Next Article in Journal
Study of the Impact of Aerodynamic Model Fidelity on the Flight Characteristics of Unconventional Aircraft
Next Article in Special Issue
Atomic and Electronic Structure of Metal–Salen Complexes [M(Salen)], Their Polymers and Composites Based on Them with Carbon Nanostructures: Review of X-ray Spectroscopy Studies
Previous Article in Journal
Vector Beams with Only Transverse Intensity at Focus
Previous Article in Special Issue
Towards Mirror-Less Graphene-Based Perfect Absorbers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Application of Nano Titanium Dioxide for Hydrogen Production and Storage Enhancement

by
Angelantonio De Benedetto
1,
Agnese De Luca
1,
Paolo Pellegrino
1,2,
Rosaria Rinaldi
1,2,
Valeria De Matteis
1,2,*,† and
Mariafrancesca Cascione
1,2,*,†
1
Department of Mathematics and Physics “E. De Giorgi”, University of Salento, Via Monteroni, 73100 Lecce, Italy
2
Institute for Microelectronics and Microsystems (IMM), National Research Council (CNR), Via Monteroni, 73100 Lecce, Italy
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Appl. Sci. 2023, 13(22), 12521; https://doi.org/10.3390/app132212521
Submission received: 13 October 2023 / Revised: 14 November 2023 / Accepted: 15 November 2023 / Published: 20 November 2023

Abstract

:
The utilization of hydrogen (H2) as a renewable and clean energy carrier, free from the reliance on fossil fuels, represents a significant technological challenge. The use of renewable energy sources for hydrogen production, such as photocatalytic hydrogen generation from water under solar radiation, has garnered significant interest. Indeed, the storage of hydrogen presents another hurdle to the ongoing advancement of hydrogen energy. Concerning solid-state hydrogen storage, magnesium hydride (MgH2) has emerged as a promising option due to its high capacity, excellent reversibility, and cost-effectiveness. Nevertheless, its storage performance needs improvement to make it suitable for practical applications. Titanium dioxide (TiO2) has distinguished itself as the most extensively researched photocatalyst owing to its high photo-activity, good chemical and thermal stability, low toxicity, and affordability. This review highlights the application of TiO2 for hydrogen production under visible and solar light, with a particular focus both on its modification without the use of noble metals and its utilization as a catalyst to enhance the hydrogen storage performance of MgH2.

1. Introduction

The extensive consumption of fossil fuels, particularly over the past century, has significantly contributed to environmental pollution. As environmental and socio-economic consciousness continues to grow, governments and legislative bodies worldwide are expressing concern and examining the factors associated with pollution that affect the energy landscape. Consequently, there is an urgent requirement to replace fossil fuels with renewable, eco-friendly, and alternative energy sources [1,2,3,4,5].
Therefore, to date, the most pressing scientific and technological challenge is the development of a renewable and clean energy that does not rely on fossil fuels. From this perspective, one of the most appealing options for large-scale utilization is hydrogen as a recyclable energy carrier. Consequently, green approaches to hydrogen production, such as the photocatalytic process from water under solar radiation, has garnered considerable interest [6,7]. For photocatalytic hydrogen production, titanium dioxide (TiO2) represents a promising candidate among various inorganic semiconductor photocatalysts [8,9] due to its high efficiency, high stability, non-toxicity, and low cost. However, the practical application remains constrained, since TiO2 responds solely to ultraviolet light, which constitutes approximately 4% of the solar spectrum. Therefore, it is mandatory to design and synthesize TiO2-based photocatalytic systems capable of reacting to the whole visible light range, accounting for ~43% of the solar spectrum, thus improving the efficiency [10,11]. The storage of hydrogen to provide grid energy from intermittent energy sources is another major challenge for the further development of hydrogen energy [12]. On the other hand, magnesium hydride (MgH2) stands out as one of the most promising candidates in solid-state hydrogen storage thanks to its high capacity (7.6 wt%), excellent reversibility, and cost-effectiveness. Nonetheless, its practical application is constrained by its high operating temperature requirements and suboptimal kinetic performance [13]. An effective approach to enhance the hydrogen storage performance of MgH2 involves the introduction of catalysts, which may encompass metals [14,15,16], metal oxides [16,17,18,19], metal sulfides [20,21], and metal halides [19,22].
Focusing on work published in the last decade, this review firstly delves into the utilization of TiO2 for hydrogen production under visible and solar light, as well as its role in enhancing the hydrogen storage performance of MgH2. Successively, in the domain of hydrogen production, various strategies aimed at enhancing TiO2 performance without the use of noble metals are explored.

