Sunesh S.
Mani
a,
Sivaraj
Rajendran
a,
Thomas
Mathew
*a and
Chinnakonda S.
Gopinath
*bc
aDepartment of Chemistry, St. John's College (Affiliated to University of Kerala), Anchal, Kerala – 691306, India. E-mail: thomasm74@gmail.com
bCatalysis and Inorganic Chemistry Division, CSIR – National Chemical Laboratory, Dr Homi Bhabha Road, Pune 411 008, India. E-mail: cs.gopinath@ncl.res.in
cAcademy of Scientific and Innovative Research (AcSIR), Ghaziabad 201002, India
First published on 10th June 2024
The major issues that determine the efficiency of photocatalyst composite materials for solar hydrogen production, with or without a sacrificial agent, are efficient visible light harvesting properties, efficient separation of charge carriers and their utilization of redox sites, and stability. Thus, significant efforts have been devoted in the past few decades to modify the above characteristics by integrating constituent components of composites using different approaches. In the present review, we aim to summarize the recent advances, predominantly, in the area of TiO2-based photocatalyst composites for solar hydrogen production. Firstly, we present the recent progress in material integration aspects by discussing the integration of TiO2 with different categories of materials, including noble/3d metals, metal oxides/sulphides/selenides, other low bandgap semiconductors, C-based materials, and dye sensitizers. Furthermore, we discuss how material integration helps in tailoring the electronic and optical properties for activity tuning in solar H2 production. Subsequently, critical changes in the physico-chemical and electronic properties of composites with respect to their preparation methods, morphology, crystallographic facets, particle size, dopant, calcination temperature, and structure–activity relationship to solar hydrogen production are addressed in detail. Moreover, we discuss the importance of fabricating a photocatalyst in a thin film form and performing solar hydrogen production in different reactor set-ups for enhancing its photocatalytic performance, while addressing device scalability. Despite the significant advancements made in this field, solar-to-hydrogen conversion efficiency still needs to be improved to realise the practical application of solar hydrogen production. In this case, the direct conversion of water to hydrogen via overall water splitting and renewable H2 production from wastewater or biomass components by employing suitable photocatalysts are some possible ways to improve the energy efficiency, and continuous research in the above directions is highly desirable.
Since the potential of TiO2 for the photolysis of water was revealed by Fujishima and Honda for the first time in 1972,6 its performance has been widely explored for a variety of applications, such as photocatalytic degradation of pollutants,7 supercapacitors,8 solar cells,9 carbon dioxide reduction,10 lithium-ion batteries,11 biomedical devices,12 self-cleaning,13 and water splitting.14 However, the wide bandgap (3.2 eV) of TiO2 restricts its light-harvesting ability to mainly the UV region (which is ∼5% of sunlight), and the fast recombination of photo-generated electron–hole pairs in bare TiO2 limits its photocatalytic functionality.6 Thus, to overcome the above-mentioned inherent drawbacks of TiO2 and enhance its photocatalytic performance for solar H2 production, significant efforts have been devoted to designing catalysts, such as doping with metals and non-metals,15,16 dye sensitization,17 use of noble metals (Pt, Pd, Au and Ag) as a co-catalyst,18 engineering the band structure to match particular energy levels,19 and fabrication of semiconductor heterojunction and/or Schottky junctions.20
In the past few years, several excellent review articles have been published on photocatalytic water splitting for hydrogen production, which are based on different catalyst systems including semiconductor-based catalyst systems,21,22 metal-free photocatalysts,23 spinel materials,24 ionic carbon nitride,25–28 carbon-based materials,29 transition metal complexes,30 and TiO2-based semiconductor materials.31–36 Also, although there are few reviews available on TiO2-based photocatalysts for solar hydrogen production,31–36 they emphasized the general aspects of photocatalytic H2 production. Thus, readers interested in the above-mentioned aspects can refer those references. In contrast, the present review emphasizes the structure–activity correlation and how material integration is beneficial for tuning the activity of photocatalysts. The various critical changes in the physico-chemical and electronic properties of TiO2-based materials fabricated by integrating TiO2 with a variety of dopants and/or materials and its structure–activity relation for solar hydrogen production are addressed in detail. Also, the efficient and concurrent utilization of both photogenerated holes and electrons to improve the photocatalytic efficiency of materials is discussed. Furthermore, we discuss the importance of fabricating photocatalysts in thin film form and performing solar hydrogen production in different reactor set-ups for enhancing the photocatalytic performance of materials and their scalability.
A Scopus-based survey was performed using the following three sets of keywords to determine the number of publications related to TiO2-based photocatalysts for water splitting, as follows: (i) TiO2 photocatalysts and hydrogen production, (ii) TiO2 photocatalysts and water splitting, and (iii) TiO2-based photocatalysts, and hydrogen production and water splitting. This survey revealed that the articles published in this area have been steadily increasing in the past ten years and the result is shown in Fig. 1. As shown in Fig. 1, an increase in the number of articles on TiO2-based photocatalysts for H2 production appeared after 2017, highlighting the importance of sustainable hydrogen production from renewable sources and the worldwide focus on it.
Several variations and advances have been made in the past decade, especially regarding TiO2-based semiconductor photocatalysts for solar hydrogen production using water as the main hydrogen source. Therefore, in this review, we comprehensively summarize and highlight the recent progress in the design and structure–activity relation of TiO2-based photocatalysts towards solar hydrogen production with water as the main hydrogen source, where the works employing sacrificial agents are also included to a significant extent. Initially, we provide the fundamental aspects of integrating different materials with TiO2, such as noble metals, non-noble metals, metal oxides/sulphides/selenides, and low bandgap semiconductors and their structure–activity correlation is highlighted to understand the mechanistic aspects and the entire progress in the field of photocatalytic hydrogen production. Also, the role of TiO2-based photocatalysts is specifically reviewed based on the effects of various factors including preparation methods, morphology, crystallographic facet-dependent activity, catalyst loading, and dopant concentration. Subsequently, we discuss the importance of engineering strategies, such as the importance of fabricating photocatalysts in thin film form to overcome the drawbacks of TiO2 in powder form, resulting in improved photoactivity and scalability. Finally, the major challenges and an outlook on the future strategies in this research field are discussed from the viewpoint of the structure–activity relation of various TiO2-based photocatalysts.
TiO2 + hv → e(CB-TiO2)− + h(VB-TiO2)+ | (1) |
2H2O + 4h+ → O2 + 4H+ | (2) |
4H+ + 4e− → 2H2 | (3) |
Fig. 2 Flowchart illustrating the fundamental and critical processes involved in semiconductor based photocatalytic water splitting processes. |
The bandgap and band position of the photocatalyst,3 the ability of the photocatalyst in enhancing the charge (electron–hole pair) separation and charge diffusion to the redox sites,38–40 presence of sacrificial reagents,41etc. are crucial factors that determine the efficiency of solar hydrogen production. The conversion efficiency of solar energy into hydrogen can be calculated by the apparent quantum yield (AQY) according to eqn (4).42
(4) |
Fig. 3 Schematic representation of different types of chemical components used to improve the photocatalytic efficiency of TiO2. |
Noble metals including Ru,79 Rh,80 Pd,44,48,81 Au,49,82–84 Ag,85–87 and Pt50,88–90 are widely used as a co-catalyst with TiO2 in photocatalytic hydrogen production due to their unique properties such as formation of a Schottky barrier,43,91 efficient interfacial electron transfer, photostability, ability to show surface plasmon resonance,92 formation of impurity energy levels,79 and oxygen vacancies.80 There are two types of noble metals, plasmonic (Ag and Au) and non-plasmonic (Pt, Pd, Ru, and Rh).48,76 In this case, a Schottky barrier is more critical for non-plasmonic metals such as Pt and Pd, which functions as an effective electron trap, imparting a high density of states to the Fermi level and facilitating charge separation and utilization. For example, non-plasmonic noble metals are not capable of exhibiting visible light absorption; however, they show very high cocatalyst activity.66,76 Noble metals such as Au and Ag on the surface of TiO2 increase the photon absorption in the visible region through the localised surface plasmon resonance (LSPR) effect.93,94 The surface plasmon resonance (SPR) effect increases the production of hot electrons at the interface between the metal and TiO2. Given that the SPR level of the metal is higher than the conduction band of TiO2, electrons are transferred from the SPR level of the metal to the conduction band of TiO2. At the same time, the Schottky barrier can prevent the back transfer of electrons to metal nanoparticles.49 The formation of an intermediate energy level enables charge separation in Ru and extends the light absorption to the visible region,79 whereas Rh induces the generation of oxygen vacancies in TiO2, which allows the withdrawal of electrons from the metal, leading to an enhancement in hydrogen production activity.80 A schematic illustration of the LSPR effect occurring on the surface of a plasmonic metal nanoparticle and H2 production augmented by electron transfer via the Schottky barrier at the interface between a metal semiconductor, such as Au nanoparticle, is shown in Fig. 4(a and b).
