytic reforming of biomass on NiS-CdS composites (black column: lactic acid; red column: lignin acid) (Adapted from Ref. [91]).
Figure 8. AQE for photocatalytic reforming of biomass on NiS-CdS composites (black column: lactic acid; red column: lignin acid) (Adapted from Ref. [91]).
Catalysts 10 00335 g008
Figure 9. The time courses of photocatalytic hydrogen production over BYV (x = 0.5) under visible light at different pH. (Adapted from Ref. [96]).
Figure 9. The time courses of photocatalytic hydrogen production over BYV (x = 0.5) under visible light at different pH. (Adapted from Ref. [96]).
Catalysts 10 00335 g009
Table 1. Summary of photocatalytic reforming of biomass for hydrogen evolution.
Table 1. Summary of photocatalytic reforming of biomass for hydrogen evolution.
CatalystBiomassEfficiencyLightRef
Pt-TiO2-W1.25 mmol/L glucose1.0 mmol h−1 gcat−1sunlight[88]
Pt/P2520 vol% methanol, 20 vol.% propanetriol, 1.3 mmol/L glucose0.749, 7.094, 7.784 mmol h−1 gcat−1300 W Xe lamp[89]
Pd/TiO2250 mg/L glucose5.7 mmol h−1 gcat−1125 W high-pressure mercury lamp[85]
Rh/TiO21.25 mmol/L glucose1.5 mmol h−1 gcat−1300 W Hg lamp[90]
NiS/CdS0.10 g/L lignin & 2.0 vol.% lactic acid1.5 (both lignin and lactic acid) & 1.086 (only lactic acid) mmol h−1 gcat−1300 W Xe lamp[91]
CdS/MoS20.1 mol/L glucose55.0 mmol h−1 gcat−1300 W Xe lamp [92]
O-g-C3N41 mol/L glucose1.37 mmol h−1 gcat−1simulated solar light[93]
ZnS-g-C3N450ppm glucose69.8 μmol h−1 gcat−1300 W Xe lamp[94]
S, N-GO0.35 mol/L Sugar & 0.35 mol/L glucose12.5 & 8.75 μmol h−1 gcat−1300 W Xe lamp [95]
Bi0.5Y0.5VO41 mol/L glucose50μmol h−1 gcat−1350 W Xe lamp [96]

Share and Cite

MDPI and ACS Style

Yao, Y.; Gao, X.; Li, Z.; Meng, X. Photocatalytic Reforming for Hydrogen Evolution: A Review. Catalysts 2020, 10, 335. https://doi.org/10.3390/catal10030335

AMA Style

Yao Y, Gao X, Li Z, Meng X. Photocatalytic Reforming for Hydrogen Evolution: A Review. Catalysts. 2020; 10(3):335. https://doi.org/10.3390/catal10030335

Chicago/Turabian Style

Yao, Yuan, Xinyu Gao, Zizhen Li, and Xiangchao Meng. 2020. "Photocatalytic Reforming for Hydrogen Evolution: A Review" Catalysts 10, no. 3: 335. https://doi.org/10.3390/catal10030335

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

Article Metrics

Back to TopTop