2. Titanium Dioxide

TiO2 is the most widely investigated photocatalyst due to its remarkable photo-activity, strong chemical and thermal stability, low toxicity, and cost-effectiveness [23]. In nature, four polymorphs of TiO2 are found: anatase (tetragonal), rutile (tetragonal), brookite (orthorhombic), and TiO2(B) (monoclinic). Additionally, two high-pressure forms have been synthesized starting from rutile: TiO2(II) and TiO2(H) [24]. However, rutile and anatase are indeed the two most common crystalline forms of TiO2, and they find numerous applications owing to their unique properties [25]. TiO2 is an n-type semiconductor, and for bulk materials, anatase has a band gap of 3.20 eV, which corresponds to an absorption threshold at a wavelength of 384 nm, while rutile has a band gap of 3.02 eV, which corresponds to an absorption threshold at the 410 nm wavelength [24]. Therefore, near-ultraviolet (UV) radiation is necessary to excite anatase, whereas the photo-response of rutile slightly extends into the visible light spectrum [26]. These varying band gap values can be attributed to differences in the lattice structures of anatase and rutile TiO2, resulting in distinct densities and electronic band structures [24]. The TiO2 photocatalytic activity depends on several factors, including phase structure, specific surface areas, crystallite size, and pore structure. Moreover, the photocatalytic activity of anatase is generally considerably higher than that of rutile [27]. The absorption of photons with energy equal to or greater than the band gap results in the generation of electron–hole pairs through the excitation of electrons from the valence band to the conduction band of the semiconductor. When both electrons and holes migrate to the surface of semiconductor without recombination, the photogenerated electrons can participate in reduction reactions, while the holes engage in oxidation reactions. These processes form the basis for photocatalytic water splitting and the photodegradation of organic pollutants [28].

2.1. Hydrogen Production Exploiting TiO2 under Visible and Solar Light

Water splitting is an uphill reaction that needs the standard Gibbs free energy change of 237 kJ mol−1 [2]. It consists of two half reactions [3]:
2H+ + 2e → H2
2H2O + 4h+ → O2 + 4H+
Therefore, the overall reaction is [2]:
H2O → H2 +1/2 O2
In order for a semiconductor to function as a photocatalyst for water splitting, it is imperative that the bottom of its conduction band exhibits a more negative potential than the reduction potential of H+/H2 at 0 V versus the normal hydrogen electrode (NHE). Simultaneously, the top of its valence band should possess a more positive potential than the oxidation potential of O2/H2O at 1.23 V versus NHE. Therefore, for efficient water splitting, the band gap of the semiconductor should exceed 1.23 eV; conversely, to harness visible light, the band gap should be less than 3.0 eV [29]. Accordingly, a band gap falling within the range of (1.23–3.0) eV is a necessary requisite. In addition, photocatalytic H2 generation is influenced by some factors such as surface area and particle size, band gap energy, corrosion resistance, and the presence of a sacrificial agent [3] such as glycerol [1], triethanolamine (TEOA) [10], ethanol [30], or methanol [11].
For H2 production, the performance of TiO2 is generally limited by the high recombination rate of the photogenerated electrons and holes, an inability to utilize visible light, and fast backward reactions [31]. Loading TiO2 with noble metals such as Pt, Au [32,33,34], or Pd [35] allows the photocatalytic efficiency of TiO2 to be enhanced by reducing the fast recombination of photogenerated charge carriers [11] and allows the TiO2 response to the entire visible light spectral range to be extended [3]. However, the use of noble metals is expensive, and their availability is limited; consequently, their use is unsuitable for large-scale energy production [11]. For these reasons, other strategies have been optimized to improve the photocatalytic efficiency of TiO2, such as employing non-noble metals, metal oxides, cadmium sulfide (CdS), and molybdenum disulfide (MoS2).
Below, different TiO2 modification strategies that have been evaluated under visible and solar light are presented.