Fig. 4 (a) Schematic representation of LSPR effect showing the oscillation of an electric field of incident light at the resonance frequency on the surface of an Au nanoparticle. (b) Schematic diagram of the proposed mechanism for photocatalytic H2 production from ethanol/water mixtures over Au/P25 TiO2. Reproduced with permission from ref. 92. Copyright©2015, Wiley-VCH. |
Although noble metals are highly efficient as co-catalysts for photocatalytic hydrogen generation, their high cost and low abundance limit their application, leading to the investigation of earth abundant and cheap non-noble metals co-catalysts.95–97 Non-noble metal nanoparticles such as Cu and Ni as co-catalysts with TiO2 have been found to be highly effective in enhancing the rate of hydrogen production in the water splitting reaction.16,98–102 Heterogeneous surface-distributed non-noble metal nanoparticles on TiO2 increase charge separation and facilitate charge transfer.103 Copper is one of the most studied non-noble metal cocatalysts for solar hydrogen evolution because of its low cost, high conductivity, capability to show SPR effect, etc. The presence of metallic copper substantially changes the electronic structure of TiO2 because of the formation of structural defects or energetic electron trap centres, which prevents electron–hole recombination.95,98,104 Nickel is another attractive non-noble metal co-catalyst because of its high work function, availability and low cost. The high work function of Ni is favourable for preventing the migration of electrons back into the conduction band of TiO2 in Ni/TiO2.105 Metallic Ni nanoparticles dispersed on TiO2 as small-sized clusters resulted in an improved photocatalytic hydrogen evolution rate.105 A schematic description of the properties related to various metal nanoparticles employed for fabricating metal-integrated TiO2 photocatalyst systems for photocatalytic H2 production is provided in Fig. 5.
Fig. 5 Schematic representation of various noble metals used for improving the photocatalytic hydrogen evolution activity of TiO2. |
Fig. 6 (a) Schematic synthesis of Cu/Ag bimetallic deposition on titanium nanotubes (TNTs) and (b) charge separation and hydrogen evolution mechanism of as-synthesized Cu/Ag@TiO2. Reproduced with permission from ref. 99. Copyright©2017, Elsevier Ltd. |
Similarly, NiCu alloy was also found to enhance the solar hydrogen production with a quasi-artificial leaf device (QuAL) fabricated using Mn-doped CdS integrated with mesoporous TiO2.106 Mn-doping in CdS was shown to enhance the visible light absorption to a longer wavelength compared to virgin CdS. In the presence of the NiCu-alloy co-catalyst, the QuAL device with an area of 1 cm2 was demonstrated to produce 10.5 mL h−1 of H2 with a power conversion efficiency of 4.8%. Another interesting aspect of this QuAL device is the increase in the visible light absorption efficiency due to different-size integrated Mn-CdS, which exhibited a wide range of band-gap values due to the different QD sizes. The fluorescence is emitted from the smaller QDs and apparently absorbed by the significantly larger QDs or the same Mn-CdS nanoparticles.
Fig. 7 (a) UV-vis absorbance spectra and (b) Tauc plot obtained from the reflectance spectra and (c) valence band edge analysis of TiO2, N-TiO2-450, N-TiO2-550. (d) Proposed band diagrams of TiO2 and N-doped TiO2 catalysts. Reproduced with permission from ref. 115. Copyright©2016, Elsevier Ltd. |
The electron hole separation and electron transfer process through the formation of a heterojunction with a low band gap semiconductor depend on the band position of both TiO2 and the supporting semiconductor. If the CB edge potential of the supporting semiconductor is more negative than that of TiO2, the photogenerated electrons in the conduction band of the semiconductor usually transfer to the lower-lying conduction band of TiO2.54,55,123 In contrast, the CB edge potential of the supporting semiconductor is less negative than that of TiO2, and the electrons from the conduction band of TiO2 will be transferred to the lower-lying conduction band of the supporting semiconductor.20,63,124,125
Fig. 9 Schematic representation of various metal oxides used for improving the photocatalytic hydrogen evolution activity of TiO2. |
The activity of TiO2 can be further improved by surface modification with Ni (metal) and P (nonmetal), resulting in more effective heterojunctions.55 For example, after surface decoration with Ni–P, the CeO2–TiO2 photocatalyst became an effective photocatalyst for hydrogen evolution due to the lowering of its bandgap to 2.4 eV and powerful electron–hole separation.55 In this photocatalyst, Ni–P acts as a bridge for electron transportation. Also, SiO2 is an interesting substrate used for dispersing and maintaining the active phase, imparting thermal stability and improving the surface area of TiO2-based photocatalysts.58 For example, the pore channels of SiO2 were successfully utilized for anchoring TiO2 quantum dots (TiO2-QDs) in a highly dispersed state on the surface of silica via the in situ hydrolysis of Ti-alkoxide.58 The anchoring effect between TiO2-QDs and the pore-wall of SiO2 foam provided extra stability to TiO2-QDs, which helped in maintaining TiO2-QDs in the anatase phase without undergoing any phase transformation and particle growth even after exposure to high temperature of up to 1000 °C.
NiO is a p-type semiconductor with a bandgap of 3.5 eV, which possesses high charge carrier concentration and high mobility of charge carriers. It can form an efficient p–n heterojunction with anatase TiO2, which facilitates the red shifting of the bandgap energy of the NiO/TiO2 heterostructure.123 The internal electric field developed at the interface of the NiO/anatase TiO2 p–n heterojunctions facilitates the dissociation efficiency of photogenerated electron–hole pairs and enhances the photocatalytic efficiency. For example, mesoporous heterostructure of 1 wt% NiO on anatase TiO2 nanoparticulate photocatalyst showed an H2 production rate of 2693 μmol h−1 g−1 by methanol photo-reforming, which is higher than that of pure anatase TiO2 and commercial P25.
Fig. 10 (a) Schematic illustration of the assembly process of the MoS2@TiO2 heterostructure. (b) SEM pictures of MoS2(10)@TiO2, MoS2(20)@TiO2 and MoS2(30)@TiO2, respectively (scale bar: 100 nm). (c) TEM and HRTEM images of MoS2(10)@TiO2 and MoS2(20)@TiO2 (scale bar: 20 nm). (d) Cross-sectional HAADF-STEM images and EDS elemental mapping of the MoS2@TiO2 heterostructure (first row, scale bar: 200 nm, and second row, scale bar: 50 nm). (e) Schematic diagram of the energy band structure, plasmonic resonance and electron transfer pathway in the MoS2@TiO2 heterojunction. (f) Band structure of monolayer 2H-MoS2 with 0%, 8% and 16% S-vacancies, with the valence band maximum and conduction band minimum both at the K point. (g) Model used for band gap computation via DFT. (h) FEM simulation of the near-field electric field distribution inside MoS2@TiO2 heterostructures excited by a 400 nm laser and (i) 3D-simplified model used for FEM simulation. Reproduced with permission from ref. 139. Copyright ©2018, The Royal Society of Chemistry. |