2.1.1. Modification of TiO2 with Non-Noble Metals and Metallic Oxides

Numerous approaches have been developed to improve the photocatalytic response of TiO2, in terms of both broadening the reaction spectrum to the entire visible spectral range [11] and mitigating the rapid recombination phenomena of electron–lacuna pairs [36].
The incorporation of non-noble metals (such as Fe, Cu, Co, Ni, and Cr) or metallic oxides onto TiO2 represents a valid alternative to noble metals. In particular, the use of this approach can effectively hinder the recombination of electron–hole pairs and promote the separation of charge carriers [37].
Subha et al. [38] synthesized a photocatalyst system based on TiO2 doped with Cu and Zn (CuZn-TiO2). TEM images and XPS spectra showed the presence of Cu2O, CuO, ZnO, and Cu0 on the TiO2 surface. Metal–semiconductor junctions and heterojunctions between semiconductors were formed due to the growth of Cu2O, CuO, ZnO, and Cu0 on the surface of TiO2. Effective charge separation due to the heterojunctions and high electron mobility due to Cu0 improved the H2 production. In addition, Cu0, CuO, and Cu2O allowed the absorption of TiO2 to shift to visible light. In particular, the CuZn-TiO2 with an optimum loading of Cu and Zn of 0.5 wt% achieved a H2 production of about 14,521 μmol h−1 g−1cat under solar light, using glycerol as sacrificial reagent. A plausible mechanism for H2 production is reported in the work of Subha et al. [38] and is schematically represented in Figure 1.
A cobalt (Co)-doped TiO2 photocatalyst was synthesized by Sadanandam et al. [36] through the impregnation method. The presence of cobalt species on the TiO2 surface of the synthesized photocatalyst was observed. The Co/TiO2 photocatalyst, with 1 wt% of Co doped on TiO2 P25, achieved a maximum hydrogen production of 11,021 μmol h−1 g−1 from aqueous glycerol solutions under solar light irradiation, showing that Co2+ ions can extend the photo-response of TiO2 into the visible region [36].
Díaz et al. [11] prepared a series of photocatalyst systems consisting of TiO2 P25 combined with Fe, Co, Ni, Cu, and Zn (denotated as Fe/TiO2, Co/TiO2, Ni/TiO2, Cu/TiO2, and Zn/TiO2, respectively) using the impregnation method. The TiO2 was loaded with 2 wt% of the selected transition metal and evaluated under both visible and UV light irradiation. During visible light exposure, with methanol serving as a sacrificial reagent, the photocatalysts Ni/TiO2, Co/TiO2, Zn/TiO2, and Fe/TiO2 exhibited lower H2 production rates than the pure TiO2. Conversely, Cu/TiO2 demonstrated a substantial enhancement in photocatalytic activity, yielding H2 production rates equal to 220 μmol h−1 g−1 [11].
Kotesh Kumar et al. [39] developed a photocatalyst system comprising bimetallic Cu-Ni alloy NPs decorated on TiO2 P25. The Cu-Ni/TiO2 photocatalyst, with a Cu loading of 2 mol% and Ni loading of 5 mol%, achieved a remarkable H2 evolution rate of 35.4 mmol g−1 h−1 under solar light conditions, employing methanol as a sacrificial reagent. Additionally, the monometallic photocatalysts Cu-doped TiO2 (Cu/TiO2) and Ni-doped TiO2 (Ni/TiO2) were also synthesized and assessed. It was observed that the monometallic photocatalysts exhibited lower H2 yields compared to the Cu-Ni/TiO2 systems [39]. Reddy et al. [40] fabricated mesoporous nano TiO2 and introduced Fe/TiO2 catalysts by synthesizing and introducing Fe dopants onto the TiO2 surface through the impregnation process. These catalysts were tested under solar light irradiation and the H2 production rate of 270 μmol/h was attained with 0.5 wt% of Fe3+ loading. A comparison between Fe doping on the TiO2 surface and Fe doping into the TiO2 lattice revealed that metal ions should be doped in the proximity to the TiO2 surface to achieve efficient photocatalytic activity through enhanced charge transfer [40]. Police et al. [1] prepared a Cu2O/TiO2/Bi2O3 ternary photocatalyst: the presence of Cu2O and Bi2O3 allows the visible-light absorption to be extended and helps minimize the electron–hole recombination of TiO2, respectively. When exposed to solar light irradiation, the ternary photocatalyst, with 2 wt% Cu and 2 wt% Bi, demonstrated an impressive H2 production rate of 3678 μmol/h in glycerol aqueous solution. Furthermore, the dependence of photocatalytic H2 production on the glycerol and catalyst content was investigated. It was observed that under an optimal glycerol concentration and catalyst amount, a H2 production rate of 6727 μmol/h was achieved [1].
A photocatalyst consisting of mesoporous ZrO2-TiO2 NPss anchored on reduced graphene oxide (rGO) was synthesized by Subha et al. [41]. The ZrO2-TiO2/rGO photocatalyst, with a loading of 1.0 wt% of ZrO2 and rGO on TiO2, exhibited remarkable H2 production (~7773 μmol h−1 g−1) in a glycerol aqueous solution under solar light irradiation. Furthermore, a ZrO2-TiO2 photocatalyst, with 1.0 wt% of ZrO2 loading on TiO2, was also examined, and it achieved an H2 production rate of 5781 μmol h−1 g−1. The ZrO2-TiO2 heterojunction anchored on rGO sheets demonstrated the suppression of electron–hole pair recombination and an enhancement in the efficiency of interfacial charge transfer [41].
Praveen Kumar et al. [42] synthesized a photocatalyst composed of CuO deposited onto TiO2 nanotubes. The photocatalyst, with a copper loading of 1.5 wt%, exhibited the highest photocatalytic hydrogen evolution rate under solar light using glycerol as a sacrificial agent. It achieved an impressive H2 production rate of 99,823 μmol h−1 g−1 [42].
Subha et al. [43] synthesized a CuxO/TiO2 nanocomposite loaded with layered Ni(OH)2. Under solar light irradiation and in the presence of glycerol (5 vol%), the Ni(OH)2-CuxO-TiO2 photocatalyst, with an optimal loading of 1.0 wt% Ni(OH)2 and 0.5 wt% CuxO, showed a H2 production of ~15,789 μmol h−1g−1. Furthermore, Cu-doped TiO2 photocatalysts (CuxO/TiO2) were tested. It was shown that the photocatalyst with 0.5 wt% Cu loading on TiO2 allowed an advancement in H2 production (~7679 μmol h−1 g−1) [43].

2.1.2. Coupling TiO2 with CdS

The CdS semiconductor has been extensively studied as a sensitizer for photocatalytic water splitting due to its narrow band gap and strong absorption in the visible-light spectral region [44]. Combining TiO2 with a CdS semiconductor can further enhance the photocatalytic performance of TiO2 [45]. Luo et al. [46] synthesized a CdS/TiO2(B) type II heterojunction via photo deposition, with CdS dots anchored on TiO2(B) nanosheets. Under visible light, a H2 evolution rate of 1577 μmol g−1 h−1 was achieved. Additionally, a CdS/TiO2(B) type I heterojunction was synthesized via the hydrothermal method; however, it resulted in a H2 evolution rate of 48 μmol g−1 h−1 under visible light, significantly lower than that obtained with the type II heterojunction. The proposed mechanism for H2 production by CdS/TiO2(B) reported by Luo et al. [46] is schematically represented in Figure 2.
Zhu et al. [47] utilized TiO2 nanotube arrays, obtained by anodizing titanium foils, in conjunction with CdS quantum dots to create CdS quantum dots decorated TiO2 nanotube arrays. Under visible light and in the presence of Na2SO3 and Na2S as sacrificial reagents, a H2 production rate of 1.89 μmol h−1 cm−2 was achieved [47].
Gao et al. [44] prepared TiO2 microspheres with highly exposed (001) facets loaded with CdS. They observed that CdS/TiO2 microsphere composites with a CdS:TiO2 molar ratio of 1 indicated the highest activity, resulting in a H2 production rate of 76.55 μmol/h when lactic acid was used as the scavenger for photo-generated holes under visible light [44].
A porous TiO2 co-doped with CdS and WO3 was synthesized by Qian et al. [45]. The ternary porous CdS/WO3/TiO2 photocatalyst, with a molar ratio of CdS:WO3:TiO2 equal to 8:8:100, achieved a H2 evolution rate of 2106 μmol g−1 h−1 under visible light using Na2S and Na2SO3.
Furthermore, the binary CdS/TiO2 and WO3/TiO2 porous photocatalysts were tested, resulting a H2 evolution rate of 1056 μmol g−1 h−1 and 981 μmol g−1 h−1 for CdS/TiO2 and WO3/TiO2, respectively. The possible mechanism for H2 production using the CdS/WO3/TiO2 photocatalyst proposed by Qian et al. [45] is schematically represented in Figure 3.

2.1.3. Coupling TiO2 with MoS2

The efficiency of photocatalytic hydrogen production can also be improved by exploiting MoS2 as a co-catalyst for TiO2 [48]. Several strategies have been developed to improve the performance of MoS2/TiO2 photocatalysts under visible light. Yuan et al. [49] introduced a composite photocatalyst system based on TiO2 nanosheets and MoS2, in which the visible light absorption of TiO2/MoS2 was expanded by incorporating CuInS2 quantum dots as a light-harvesting inorganic dye. The ternary CuInS2/TiO2/MoS2 photocatalyst with 0.6 mmol g−1 of CuInS2 and 0.5 wt% of MoS2 exhibited a H2 evolution rate of 1034 μmol h−1 g−1 under visible light in the presence of Na2S and Na2SO3 as sacrificial reagents. A possible mechanism for H2 production is reported in the work of Yuan et al. [49] and is schematically represented in Figure 4. A Zn(II)-5,10,15,20-tetrakis(4-carboxyphenyl)-porphyrin (ZnTCPP)-sensitized MoS2/TiO2 photocatalyst was synthesized by Yuan et al. [10], exploiting the ZnTCPP to improve the absorption in visible light and the MoS2 to enhance the photocatalytic activity. The ZnTCPP-MoS2/TiO2 photocatalyst was prepared by synthesizing MoS2/TiO2 composites, and then adsorbing the ZnTCPP dye on the TiO2 surface. TiO2 P25 was used. A H2 production rate of 10.2 μmol h−1, under visible light, was achieved using the photocatalyst loaded with 1.00 wt% of MoS2. TEOA was used as the sacrificial electron donor in the photocatalytic hydrogen evolution experiments. Shen et al. [50] synthesized MoS2 nanosheet–porous TiO2 nanowire hybrid nanostructures using the hydrothermal method. A H2 evolution rate of 16.7 mmol h−1 g−1 was achieved under visible light in the presence of Eosin Y dye in TEOA aqueous solution. Flower-like MoS2@TiO2 catalysts were synthesized by Ma et al. [51] using a hydrothermal method, using a metal–organic framework (MOF) as a precursor. The MoS2@TiO2 photocatalysts, with a loading of 14.6 wt% of MoS2, allowed a H2 production rate of 10,046 μmol h−1 g−1 under visible light to be achieved in the presence of fluorescein as photosensitizer in an aqueous acetone solution and using TEOA as a sacrificial donor.