Fig. 11 (a) Schematic illustration of the interfacial charge transfer process in FeS2–TiO2 catalyst under UV, visible and near-IR light irradiation. (b) HRTEM image showing the intimate contact between the (210) plane of FeS2 and the (101 plane) of TiO2 in FeS2–TiO2. (c) Rate of hydrogen evolution from TiO2, FeS2 and the FeS2–TiO2 nanocomposite under a 300 W xenon lamp. Reproduced with permission from ref. 64. Copyright ©2018, the Royal Society of Chemistry. |
Fig. 12 (a) Schematic illustration of the synthesis of TCNx composites. (b) UV-vis absorption spectra of TiO2 and TCNx composites. SEM images of (c) g-C3N4 NS and (d) TCN50. (e) and (f) HRTEM images of TCN50. (g) Proposed mechanism for visible light photocatalytic activities of TCNx composites. (h) VB XPS of TiO2 and g-C3N4 NS. Reproduced with permission from ref. 151. Copyright ©2017, the Royal Society of Chemistry. |
Fig. 13 Mechanism for photocatalytic H2 evolution over eosin Y-sensitized C-TiO2 hollow nanoshells with Mie resonance. Reproduced with permission from ref. 74. Copyright ©2021, Elsevier Ltd. |
Fig. 14 Schematic illustration of the mechanism of photocatalytic hydrogen evolution from an Au–Pd/rGO/TiO2 triple composite, in which an SPR electron from Au is transferred to TiO2via rGO. Reproduced with permission from ref. 66. Copyright ©2019, ACS. |
No. | Material | Experimental conditions | H2 yield/mmol h−1 g−1 | Ref. |
---|---|---|---|---|
Metal-integrated TiO2 | ||||
1 | 0.5Rh/TSG (TiO2 sol–gel) | 50 mg of 0.5Rh/TSG photocatalyst in aqueous solution of ethanol (50% v/v, water), UV Pen-Ray Hg lamp (λ = 365 nm) | 7.2246 | 51 |
2 | Ru 8.0/TiO2 NBs-400 | 10 mg of photocatalyst was suspended in 10 mL of water containing 0.1 g of ethylenediamine tetra acetic acid disodium salt (EDTA-2Na), 300 W xenon lamp | 25.34 | 47 |
3 | TiO2Ag-F | 200 mg catalyst in 200 mL of distilled water, 4 mL of methanol, 250 W mercury lamp | 0.180 in 4 h | 38 |
4 | TiO2–Au-1% Pt | 10 mg of TiO2–Au photocatalyst into an aqueous methanol solution (120 mL, 25 vol%), 300 W Xe Lamp UV cut-420 nm filter | 0.0924 | 49 |
5 | 3 wt% Au/P25 TiO2 | (6.5 mg) was placed in ethanol (15 mL) and Milli-Q water (3.75 mL), 200 W, 365 nm with 6.5 mW cm2 | 31.5 | 92 |
6 | Ag@TiO2 | 0.05 g catalyst was suspended in deionized water and methanol mixed solutions (40 mL, 3:1), 300 W xenon lamp | 0.5319 | 164 |
7 | Pd/P25 1 wt% of Pd | Thin film forms of photocatalysts (1 mg) with 25% v/v aqueous methanol solution, direct sunlight 50.2 mW cm−2 | 104 | 165 |
8 | Au(2%)@TiO2 | 15 mg of photocatalyst was dispersed in 25 mL total volume containing 20% methanol (v/v) in aqueous solution, natural sunlight | 3.99 μmol h−1 with 100 mg | 166 |
9 | The TiO2–Pd NSs | 15 mg photocatalyst was dispersed in 50 mL of methanol/H2O mixture (20 vol% methanol), 300 W Xe UV lamp and vis-NIR light were used as illumination source, 400 nm cutoff filter 2.7 and 100 mW cm−2 | 2.80 | 167 |
10 | Au@TNT | 50 mg of photocatalyst was added to 100 mL of methanol-deionized water mixed solution (VMeOH/VH2O = 1:10), Xe lamp (λ > 400 nm), intensity of incident visible light (400 nm) was 0.1 W cm−2 | 0.482 | 168 |
11 | 1.5% Ag/TiO2 | 20% (v/v) aqueous solution of Na2S + Na2SO3, 254 nm wavelength UV light, intensity of light was 4.40 mW cm−2 | 23.496 | 112 |
12 | TiO2–Au 9 wt% | 30 mg of catalyst was dispersed in 150 mL of water/methanol mixture with the volume ratio of 9:1, 300 W xenon arc lamp intensity was 380 mW cm−2 | 12.440 | 169 |
13 | Pd/TiO2 nanosheets (0.18 At% Pd NPs) | 50 mg of photocatalyst in 50 mL of an aqueous solution containing 20% methanol in volume, 300 W Xe lamp equipped with a cut-off filter i (λ > 420 nm), light intensity (50 mW cm−2) | 3.096 | 170 |
14 | 0.1 mol% Ru–TiO2 | 0.5 g of fine powder photocatalyst was suspended in 550 mL water and 50 mL methanol, 500 W Hg mid-pressure immersion lamp λ > 320 nm | 3.400 | 79 |
15 | 1 wt% Ag/TiO2(TiAg-1) | 1 mg photocatalyst was dispersed in 1 mL ethanol drop-casted on a thin film, 25% (v/v) methanol/water mixture, direct sunlight | 4.59 | 43 |
16 | TiO2/Cu | 20 mg of TiO2 and a given amount of metal particles were added to 30 mL, 15 W black light with emission of ∼352 nm; 1.0 mW cm−2 | 0.850 mmol g−1 in 3 h | 171 |
17 | Cu/Ag@TNT 0.1 M Cu and 0.1 M Ag | 5 mg of catalyst was dispersed in 50 mL of 5 vol% glycerol aqueous solution, 300 W Xe lamp with a UV cutoff filter with wavelength of >400 nm | 56.167 | 99 |
18 | TiO2–Ni-1% | 50 mg photocatalyst was added to methanol solution (20 vol%), 300 W Xe lamp | 1.433 | 16 |
19 | Cu(3%)–TiO2/ErB | 5 mg of catalyst in 70 mL water containing 10 vol% triethanolamine, 300 W Xe lamp 0.15 W cm−2 | 13.4 | 104 |
20 | 0.5 wt% Ni/P25 TiO2 | Photocatalyst (6.5 mg) was placed in the reactor and 20 mL of an aqueous glycerol mixture (10 vol%), 100 W, 365 nm at a distance of 10 cm from the reactor. The photon flux at the sample was 6.5 mW cm−2 | 26.0 | 101 |
21 | Ni-a/TiO20.46 wt% | 50 mg of the sample disperses in 100 mL of 10 vol% methanol aqueous solution, 300 W Xe lamp | 0.0945 mmol h−1 with 50 mg | 172 |
22 | TiO2–NT/Pd–ND | 1 mg catalyst in H2O/MeOH solution (2:1), 300 W Xe lamp, maximum intensity of 203.3 mW cm−2 at wavelength of 365 nm | 0.143 mmol h−1 with 1 mg | 173 |
TiO2 with non-metals | ||||
23 | 15 wt% S-modified TiO2/β-SiC | 0.05 g of catalyst was suspended in 50 mL of an aqueous solution containing 10 vol% of methanol solution, 125 W medium pressure Hg visible lamp was used, 1 M NaNO2 solution as a UV filter under visible light irradiation (λ ≥ 400 nm) | 1.254 | 45 |
24 | N-doped TiO2 hollow fibres nitrogen up to 5 at% (N-TiO2) | Photocatalyst sample (0.02 g) was immersed in an aqueous methanol solution (5 mL methanol and 20 mL deionized water), 150 W Xe lamp optical filter, which allowed only wavelengths higher than 420 nm | 0.185 μmol g−1 in 6 h | 108 |
25 | TiO2/C | 5 mg catalyst powder in aqueous solution (10 mL) consisting of triethanolamine (10 vol%), 300 W Xe lamp, cut-off filter (λ > 400 nm) | 57.2 μmol h−1 with 5 mg | 174 |
TiO2 with metal oxide semiconductor | ||||
26 | 5 wt% Fe2O3/TiO2 | 50 mg Fe2O3/TiO2 nanocomposite was suspended with magnetic stirring in 200 mL aqueous solution 10 vol% glycerol, 500 W Xenon lamp with UV cutoff filter (λ > 420 nm) | 1.0 mmol h−1 g−1 | 54 |
27 | Ni–P/CeO2–TiO2 | 0.5 g of the prepared catalyst was directly dispersed in 20 mL deionized water, 1 Sun (1000 mW cm−2) was applied by a solar simulator | 1.30 mmol in 5 h with 0.5 g | 55 |
28 | Pt/NiO–TiO2 | 0.1 g of catalyst was suspended in water + methanol mixtures (33% v/v, 15 ml), 400 nm, λmax = 536 nm (medium-pressure Hg lamp, 400 W) and UV-vis light source (Hg lamp, 400 W) | 0.537 mmol g−1 h−1 | 56 |
29 | TiO2-QDs/SiO2 | 50.0 mg of TiO2/SiO2 composite photocatalyst was suspended in 80.0 mL of aqueous solution containing methanol (25.0 vol%), UV-LEDs (3 W, 365 nm) | 10.399 mmol g−1 h−1 | 58 |
30 | TiO2–ZnO-(0.6%) | 30 mg catalyst powder was ultrasonically dispersed into 80 mL H2O, and then 20 mL methanol, 300 W Xe lamp, the light intensity was kept at around 244 mW cm−2 | 0.3135 mmol h−1 with 30 mg | 59 |
31 | 2.82 wt% Ag2O·TiO2 | 0.2 g catalyst was suspended in 100 mL glycerol aqueous solution (with 7 vol% of glycerol), 300 W Xe arc lamp (320–780 nm) | 336.7 μmol h−1 g−1 | 131 |
32 | 1 wt% NiO/anatase TiO2 | (50 mg) and the mixture of methanol/H2O (1:1 v:v, 10 mL), UV-light (Hg vapor light source (LUMATEC SUPERLITE 400)) | 2.693 | 123 |
33 | 25 wt% of Y2O3 in TiO2 | Photocatalyst (10 mg) was dispersed in a 7.5 mL aqueous solution. After sonication, 2.5 mL ethanol was added as a sacrificial reagent, 300 W Xenon/Mercury lamp | 1.380 in 2.5 h | 124 |
34 | 1 wt% Cu/TiO2 | Photocatalyst (2.5 mg) was loaded in the reactor containing 25 mL of a glycerol–water mixture (5 vol% glycerol), 100 W, 365 nm UV light, excitation at ∼6.5 mW cm−2 | 15.32 | 175 |
35 | 2.5-Cu2O/TiO2 | 450 mL aqueous solution, 10 vol% ethylene glycol as a scavenger in reaction solution, 500 W Xe arc lamp bandpass filter (λ = 365 nm with the photon flux of 3.6 mW cm−2) | 2.048 | 176 |
36 | TiO2–(0.1 wt%)CuO | 0.1 g of the sample was suspended in 1 M KOH, 300 W xenon lamp | 2.715 in 5 h | 177 |
37 | Cu2O/TiO2 | 0.1 g of the Cu2O/TiO2 photocatalyst was dispersed in 100 mL 10 vol% aqueous methanol, 300 W xenon lamp with or without a bandpass filter (λ > 420 nm) | 24.83 | 178 |
TiO2 with metal chalcogenides | ||||
38 | S-doped hetero nanostructured TiO2/Cu2S | Aqueous solution containing 0.35 M Na2SO3 and 0.35 M Na2S, visible light irradiation conditions (λ > 420 nm) | 1.280 with 50 mg | 6 |
39 | NiSe/TiO2 | 50 mg of the as-prepared photocatalyst powder in 100 mL aqueous solution containing 10 vol% methanol | 55.4 μmol h−1 | 20 |
40 | FeS2–TiO2 | 1 g catalyst (50% aqueous methanol solution), mercury arc lamp (400 W) | 0.331 | 64 |
41 | MoS2/TiO2 (0.14 wt%) | 50 mg photocatalyst was suspended in 80 mL solution containing 20 mL methanol, four low-power LEDs (3.5 W cm−2, λ = 365 nm) | 2.443 | 125 |
42 | 10 mmol NiS/TiO2 nanosheet films | 80 mL 10% ethanol/H2O (v/v) solution, 500 W Xe lamp | 4.31 μmol cm−2 in 3 h | 135 |
43 | MoS2/TiO2 | MoS2@TiO2 films with a size 7 × 7 mm were submerged in 15 mL of mixed solution made of DI water (seawater) and methanol (8:2 by volume), solar light simulator (AM 1.5, 300 W Xe, 100 mW cm−2) | 580 | 139 |
44 | 0.50 wt% MoS2, 2D-2D MoS2/TiO2 | 100 mg photocatalyst in 100 mL aqueous solution containing 10% methanol in volume, 300 W Xe-arc lamp | 2.145 | 137 |
45 | 2D/1D TiO2 nanosheet/CdS nanorods | 50 mg photocatalyst was suspended in aqueous solution (20 mL lactic acid, 210 mL water), 300 W Xenon arc light source after filtering the UV light with circulating cooling NaNO2 aqueous solution (1 M) to pass only visible light (λ > 400 nm) | 128.3 | 179 |
TiO2 with carbonaceous material | ||||
46 | g-C3N4-TiO2 (1:4) | 75 mg of photocatalyst was dispersed in 75 mL of aqueous solution containing 10% triethanolamine, 250 W visible light source | 1.041 | 68 |
47 | TiO2-100-G | Photocatalyst (100 mg) was dispersed in an aqueous solution containing H2O (80 mL) and CH3OH (20 mL), 300 W Hg lamp with a wavelength of approximately 365 nm | 1.93 | 180 |
48 | TiO2/g-C3N4 | 15 mg sample dispersed in 8 mL ethanol and 72 mL H2O, 150 W Xenon lamp with a cutoff filter (λ > 420 nm) | 0.35 | 181 |
TiO2 with dye sensitization | ||||
49 | C-TiO2 hollow nanoshells with Eosin Y sensitization | Under visible light irradiation (λ > 420 nm) | 0.468 | 74 |
50 | Ru(dcbpy)3/TiO2 | 50 mL of solution containing 20 mg of photocatalyst and EDTA (2 mmol L−1), irradiated by a xenon lamp (150 W) with a UV cutoff filter (λ > 400 nm) | 94 μmol g−1 in 5 h | 17 |
51 | Carbazole-based organic dye-sensitized Nafion-coated Pt/TiO2 system (D1@NPT) | 20 mL aqueous suspension containing 10 mg of the photocatalyst and 10 vol% of TEOA as SED, Xenon arc lamp (400 W) was used as a light source | 67.9 μmol h−1 | 154 |
52 | BE-Au(1 wt%)-TiO2(modified poly(benzothiadiazole) flake denoted as BE) | 30 mg catalyst was dispersed in 30 mL 10% vol TEOA aqueous solution, filters (λ = 420, 500 nm, etc.) are used | 781.2 μmol h−1–30 mg | 182 |
TiO2-ternary composite material | ||||
53 | Au–Pd/rGO/TiO2 | 25 mg of the catalyst was suspended in 40 mL of an aqueous methanol solution (25% v/v), (300 W xenon arc lamp) with AM 1.5 filter under 1 sun conditions (100 mW cm−2) | 21.50 | 66 |
54 | Pt0.5–Au1/TiO2 | (20 mg) of catalyst was added to 30 mL of distilled water and 10 mL methanol, 300 W power AM1.5 (100 mW cm−2) or >400 nm (92 mW cm−2) filter | 1.275 | 76 |
55 | 0.5 wt% Pt/0.1 wt% Cu/TiO2 | 20 vol% methanol–water, photocatalyst powder (14 mg) was mixed with distilled water (1.2 mL), deposited on 0.85–0.42 mm quartz beads (3 g), xenon arc lamp, 300 W, 40.0 mW cm−2 irradiation intensity | 27.2 | 117 |
56 | CuInS2/TiO2/MoS2 photocatalyst with 0.6 mmol g−1 CuInS2 and 0.5 wt% MoS2 | 50 mg photocatalyst was suspended in 250 mL of an aqueous solution containing 0.1 M Na2S and 0.1 M Na2SO3, 300 W Xenon lamp equipped with a UV cutoff filter (λ > 420 nm) | 1.034 | 120 |
57 | rGO/TiO2/ZIS (2.0 wt% rGO and 50 wt% ZIS) | 0.01 g of sample is added into 10 mL of aqueous solution containing 0.35 M of Na2S and 0.25 M of Na2SO3, with a light source (300 W Xe lamp) | 0.4623 | 142 |
58 | 0.5 wt% Au/(TiO2–g-C3N4) (95/5) | 1 L of Milli-Q water and triethanolamine (TEOA) (1 vol%), solar light | 1.750 | 157 |
59 | CuO(1%)–Co3O4 (0.05%)/TiO2 | 20 mg photocatalyst was suspended in 100 mL of pure water or methanol aqueous solution (30 vol%), 300 W Xe lamp, light intensity of 600 mW cm−2 | 0.273 with 20 mg | 159 |
60 | 2 wt% Cu2O/TiO2/Bi2O3 | 5% glycerol–water solution 0.05 g L−1 photocatalyst, natural solar light | 6.727 with 50 mg | 162 |
61 | Pt(0.02%)Co3O4(0.005%)(D)/N-TiO2 (PCNT(D)) | 100 mg of photocatalyst dispersed in 100 mL of an aqueous solution containing methanol (VH2O: VMeOH = 9:1), 300 W Xe lamp and a UV light cutoff filter (λ > 400 nm) with a light intensity of 380 mW cm−2 | 197 | 183 |
Fig. 15 Schematic representation of the various factors that determine the photocatalytic hydrogen evolution activity of TiO2-based photocatalysts. |
Fig. 16 Schematic representation of the various strategies for the synthesis of TiO2-based photocatalyst systems reported in the literature. |
Li et al. reported six distinct preparation methods for the fabrication of CuO/TiO2 photocatalysts and evaluated the performance of these differently prepared materials in the photocatalytic hydrogen evolution reaction.184 As illustrated in Fig. 17, among the methods, including CAD, CP, impregnation (EI, SWI, and SI) and SG, CuO/TiO2 prepared via the CAD method showed the highest rate of hydrogen evolution (3155.7 μmol g−1 h−1). The highest activity of the material prepared via CAD is due to the formation of TiO2 in the anatase/rutile mixed-phase. The advantages of anatase/rutile mixed-phase include enhanced electron–hole separation, transfer of electrons from rutile to anatase phase, and formation of catalytic hotspots at the interface of the anatase and rutile phases.