2.1.4. Coupling TiO2 with MoS2 and CdS

Du et al. [52] developed a photocatalyst based on a porous TiO2 monolith with MoS2 nanosheets and CdS nanoparticles grown on the porous TiO2. The MoS2-CdS-TiO2 photocatalyst was designed to exploit both CdS and MoS2 to extend the absorption to visible light and to improve the hydrogen production efficiency. The MoS2-CdS-TiO2 photocatalyst with 3% of MoS2 and 10% of CdS (molar ratio to TiO2) showed a H2 generation rate of 4146 μmol h−1 g−1 using Na2S and Na2SO3 as sacrificial reagents under visible light. In addition, CdS-TiO2 and MoS2-TiO2 photocatalysts were also tested. A H2 generation rate of 875 μmol h−1 g−1 and 1360 μmol h−1 g−1 was obtained with CdS-TiO2 and MoS2-TiO2, respectively. Qin et al. [53] prepared MoS2/CdS-TiO2 nanofibers via an electrospinning-mediated photo-deposition method. The MoS2/CdS-TiO2 photocatalyst, with a loading of 1 wt% of MoS2, showed a H2 evolution rate of 280.0 μmol h−1 in a lactic acid aqueous solution under visible light.
The main works that describes TiO2-based photocatalysts modified with non-noble metals, metallic oxides, CdS, and MoS2 for H2 production under visible and solar light were summarized in Table 1.

2.2. Exploiting the TiO2 Properties for Hydrogen Storage

For solid-state hydrogen storage, several materials such as MgH2 [13,54,55,56], NaAlH4 [57], LiAlH4 [58], and LiBH4 [59,60] have been considered due to their higher gravimetric and volumetric hydrogen densities [61]. However, their practical applications are limited due to their sluggish kinetic and high dehydrogenation temperature [54,57,58,62]. Concerning MgH2 [13,54,55,56,63,64,65,66,67,68], NaAlH4 [57,69,70], LiAlH4 [71,72] and LiBH4 [60,62], the use of TiO2 as a catalyst has been investigated in order to improve the dehydrogenation/rehydrogenation kinetics of these materials, showing an improvement in performance by adding TiO2. Moreover, the use of TiO2 has also been investigated for MgH2-NaAlH4 composite [61] and 4MgH2-LiAlH4 composite [73]. Both composites showed significant improvements in the dehydrogenation/rehydrogenation kinetics with the introduction of TiO2.

Application of TiO2 to Enhance the Hydrogen Storage Performance of MgH2

Several strategies have been adopted to improve the hydrogen storage performance of MgH2 by exploiting TiO2 as catalyst. A carbon-supported nanocrystalline TiO2 was synthesized by Zhang et al. [63] as a catalyst precursor. The carbon-supported nanocrystalline TiO2 (TiO2@C) exhibited good catalytic activity in the hydrogen storage reaction of MgH2. It was observed that the TiO2@C-containing MgH2 composite with a TiO2@C content of 10 wt% released 6.5 wt% hydrogen within 7 min at 300 °C. Moreover, the dehydrogenated sample took up approximately 6.6 wt% hydrogen within 10 min at 140 °C under 50 bar of H2. Chen et al. [54] explored the catalytic effect of synthesized Co/TiO2 nanocomposite on the hydrogen desorption/absorption properties of MgH2. It was observed that using Co/TiO2 highly improved the hydrogen desorption/absorption kinetics of MgH2 in comparison to using Co or TiO2 separately. The MgH2-Co/TiO2 composite showed an absorption of 6.07 wt% H2 within 10 min at 165 °C under 60 bar of H2 pressure and a dehydrogenation peak temperature of 235.2 °C. A sandwich-like structure composed of Ti3C2 and carbon-supported anatase TiO2 (Ti3C2/TiO2(A)-C) was developed by Gao et al. [55] in order to exploit the laminar MXene and the corresponding metal oxide to enhance the hydrogen absorption and desorption kinetics of MgH2. The Ti3C2/TiO2(A)-C structure was prepared using a heat treatment. The catalyst-doped MgH2 composites with 5 wt% of Ti3C2/TiO2(A)-C released approximately 5 wt% hydrogen within 1700 s at 250 °C under 0.05 MPa of hydrogen pressure, with a larger rate constant of 0.258 wt% min−1. An uptake of approximately 4 wt% hydrogen within 800 s at 125 °C, under 3 MPa hydrogen pressure, was also observed. Unlike the preparation of the TiO2/MXene heterostructure via the heat treatment of MXene, Gao et al. [13] also developed self-assembled TiO2/Ti3C2Tx heterostructures through a one-step ultrasonic method at room temperature. The TiO2 nanoparticles were self-assembled onto a few layers of Ti3C2Tx. The TiO2/Ti3C2Tx-MgH2 composite with 5 wt% of TiO2/Ti3C2Tx achieved the fastest dehydrogenation and hydrogenation kinetics, showing a hydrogen desorption of 5.98 wt% within 600 s at 300 °C and a hydrogen absorption of 5.90 wt% within 1200 s at 175 °C. A catalyst composed of TiO2 nanoparticles deposited on the pore wall of a three-dimensionally ordered macroporous (3DOM) structure was synthesized by Shao et al. [56]. It was observed that the 3DOM TiO2 catalyst improved the hydrogen storage properties of MgH2. The MgH2-3DOM TiO2 composite showed a hydrogen desorption of 5.75 wt% within 1000 s at 300 °C. A hydrogen absorption of 4.17 wt% within 1800 s even at 100 °C was observed. Jardim et al. [64] investigated the effects of TiO2 nanorods on the absorption/desorption properties of MgH2. The TiO2 nanorods were produced through titanate nanotube heat treatment, whereas titanate nanotubes were synthesized via the alkaline hydrothermal route using commercial TiO2 anatase as a starting material. At 350 °C, the MgH2-TiO2 nanorod composite showed a hydrogen absorption of 5.5 wt% after 10 min under 10 bar of hydrogen pressure. Furthermore, it was observed that after 5 min at 350 °C under 0.1 bar of hydrogen pressure, the composite desorbed around 100% of the absorbed hydrogen. A Ni/TiO2 cocatalyst with a Ni@TiO2 core–shell structure was synthesized by Zhang et al. [65] via a modified hydrothermal synthesis method. The MgH2-Ni/TiO2 composite with 5 wt% of Ni/TiO2 allowed a hydrogen absorption of 4.50 wt% to be obtained at 50 °C under an initial hydrogen pressure of 3.0 MPa and a hydrogen desorption of 5.24 wt% in 1800 s at 250 °C under 0.005 MPa of hydrogen pressure [65]. Zhang et al. [66] synthesized and investigated three-dimensional flower-like TiO2 (fl-TiO2) and three-dimensional flower-like carbon-wrapped TiO2 (fl-TiO2@C) as the catalysts for MgH2. Both catalysts enhanced the hydrogen sorption kinetics of MgH2. Superior desorption kinetics of the MgH2-fl-TiO2@C composite compared to the MgH2-fl-TiO2 composite were observed. The MgH2-fl-TiO2@C composite allowed a hydrogen desorption of 6.0 wt% to be obtained in 7 min at 250 °C under a hydrogen pressure of 1 kPa. Concerning the absorption, the MgH2-fl-TiO2@C composite showed a hydrogen absorption of 6.3 wt% at 150 °C within 40 min under 5 MPa of hydrogen pressure. Liu et al. [67] synthesized graphene-supported TiO2 nanoparticles (TiO2@rGO) using the solvothermal method. It was observed that the MgH2-TiO2@rGO composite desorbed 6.0 wt% hydrogen within 6 min at 300 °C and absorbed 5.9 wt% hydrogen within 2 min at 200 °C. The enhancement of the catalytic activity can be due to fine, uniform TiO2 nanoparticles that were obtained by exploiting the combined effect of ethylene glycol and graphene during the solvothermal process. In addition, rGO can act as an electronic conductive channel to enhance the catalytic effect. Zhang et al. [68] synthesized TiO2 nanosheets with exposed {001} facets in order to exploit the nanometer-size and highly active {001} facets of anatase TiO2 to enhance the reaction kinetics of MgH2. The MgH2-TiO2 nanosheet composite allowed a hydrogen desorption of 6.0 wt% to be obtained at 260 °C within 192 s and, in addition, a hydrogen desorption of 1.2 wt% within 300 min at 180 °C was also observed. Furthermore, the composite absorbed 6.1 wt% hydrogen within 10 s at 150 °C. Also, a hydrogen absorption of 3.3 wt% hydrogen at 50 °C within 10 s was obtained.
The general strategy to improve H2 storage performance of MgH2 using TiO2-based catalysts are reported in Table 2.