Fig. 17 Comparison of the photocatalytic hydrogen evolution rate on pure P25 and CuO/TiO2 samples produced by different preparation methods, namely chemical adsorption decomposition (CAD), composite precipitation (CP), ethanol impregnation (EI), simple wet impregnation (SWI), stepwise impregnation (SI), and sol–gel (SG). Reproduced with permission from ref. 184. Copyright ©2014, Springer. |
A comparison of the activity of the TiO2/CuO nanocomposites prepared via the ILHAM, hydrothermal, sol–gel and ionic liquid co-precipitation methods for photocatalytic H2 production was reported, and this study revealed that the TiO2/CuO nanocomposite produced via ILHAM showed the highest hydrogen evolution activity (8.670 mmol g−1 within a period of 2.5 h) under a xenon/mercury lamp using aqueous methanol.134 The highest activity of the TiO2/CuO nanocomposite obtained by the ILAHM method is attributed to the versatility of this synthesis strategy to control the morphology and other physico-chemical characteristics of the catalyst by utilizing the unique properties of ionic liquids such as thermal stability and low vapor pressure.134
A comparison of the photocatalytic activity of mesostructured Ag@TiO2 prepared by the in situ electrospinning method and electrospinning in combination with the photo-deposition approach revealed that Ag@TiO2 nanofibers prepared through the former method displayed a higher H2 yield of 531.9 μmol g−1 h−1 under one sun conditions with 25% aq. methanol.164 The high activity of the Ag@TiO2 nanofibers fabricated via the in situ electrospinning method is attributed to its relatively large BET surface area (39.8 m2 g−1) and smaller particle size (16.5 nm) than that prepared by the other method.164
Xiao et al. developed a new molten salt technique to synthesize atomically dispersed Ni species on TiO2, as illustrated in Fig. 18a for application in photocatalytic water splitting to produce H2, which exhibited 4-times higher activity (94.5 μmol h−1 for 50 mg of sample) than Ni/TiO2 prepared by the conventional impregnation method.172 The experimental and theoretical studies (Fig. 18b–d) revealed that the higher activity of Ni/TiO2 produced by the molten salt technique is due to the fact that it enabled the atomic distribution of Ni ions on TiO2, which favoured the formation of strong Ni–O bond and large number of oxygen vacancies. Consequently, this made the material highly integrated and was beneficial for efficient charge carrier utilization. The chemical bath deposition process is another versatile synthesis strategy employed in the literature for achieving the uniform deposition of metal species on TiO2, as reported by Chang et al., for the fabrication of Cu2O on mesoporous TiO2 beads (MTBs), which facilitated the essential charge separation for an improved hydrogen generation rate (223 mmol h−1 g−1).9
Fig. 18 (a) Schematic illustration of molten salt-mediated preparation of atomic Ni on TiO2. For the TiO2 molecular structure, (b) EXAFS and XANES spectra of Ni/TiO2, respectively. Ni loading amount: 0.46 wt% (c) Comparison of H2 evolution activity on Ni-np/TiO2, Ni-a/TiO2 and Pt/TiO2 with the same loading amount (∼0.5 wt%). Experimental condition: 50 mg of sample dispersed in 100 mL of 10 vol% methanol aqueous solution (278 K) under an irradiation of a 300 W Xe lamp. (d) Free energy versus the reaction coordinates of different active sites. The simulation is based on the (101) facet of anatase TiO2. (e) Molecular models of OV on Ni/TiO2 and the corresponding formation energy of OV. Reproduced with permission from ref. 172. Copyright© 2020, Wiley-VCH. |
The synthesis strategy leading to the formation of TiO2 with a sufficient amount of Ti4+ ions in tetrahedral coordination enhances the photocatalytic H2 production rate due to fact that the photocatalytic activity of TiO2-based catalysts depends on the quantity of Ti4+ ions present in a tetrahedral environment.185 Kumari et al. reported an impregnation method for the fabrication of functionalized carbon nanotubes (FCNTs) integrated with TiO2 nanotubes (TiNTs), which showed a high performance (7476 μmol h−1 g−1) for the photocatalytic hydrogen evolution reaction due to the strong interaction between FCNTs and TiNTs and the well-dispersed nature of the catalyst.67
Careful control of the synthesis parameters and process variation are essential given that even a slight variation in the synthesis process may introduce a variety of defects and impurities that could improve or worsen the photocatalytic activity and stability of materials. Generally, anatase-phase TiO2 exhibits higher activity than the rutile phase, but synthesis frequently results in mixed-phase materials, making property optimisation more challenging.
Fig. 19 Amount of hydrogen produced vs. surface area of photocatalyst for Au@TiO2:0.5Au-450, 1Au-450, 2Au-450 and 3A-450 under natural solar light. Adopted from ref. 166. Copyright ©2018, the Royal Society of Chemistry. |
In another variation, among different TiO2-NT (nanotube)-based photocatalysts (TiO2-NT, TiO2-NT/ZnO-NR (nanorod), TiO2-NT/ZnO-NR/Ag-NP (nanoparticle), TiO2-NT(nanotube)/Pd-NDs (nanodendrites)) employed for photocatalytic H2 production under one sun conditions using an H2O/MeOH (1:2) mixture, the TiO2-NT(nanotube)/Pd-ND catalyst having the highest surface area showed the highest hydrogen generation rate of 143.1 μmol h−1.173 Mishra et al. synthesized 15 wt% S-modified TiO2/β-SiC and showed that the high photocatalytic hydrogen evolution activity of the catalyst is due to its high specific surface area.45
A 2D TiO2/g-C3N4 photocatalyst with 20 wt% TiO2 having a high specific surface area showed a 2.4-times higher photocatalytic hydrogen evolution rate than g-C3N4 nanosheets having a relatively low specific surface area. The high photocatalytic activity of 2D TiO2/g-C3N4 is due to its high surface area, and hence higher number of available active sites for hydrogen production.181 Another way of utilizing the advantage of the high surface area in improving the photocatalytic H2 production rate of titania is by modifying it in the form of silica-embedded titania and/or titania silica mixed oxide.87 Although the correlation between surface area and photocatalytic activity has been demonstrated in the literature, there are some reports showing that surface area hardly had any influence on the photocatalytic activity and H2 production rate in comparison with other favourable factors, as illustrated in the case of CuO/TiO2 heterojunction and Pt@CuO/TiO2 composites synthesized employing TiO2 nanosheets and nanorod precursors.186 Further, a high surface area is invariably associated with high porosity (with micro and mesopores) and large number of surface defects and can act as recombination centers. Considering these contradicting views, caution should be exercised when deriving a proper conclusion and proceeding further in photocatalysis.
In general, the photocatalytic activity of TiO2 is mainly contributed by the anatase phase due to its high surface area and small particle size; however, as the calcination temperature increases, the content of the less photocatalytically active rutile phase increases.190 Thus, TiO2-based photocatalytic materials show the maximum hydrogen generation at a calcination temperature range of 300–500 °C.190 Besides, as the calcination temperature increases, the probability of charge carrier recombination at the bulk traps also increases.53 The contribution of the anatase phase to enhancing the photocatalytic H2 evolution reaction rate has been exemplified in the literature.191 Among the different N-doped TiO2 samples prepared at different calcination temperatures ranging from 300 °C to 800 °C, NTP-400 (N-doped TiO2 calcined at 400 °C) showed the highest rate of photocatalytic hydrogen production due to the formation of highly crystallized anatase phase and incorporation of N in the crystal lattice.189
Sun et al. reported that the calcination temperature affects the photocatalytic hydrogen evolution activity of TiO2-based photocatalysts because the anatase-to-rutile ratio in TiO2 varies with the calcination temperature.191 Camposeco et al. prepared 0.5Rh/TSG(TiO2) via the sol–gel method and studied its photocatalytic performance as a function of calcination temperature (100 °C, 200 °C, 300 °C, 400 °C and 500 °C) for the water splitting reaction using 50% (v/v) ethanol–water mixture under a UV Pen-Ray Hg lamp (λ = 365 nm). As illustrated in Fig. 20a, 0.5Rh/TSG prepared at 400 °C (anatase) showed the highest H2 production rate of 7246 μmol h−1 g−1.51
Fig. 20 (a) Variation in the rate of hydrogen production on 50 mg of 0.5Rh/TSG(TiO2) as a function of calcination temperature (reproduced with permission from ref. 51. Copyright ©2018 Elsevier Ltd). (b) Variation in the rate of hydrogen production as a function of calcination temperature for the Ru (wt% = 8.0)/TiO2 NB photocatalyst (reproduced with permission from ref. 47. Copyright ©2016, Wiley-VCH). |
A comparison of the photocatalytic H2 evolution activity of Ru (wt% = 8.0)/TiO2NB (nanobelts) non-annealed sample and the same catalyst calcined at different temperatures (200 °C, 400 °C, 600 °C, and 800 °C), as illustrated in Fig. 20b, revealed that Ru 8.0/TiO2 NBs calcined at 400 °C showed the highest photocatalytic hydrogen production rate of 25.34 mmol h−1 g−1.47 The improved activity of Ru 8.0/TiO2 NBs@400 °C is due to its improved crystallinity with the formation of the crystalline-phase Ru/RuO2 and formation of an intimate junction between Ru/RuO2 and TiO2. With a further increase in the calcination temperature, an increased amount of RuO2 was formed due to the oxidation of Ru, which had a detrimental effect on the rate of hydrogen evolution.47
The majority of researchers tried to maintain a greater anatase phase content in TiO2. This is because although the rutile phase is stable, it has poorer photocatalytic activity compared to anatase because of its larger bandgap and smaller surface area. Thus, if this phase transition is not properly managed, the photocatalytic efficiency may significantly decline. Furthermore, high calcination temperatures can lead to particle growth and agglomeration, which reduces the surface area and active sites for photocatalytic reactions. Alternatively, inadequate calcination may result in the presence of leftover organic matter or partially degraded precursors, which can serve as recombination sites for charge carriers produced during the photocatalytic process and hinder the activity.
Wang et al. fabricated a 2D/1D heterojunction of TiO2/CdS composite with (001) facets of TiO2 and CdS nanorods, which showed a photocatalytic hydrogen evolution rate of 128.3 mmol g−1 h−1 with acetic acid-water mixture (20 mL acetic acid in 210 mL water) under visible light (300 W Xe arc lamp; λ > 400 nm), which is much higher than that by the conventionally prepared P25/CdS (35.3 mmol g−1 h−1).179 The outstanding photocatalytic performance of 2D/1D TiO2/CdS is attributed to the morphological features of TiO2 and the formation of an optimized 2D/1D heterojunction, leading to efficient charge separation and enhanced electron transport. The morphology of the cocatalyst also plays a vital role in enhancing the photocatalytic H2 production rate. Thus, between TiO2–Pd NSs (nanosheets) and TiO2–Pd NTs (nanotetrahedrons), the photocatalytic hydrogen production rate was higher on TiO2 integrated with ultrathin nanosheets of Pd due to the rapid generation of hot electrons and their transfer from Pd nanosheets.167 A comparison of the morphological features, optical properties and photocatalytic activities of TiO2–Pd nanosheets and TiO2–Pd nanotetrahedrons is shown in Fig. 21.