3. Conclusions

This review has predominantly focused on the utilization of titanium dioxide (TiO2) in the dual capacity of hydrogen (H2) production and H2 storage. Concerning the aspect of H2 production, extensive efforts have been dedicated to enhancing the photocatalytic efficiency of TiO2 under both visible and solar light, all while abstaining from the utilization of precious metals. In the realm of solar-driven H2 production, TiO2 has frequently undergone modifications involving non-precious metals and metal oxides. Glycerol has consistently featured as the sacrificial agent in the majority of the documented research endeavors.
In the context of visible-light-driven H2 production, the strategies have commonly entailed the amalgamation of TiO2 with MoS2, complemented by the incorporation of a photosensitizer, and the fusion of TiO2 with CdS. Turning our attention to the subject of H2 storage, this review has unequivocally demonstrated the catalytic prowess of TiO2 in augmenting the H2 storage performance of magnesium hydride (MgH2). A myriad of strategies has been employed, encompassing the synthesis of TiO2 nanosheets, flower-like TiO2 structures, TiO2 nanorods, TiO2 nanoparticles, TiO2/Ti3C2Tx heterostructures, graphene-supported TiO2 nanoparticles, carbon-enveloped nanocrystalline TiO2, Co/TiO2 nanocomposites, and Ni@TiO2 core–shell structures.
The integration of MgH2 with TiO2-based catalysts has yielded notable enhancements in H2 absorption and desorption kinetics, leading to lower operating temperatures for H2 desorption and absorption processes, especially when compared to pure MgH2.

Author Contributions

Conceptualization, V.D.M. and M.C.; writing—original draft preparation, A.D.B.; writing—review and editing A.D.L., P.P., R.R., V.D.M. and M.C.; supervision, V.D.M. and M.C. All authors have read and agreed to the published version of the manuscript.