Fig. 21 (a) Schematic illustration, TEM (b and c), and HRTEM (d) images of TiO2–PdNSs. (e) Schematic diagram, TEM (f), and HRTEM (g) images of TiO2–Pd NSs with TiO2 nanosheets. (h) Schematic illustration, TEM (i) and (j), and HRTEM (k) images of TiO2–Pd NTs. (l) Schematic illustration and TEM (m) image of TiO2–Pd NTs with TiO2 nanosheets. UV-vis diffuse reflectance spectra (n), photocurrent response (o), PL spectra (p) and photocatalytic H2 production performance (q) of bare TiO2, TiO2–Pd NTs and TiO2–Pd NSs. Schematic illustrating the photocatalytic H2 evolution reaction on the samples of TiO2–Pd NSs and TiO2–Pd NTs (r) under UV and (s) vis-NIR light irradiation, respectively. Reproduced with permission from ref. 167. Copyright ©2016, the Royal Society of Chemistry. |
The photocatalytic H2 production activity of the Pt@CuO/TiO2 composite fabricated using TiO2 nanosheets was higher than that of Pt@CuO/TiO2 prepared using TiO2 nanorods.186 The relatively high activity of the nanosheet-based photocatalyst is due to a variety of reasons, as follows: (1) TiO2 nanosheets mainly have exposed relatively high surface energy (001) facets, whereas TiO2 nanorods are dominated by comparatively lower surface energy (101) facets, (2) CuO has stronger interactions with TiO2 nanosheets than with TiO2 nanorods, (3) CuO can form a more stable p–n heterojunction with TiO2 nanosheets, and (4) TiO2 nanosheets have a low density of defects, which reduce the possibility of charge recombination. Compared to conventional Pd/P25 TiO2, Pd/TiO2 nanosheets showed relatively higher activity for photocatalytic H2 production due to the greater number of exposed high surface energy (001) facets on nanosheets.170 In contrast, for the nanorod morphology, the aspect ratio is crucial for enhancing the photocatalytic water splitting, as reported by Fuet et al. for nanorods of rutile TiO2.192 The different works discussed above show that due to the larger surface area and more active sites, nanostructured TiO2 materials such as nanotubes, nanorods, and nanospheres generally show improved photocatalytic activity compared to bulk materials. Nevertheless, the production of these nanostructures frequently necessitates precise control of the synthesis parameters and more sophisticated and expensive techniques.
Fig. 22 (a) Schematic illustration of the preparation of TiO2–graphene nanocomposites with controllable TiO2 crystal facets exposed. (b) Band structures of TiO2 and TiO2–graphene nanocomposites. (c) H2 evolution rates from methanol solution catalyzed by TiO2 and TiO2–graphene nanocomposites. (d)–(i) TEM and HRTEM images: (d) and (e) TiO2-101-G, (f) and (g) TiO2-001-G, and (h) and (i) TiO2-100-G. Schematic illustration of atomic structures of interfaces between graphene and different TiO2 crystal facets and the photocatalytic process by TiO2–graphene nanocomposites: (j) TiO2-100-G, (k) TiO2-101-G, and (l) TiO2-001-G. Reproduced with permission from ref. 152. Copyright ©2014, Wiley-VCH. |
Exposed facets of certain dopants can also influence the rate of hydrogen generation due to the formation of appropriate energy levels and reduced rate of electron–hole recombination by trapping the photogenerated electrons. For example, The Pt/TiO2 photocatalyst having (111) facets of Pt with a higher Fermi level is highly effective for trapping the electron in the conduction band of TiO2 than that with the (100) facets of Pt.188 In another study, Qi et al. reported that a CdS-sensitized Pt/TiO2 nanosheet photocatalyst with (001) exposed facets showed outstanding water splitting activity due to the special electronic and surface properties of the (001) facets.194 Compared to {101}-TiO2/CdSe QDs, the relatively high photocatalytic production rate shown by {001}-TiO2/CdSe QDs is due to the enhanced charge separation and efficient electron transfer from CdSe QDs to the (001) facets of TiO2.136 Due to the differences in their surface energy, reactive site density, and charge carrier dynamics, different facets of TiO2 exhibit different photocatalytic activities. Hence, facet-controlled synthesis methods are required to maximise the photocatalytic performance. For example, nanocube and octahedron morphologies only expose (100) or (111) facets, respectively, while truncated octahedron exposes both (100) and (111) facets. Thus, by fine tuning these methods, it is possible to improve the performance of photocatalysts and well-defined synthetic methodologies are required.
Metal nanoparticles such as Au and Pt with a size in the range of 1–10 nm supported on TiO2 lead to an increase of energy gap between the Fermi level of metal nanoparticles and the conduction band of TiO2, resulting in improved charge separation and enhanced hydrogen production activity.195 Gold nanoparticles were synthesized via a seed-mediated growth method for the sensitization of Pt/TiO2 catalysts (Fig. 23a) and it was evident from the TEM images (Fig. 23b–e) that narrow-sized Au particles were obtained through this method. Among the Au-sensitized Pt/TiO2 photocatalysts with different particle sizes of Au (10, 20, 30, and 50 nm) for photocatalytic hydrogen production, the Au nanoparticles with a particle size of 20 nm showed the highest performance under visible light (Fig. 23g and h).200 The size of Au nanoparticles is clearly evident from the TEM images, as illustrated in Fig. 23, and the results show that the SPR effect and the electron injection into the CB of TiO2 depends on the particle size of Au (Fig. 23f and h).200 Generally, smaller particles possess a large surface area and a greater number of reaction sites per unit weight of catalyst. Nonetheless, smaller particles can sometimes lead to a decline in photocatalytic activity with time due to particle aggregation. If the size of the particles is smaller than twice the width of the space charge layer, the extent of band bending decreases, which eventually diminishes the potential to separate e−/h+ pairs. Alternatively, visible light-absorbing small quantum dots could be grown inside the pores of a wide bandgap semiconductor and this approach is likely to result in higher activity due to the inherent and inevitable formation of heterojunctions.
Fig. 23 (a) Schematic diagram of the synthesis of Au nanoparticles supported on Pt/TiO2 composite catalysts. TEM images of the synthesized gold nanoparticles with a size of approximately 50 nm (b), 30 nm (c), 20 nm (d), and 10 nm (e). TEM images of Pt/TiO2. (f) Plausible photocatalytic mechanism of Au-sensitized Pt/TiO2 under visible light. (g) Photocurrent response of Au nanoparticles with different sizes supported on Pt/TiO2 nanocomposites under visible-light irradiation (λ > 420 nm). (h) Hydrogen-production activity of different catalysts under visible-light irradiation (λ > 420 nm) in water/methanol mixture. Reproduced with permission from ref. 200. Copyright ©2017, the American Chemical Society. |
A comparison of the H2 evolution activity between Pt-deposited mesoporous TiO2 nonporous Pt/TiO2-P25 revealed that mesoporous Pt/TiO2-450 showed better activity than non-porous Pt/TiO2-P25 because the mesoporous network facilitates efficient charge transfer by forming suitable interfaces and the reactants can easily diffuse through the pores.202 Li et al. reported that the CuO/TiO2 photocatalyst prepared by the chemical adsorption decomposition method (CAD) consisted of a large number of small pores, which facilitated the absorption of a large number of reacting molecules on its surface due to the presence of more active sites and promoted the hydrogen generation activity.184 One of the reasons for the enhanced H2 production rate displayed by the Au-deposited mesoporous S,N-TiO2 (SNT) photocatalyst with a pore diameter in the range of 8 to 13 nm is due to its mesoporous nature, which enhanced the light-harvesting ability of the material.199 Further, using green leaves and highly porous architecture was proved to be beneficial for photocatalysis and light harvesting applications.203,204 For example, Devaraji et al. demonstrated the use of a green leaf as a template to produce an inorganic leaf to produce ZnO with a nano-micro architecture and demonstrated its use for the efficient oxidation of benzene to phenol under UV light. In another work, Chen and coworkers demonstrated the importance of the porous architecture for enhanced photocatalytic H2 production by fabricating TiO2 nanotube arrays (TNTAs) grown on titanium fiber of a titanium web (TNTA-web) and titanium foil (TNTA-foil).205 The TNTA-web generated 40 mmol h−1 m−2 of hydrogen, while the TNTA-foil failed to produce hydrogen under similar conditions. Further, the addition of the same amount of Pd nanoparticles to the TNTA-web and TNTA-foil resulted in an increase in the production rate to 130 and 10 mmol h−1 m2, respectively. The enhanced photocatalytic H2 production activity of the TNTA-web material was attributed to its unique dual porosity. The synthesis procedure and morphological analysis of TNTA-web are depicted in Fig. 24. Photocatalysts having dual porosity, such as hierarchical porous-structured materials containing a combination of macroporosity and mesoporosity, are believed to be more advantageous than single porosity. However, more research in this direction is needed for proper control of the porosity, which can be achieved by careful variation of the synthesis parameters and/or adopting novel synthesis strategies. Establishing a structure–activity relation is an interesting aspect in photocatalytic research, which deserves more attention.
Fig. 24 (a) Synthesis of Pd/TNTA-web (3D TiO2 nanotube array). Scanning electron microscopy (SEM) images of the pristine titanium-web substrate (b) and (c) and TiO2 nanotubes arrays on the web (d), (e) and (f). HR-TEM image of a nanotube (g), SEM image (h), TEM image (i) and HRTEM image (j) of the TNTA-web with 0.1 mg cm−2 Pd (0.26 wt%). Reproduced with permission from ref. 205. Copyright ©2018, the Royal Society of Chemistry. |
The photocatalytic performance of Ag/TiO2 in water splitting reaction for different catalyst amounts revealed that the H2 production rate increased initially with an increase in the catalyst concentration from 15 to 20 mg L−1, and with a further increase in catalyst concentration, the activity declined (see Fig. 25a).112 The catalyst in excess amount may hinder the absorption of radiation, and in certain cases it accelerates the electron–hole recombination.112 The amount of catalyst needed for achieving the optimum activity may vary as the nature of the catalyst, catalyst composition and reaction conditions vary. For example, as illustrated in Fig. 25b, the H2 production rate on Ni/γ-Al2O3/CNT-TiO2 was the maximum with 10 mg catalyst and the activity decreased with a further increase in catalyst amount.102 The excess amount of catalyst adversely affected the light penetration, which decreased the overall catalytic performance.102 In fact, an earlier review66 listed the very high activity associated with a small amount of catalyst (1–3 mg in ∼50–100 mL solution), whereas a larger quantity of the same catalysts (10–200 mg in ∼50–200 mL solution) showed a large drop in hydrogen production activity. The large difference in activity is attributed to the significant light scattering with a larger amount of catalyst quantity, whereas efficient light absorption with a small quantity of catalyst in large volumes of solution. Based on the available literature and above discussion, it can be concluded that careful optimisation of the amount of photocatalyst is necessary to extract the best catalytic performance from a photocatalytic material. Nonetheless, how it can be scaled-up for large volume applications also needs to be addressed. Although this is an engineering issue, this thinking is likely to bring clarity and minimize the problems in the future.
Fig. 25 (a) Photocatalytic hydrogen evolution activity of Ag/TiO2 by varying the amount of catalyst. Reproduced with permission from ref. 112. Copyright ©2019, Elsevier Ltd. (b) Variation in photocatalytic H2 evolution rate with respect to the amount of Ni/γ-Al2O3/CNT-TiO2 catalyst. Reproduced with permission from ref. 102. Copyright ©2017, Elsevier Ltd. |
As illustrated in Fig. 26a, the photocatalytic H2 evolution activity of Cu-decorated TiO2 containing different wt% of Cu (1, 2, 3, 4, and 5 wt%) revealed that the amount of hydrogen produced increases up to 3 wt% of Cu (3.33 mmol g−1 h−1), and with a further increase in Cu concentration, the activity is reduced.104 This study revealed that 3 wt% Cu is the optimum concentration to achieve the maximum activity, and at a higher concentration of Cu (above 3 wt%), the aggregation of Cu particles occurs on the surface of TiO2 and these agglomerated Cu nanoparticles act as recombination centers for photogenerated electrons and holes. This shows that a suitable mass ratio of components is crucial to generate an efficient interface for excellent photocatalytic hydrogen evolution.
Fig. 26 (a) Variation in the H2 evolution activity of TiO2 decorated with different wt% of Cu. Reproduced with permission from ref. 104. Copyright ©2017, the Royal Society of Chemistry. (b) Rate of hydrogen evolution of TiO2 and TiNi samples with different loadings of Ni. Reproduced with permission from ref. 16. Copyright© 2018, the Royal Society of Chemistry. |
In another variation, the photocatalytic hydrogen production activity of Cu2O–TiO2/rGO revealed that TiO2 with the optimum Cu loading of 1 wt% and GO loading of 3 wt% showed the maximum performance because of the enhanced charge transfer and excellent light absorption capacity of the material. The highest photocatalytic H2 production activity was observed with 3 wt% Au-loaded mesoporous S,N-TiO2 when Au-deposited mesoporous S,N-TiO2 with different wt% of Au (0.5, 1, 2, 3, 4 and 5) were compared for the reaction.199 In this case, a high content of gold blocked the active sites and obstructed the penetration of light into the catalyst surface. The high concentration of metal on the surface of TiO2 may also act as electron–hole recombination centres, which reduces the rate of photocatalytic hydrogen production.85 When a combination of metal nanoparticles is employed for the fabrication of TiO2-based composites, the uniform distribution of all the metal nanoparticles on TiO2 is crucial to achieve the maximum photocatalytic performance.85
Reddy et al. reported that a bimetallic Cu/Ag@TNT-2 (CAT-2) photocatalyst with the optimum amount of Ag and Cu prepared using a mixture of 0.1 M Cu and 0.1 M Ag precursor showed the highest rate of photocatalytic hydrogen generation due to the uniform dispersion of Ag and Cu on TiO2.99 A lower number of active sites was present at a low concentration of Cu and Ag on the surface of TiO2, whereas a high concentration of metal nanoparticles led to poor electron transfer.99 Wei et al. reported that a TiO2–Ni hybrid photocatalyst with the optimum Ni amount of 1 mol% showed the highest rate of hydrogen evolution during the photocatalytic water splitting reaction.16 An excess Ni loading limited the light absorption ability and blocked the active sites of the catalyst, leading to a decrease in activity (see Fig. 26b).
The photocatalytic hydrogen evolution activity of Pd-integrated TiO2 nanosheets revealed that the photocatalyst showed the highest activity with an optimum Pd loading of 0.1 mg cm−2.205 The H2 evolution activity decreased when the amount of Pd was higher than the optimum loading due to the increase in Pd particle size and incorporation of Pd atoms in the bulk part rather than on the surface of TiO2. Moreover, the integration of a large amount of Pd in TiO2 limits the expansion of the Pd–TiO2 interface and reduces the number of charge carriers. The incorporation of a small amount of non-metal dopants such as N in anatase TiO2 nanotubes creates defect centers, while maintaining the anatase phase, and strongly promotes the photocatalytic H2 evolution performance of TiO2 nanotube layers, whereas the high-dose N incorporation leads to amorphization of the incorporated region of TiO2.114 In conclusion, the effectiveness of TiO2-based photocatalytic materials is highly sensitive to the amount of dopant used. Both insufficient and excessive dopant levels can significantly impair the photocatalytic efficiency by either failing to adequately enhance charge separation or by introducing recombination centres, respectively.
A comparison of the solar H2 production activity of titania (P25) and Pd/P25 fabricated in thin film form (4.69 cm2) on a glass plate and their powder form counterparts revealed that the Pd/P25 catalyst (1 mg) fabricated in thin film form resulted in an H2 production rate of 104 mmol h−1 g−1, which was about 11–12-times higher than the Pd/P25 catalyst (25 mg) in powder form.165 The digital photograph, FESEM images and photocatalytic performance of the prepared thin film are depicted in Fig. 27a–d. The enhanced activity of Pd/P25 in thin film form is due to its maximum exposed surface area to the reactants and sunlight and other advantageous aspects of thin film, as described above. A similar trend in the photocatalytic H2 production activity of a catalyst in thin film form compared to its powder counterpart was also observed with Ag/TiO2 nanocomposites (Fig. 27e).43
Fig. 27 (a) Digital photographs of thin films of P25 and Pd/P25. FESEM images of Pd/P25 thin films over a glass plate at a scale bar of 2 μm (b) and (c) cross-sectional view of a freshly cleaved thin film (scale bar: 10 μm). Inset in (c) shows the cracks. (d) Photocatalytic H2 evolution activity in the particulate and thin film form of different catalysts (reproduced with permission from ref. 165. Copyright©2019, the Royal Society of Chemistry). (e) Photocatalytic H2 production activity of mesoporous TiO2 and Ag/TiO2 nanocomposites, measured in thin-film form, in aqueous methanol solution under direct sunlight. Reproduced with permission from ref. 43. Copyright ©2021, Wiley-VCH. |
The excellent catalytic performance of Pd/TiO2,165 Ag/TiO2,43 Cu–Ni/TiO2,3 CuxO/TiO2,209 Cu–Ag/TiO2206 and AuPd/C/TiO266 fabricated in thin-film form compared to their powder counterparts has been demonstrated in the literature. Tudu et al. reported that the thin film form of the Cu–Ni/TiO2 (1:1 = Cu:Ni) photocatalyst exhibited an enhanced rate of hydrogen evolution (41.7 mmol h−1 g−1) compared to its powder form (1.75 mmol h−1 g−1) due to the efficient charge generation and its utilization in the film form, which mimicked the natural photosynthesis by green leaves.3 In another report, the film form of the AuPd/rGO/TiO2 photocatalyst displayed an enhanced hydrogen generation (21.5 mmol h−1 g−1) performance compared to its powder form by a factor of 43 times. The higher activity of the film form of the catalyst is attributed to its high light absorption capacity, efficient contact between the particles, and remarkable charge generation and utilization.66
An important aspect of assembling and integrating thin films is to enhance the number of heterojunctions or Schottky junctions. A recent work on integrating BiVO4 in the micro and mesopores of P25-TiO2 showed a solar to fuel efficiency of 31% under one sun condition for artificial photosynthesis to produce methanol and formaldehyde. One of the main reasons attributed is the observation of 174 trillions of BiVO4–TiO2 heterojunctions in 1 mg of material, as reported by Salgaonkar et al.210 To the best of our knowledge, this was the first time that the number of heterojunctions in a material was reported, which was calculated based on the pore-size distribution and surface area by measuring them carefully before and after integrating BiVO4 with TiO2. This indeed opens a new window to fabricate highly integrated materials for efficient charge separation and their diffusion to the redox sites. Similar works may lead to an enhancement in light to chemical conversion for a variety of reactions.
A bifunctional p–n heterojunction of NiO–TiO2 photocatalysts could simultaneously produce H2 and VAPs such as glyceraldehyde and dihydroxyacetone from a glycerol–water solution (10% v/v) under UV-visible light irradiation.211 The best-performing catalyst (7.5% Ni-loaded TiO2) produced 8000 μmol h−1 g−1 hydrogen and could achieve 20% conversion of glycerol. Interestingly, at the beginning of the reaction, the yield of glyceraldehyde was higher than that of DHA; however, after 24 h of reaction, almost the same yields of both the products were obtained.
The reduction reaction occurs on the surface of TiO2, while the glycerol oxidation occurs on NiO. The feasibility test revealed that glyceraldehyde contributes the highest share of annual earnings (89%), followed by dihydroxyacetone (11%) and H2 (0.03%). The photocatalytic H2 production performance, glycerol oxidation and corresponding reaction mechanism of NiO–TiO2 photocatalysts are provided in Fig. 28. In another work, Bajpai et al. achieved an improved H2 production rate of 18 mmol h−1 g−1 with the generation of three VAPs, glycolaldehyde, DHA, and formic acid, from a glycerol–water solution with Au integrated with P25-TiO2 (Au@TiO2).52 The Au@TiO2 catalyst system achieved 4–10% conversion of 0.05 M glycerol to VAPs. The photocatalytic experiments were carried out under three different reaction conditions (aerobic, anaerobic and dry air) and under 2 different light sources (direct sunlight and one sun condition with 100 mW cm−2). The overall products yield was higher under aerobic conditions and sunlight irradiation compared to anaerobic and dry air conditions. However, 7.5 mmol g−1 CO2 was also observed under aerobic conditions due to the over-oxidation of glycerol. The liquid and gaseous products were the same when the photocatalytic experiments were performed under one-sun conditions. However, the CO2 production declined (5.3 mmol g−1) under one sun condition compared to direct sunlight. The dual-function of the catalyst was ascribed to the improved charge transport and efficient separation of photogenerated charge carriers due to the highly dispersed and electronically integrated Au with TiO2. The development of photocatalytic systems that can selectively produce the desired products in high yields from biomass components, without compromising the yield of H2 is a significant challenge in this area. It should also be noted that a similar concept in the electrolysis of aqueous glycerol has been growing rapidly in the last 3–4 years and a few recent works is worth mentioning. Chauhan et al.212 achieved a low-voltage pathway to water electrolysis, while producing H2 at the cathode and value-added products at the anode.
Fig. 28 (a) Schematic illustration of the photocatalytic process. Comparison of H2 production performance of samples with different nickel loadings (b). (c) Glycerol conversion over different photocatalysts. Yield of liquid products obtained with best-performing catalyst (d) and (e) plausible reaction mechanism. Reproduced with permission from ref. 211. Copyright 2023, Elsevier. |
Wu et al. developed a dual-functional photocatalyst system for the treatment of pharmaceutical-contaminated water using Co3O4-modified {001}/{101}-TiO2(TC) nanosheets. The unique aspect of this photocatalytic system is that the p-type semiconductor Co3O4 creates a p–n junction with TiO2 and the {001}/{101} facets of TiO2 form an inherent surface heterojunction, which enhances the charge carrier separation. The electrons on the surface of the TiO2 nanosheet can convert water molecules into molecular hydrogen, while the holes on the Co3O4 surface effectively oxidise pharmaceutical contaminants in the wastewater.213 The interesting idea of the concurrent utilization of electrons and holes was employed by Salgaonkar for different isomers of butanol oxidation to the corresponding butanal/butanone and hydrogen with TiO2–Pd and Pd coated with half-a-monolayer of Pt under visible light.216 This generic procedure for the selective oxidation of alcohol to aldehyde/ketone with H2 is worth pursuing for many different substrates. The mineralization of a serious endocrine disruptor, namely endosulphan, was investigated by Devaraji et al. with a solid solution of ZnO–ZnS under direct sunlight.214 The a-TiO2 (anatase TiO2)/b-AC (biomass activated carbon) nanocomposite synthesized through the ultrasonication technique was very effective for sulphide wastewater treatment together with excellent photocatalytic hydrogen production (400 mL h−1).217 Non-metal-doped TiO2 (NMx-TiO2, where x is the weight percentage of non-metal element) nanocomposites displayed a high hydrogen production rate and COD elimination simultaneously when employed for waste water treatment.215 In this case, using a catalyst loading of 4 g L−1 and light intensity of 5.93 mW cm−2, 7 wt% P-loaded TiO2 (P7/TiO2) could achieve a hydrogen production rate of 8.34 mmol g−1 and 50.6% COD elimination.
The simultaneous oxidation of organic waste and water reduction offer a viable solution to major environmental and energy problems. The goal of ongoing research and development is to remove financial and technological obstacles, enabling the utilization of these technologies in everyday life and their public availability. The development of more sophisticated photocatalysts with increased efficiency has been the focus of recent research. Many organic wastes, such as plastic wastes, agricultural, and industrial pollutants, can be efficiently broken down by photocatalytic reactions. This helps in decreasing the environmental problems caused by waste accumulation. Also, the photocatalytic breakdown of harmful organic compounds into less toxic chemicals may reduce the water and soil pollution.
Fig. 29 (a) Digital photograph and (b) schematic diagram of the flow membrane reactor used for photocatalytic methanol dehydrogenation and reforming. (c) H2 yield obtained with different Cu wt% on PC50. (d) Amount of H2 produced with 1% Cu/PC50 at different temperatures. (e) Schematic illustration of photocatalytic process on 1% Cu/PC50 in a flow membrane reactor. Reproduced with permission from ref. 226. Copyright©2022, the Royal Society of Chemistry. |
The integration of TiO2-based photocatalysts with a second light-absorbing component and/or co-catalyst can significantly influence their H2 production rate. Metal nanoparticles in the form of single atoms and clusters reduce the metal loading and improve the catalytic efficiency. The porosity of a photocatalyst influences the solar hydrogen production rate by providing a large number of surface exposed sites and enhances its light-harvesting ability by transferring the incident photon flux into the internal surface of mesoporous TiO2 through multiple internal reflections and increasing the mass transfer rate. The photocatalysts fabricated in thin-film form show a significantly high performance for solar hydrogen production compared to the same catalysts in particulate form because the catalyst nanoparticles fabricated in thin film are highly connected across the substrate and have a larger exposed surface area to the reactants and sunlight with relatively less mass transfer constraints. The efficient and concurrent utilization of both photogenerated holes and electrons is very important to improve the photocatalytic efficiency of materials and the overall energy efficiency of the process. We also discussed the importance of performing the TiO2-based photocatalytic water splitting reaction in different reactor setups for enhancing the photocatalytic performance of the material and its scalability.
Continuous and dedicated efforts are needed in the area of photocatalytic hydrogen generation to fabricate commercially viable catalysts in scalable form that are efficient to produce sufficient quanta of charge carriers and their utilization. Thus far, less emphasis has been placed on the scalability and long-term stability of the photocatalytic material. In fact, for real world applications, photocatalysts that can deliver the same performance for several months (or even a year) and their scalability are necessary, and hence focus should be placed on developing novel photocatalysts with exceptional stability together with scalability towards commercial applications. Despite the significant advancements in this field, the solar-to-hydrogen (STH) conversion efficiency still needs to be improved for practical applications. At least 10% STH should be achieved, which is expected to attract industrial attention. Overall water splitting leading to the direct conversion of water into hydrogen without using any sacrificial reagent and by reducing greenhouse gas emission is a promising technique.227,228 Renewable H2 production coupled with wastewater treatment with a suitable photocatalyst is another way to compensate for the energy loss at the expense of waste minimization. Renewable H2 production and the concurrent formation of value-added products by employing a waste biomass component, such as glycerol,52,212 are an attractive way to minimize waste with concurrent value addition. The oxidation of biomass components occurs at a significantly lower potential than oxygen formation; in addition, electrons and holes are simultaneously used ensure the catalyst stability for a longer period. Hence, continuous research in the above-mentioned directions is needed to improve the energy efficiency of the process by optimising the photocatalyst and reaction conditions.
This journal is © The Royal Society of Chemistry 2024 |