Funding

M.C. kindly acknowledges PRP@CERIC-CUP J97G22000400006 for sponsoring her salary and work. V.D.M. kindly acknowledges Programma Operativo Nazionale (PON) Ricerca e Innovazione 2014–2020-Azione IV.6 “Contratti su tematiche green”-DM 1062/2021 for sponsoring her salary and work. A.D.L. acknowledges PON 2014-2020, Asse I “Capitale Umano”, “Dottorati Innovativi con caratterizzazione industriale”, CUP F85F21005780001; and A.D.B. kindly acknowledges Progetto Tisma, CUP F83C21000150001 PNR–MUR. M4C2–Dalla ricerca all’impresa—1.1: Fondo per il Programma Nazionale della Ricerca (PNR) e Progetti di Ricerca di Rilevante Interesse Nazionale (PRIN).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Police, A.K.R.; Vattikuti, S.V.P.; Mandari, K.K.; Chennaiahgari, M.; Sharma, P.; Valluri, D.K.; Byon, C. Bismuth oxide cocatalyst and copper oxide sensitizer in Cu2O/TiO2/Bi2O3 ternary photocatalyst for efficient hydrogen production under solar light irradiation. Ceram. Int. 2018, 44, 11783–11791. [Google Scholar] [CrossRef]
  2. Huang, J.-F.; Lei, Y.; Luo, T.; Liu, J.-M. Photocatalytic H2 Production from Water by Metal-free Dye-sensitized TiO2 Semiconductors: The Role and Development Process of Organic Sensitizers. ChemSusChem 2020, 13, 5863–5895. [Google Scholar] [CrossRef] [PubMed]
  3. Singh, R.; Dutta, S. A review on H2 production through photocatalytic reactions using TiO2/TiO2-assisted catalysts. Fuel 2018, 220, 607–620. [Google Scholar] [CrossRef]
  4. Ge, M.; Cai, J.; Iocozzia, J.; Cao, C.; Huang, J.; Zhang, X.; Shen, J.; Wang, S.; Zhang, S.; Zhang, K.-Q.; et al. A review of TiO2 nanostructured catalysts for sustainable H2 generation. Int. J. Hydrogen Energy 2017, 42, 8418–8449. [Google Scholar] [CrossRef]
  5. Cai, J.; Shen, J.; Zhang, X.; Ng, Y.H.; Huang, J.; Guo, W.; Lin, C.; Lai, Y. Light-Driven Sustainable Hydrogen Production Utilizing TiO2 Nanostructures: A Review. Small Methods 2019, 3, 1800184. [Google Scholar] [CrossRef]
  6. Abe, R. Recent progress on photocatalytic and photoelectrochemical water splitting under visible light irradiation. J. Photochem. Photobiol. C Photochem. Rev. 2010, 11, 179–209. [Google Scholar] [CrossRef]
  7. Fajrina, N.; Tahir, M. A critical review in strategies to improve photocatalytic water splitting towards hydrogen production. Int. J. Hydrogen Energy 2019, 44, 540–577. [Google Scholar] [CrossRef]
  8. De Matteis, V.; Cannavale, A.; Blasi, L.; Quarta, A.; Gigli, G. Chromogenic device for cystic fibrosis precocious diagnosis: A “Point of Care” tool for sweat test. Sens. Actuators B Chem. 2016, 225, 474–480. [Google Scholar] [CrossRef]
  9. Wondimu, T.H.; Bayeh, A.W.; Kabtamu, D.M.; Xu, Q.; Leung, P.; Shah, A.A. Recent progress on tungsten oxide-based materials for the hydrogen and oxygen evolution reactions. Int. J. Hydrogen Energy 2022, 47, 20378–20397. [Google Scholar] [CrossRef]
  10. Yuan, Y.; Lu, H.; Ji, Z.; Zhong, J.; Ding, M.; Chen, D.; Li, Y.; Tu, W.; Cao, D.; Yu, Z.; et al. Enhanced visible-light-induced hydrogen evolution from water in a noble-metal-free system catalyzed by ZnTCPP-MoS2/TiO2 assembly. Chem. Eng. J. 2015, 275, 8–16. [Google Scholar] [CrossRef]
  11. Díaz, L.; Rodríguez, V.D.; González-Rodríguez, M.; Rodríguez-Castellón, E.; Algarra, M.; Núñez, P.; Moretti, E. M/TiO2 (M = Fe, Co, Ni, Cu, Zn) catalysts for photocatalytic hydrogen production under UV and visible light irradiation. Inorg. Chem. Front. 2021, 8, 3491–3500. [Google Scholar] [CrossRef]
  12. Yang, Y.; Gao, P.; Wang, Y.; Sha, L.; Ren, X.; Zhang, J.; Yang, P.; Wu, T.; Chen, Y.; Li, X. A simple and efficient hydrogen production-storage hybrid system (Co/TiO2) for synchronized hydrogen photogeneration with uptake. J. Mater. Chem. A 2017, 5, 9198–9203. [Google Scholar] [CrossRef]
  13. Gao, H.; Shi, R.; Shao, Y.; Liu, Y.; Zhu, Y.; Zhang, J.; Hu, X.; Li, L.; Ba, Z. One-step self-assembly of TiO2/MXene heterostructures for improving the hydrogen storage performance of magnesium hydride. J. Alloys Compd. 2022, 895, 162635. [Google Scholar] [CrossRef]
  14. Liang, G.; Huot, J.; Boily, S.; Van Neste, A.; Schulz, R. Catalytic effect of transition metals on hydrogen sorption in nanocrystalline ball milled MgH2–Tm (Tm=Ti, V, Mn, Fe and Ni) systems. J. Alloys Compd. 1999, 292, 247–252. [Google Scholar] [CrossRef]
  15. Chen, G.; Zhang, Y.; Chen, J.; Guo, X.; Zhu, Y.; Li, L. Enhancing hydrogen storage performances of MgH2 by Ni nano-particles over mesoporous carbon CMK-3. Nanotechnology 2018, 29, 265705. [Google Scholar] [CrossRef]
  16. Sadhasivam, T.; Sterlin Leo Hudson, M.; Pandey, S.K.; Bhatnagar, A.; Singh, M.K.; Gurunathan, K.; Srivastava, O.N. Effects of nano size mischmetal and its oxide on improving the hydrogen sorption behaviour of MgH2. Int. J. Hydrogen Energy 2013, 38, 7353–7362. [Google Scholar] [CrossRef]
  17. Zhang, X.; Shen, Z.; Jian, N.; Hu, J.; Du, F.; Yao, J.; Gao, M.; Liu, Y.; Pan, H. A novel complex oxide TiVO3.5 as a highly active catalytic precursor for improving the hydrogen storage properties of MgH2. Int. J. Hydrogen Energy 2018, 43, 23327–23335. [Google Scholar] [CrossRef]
  18. Patah, A.; Takasaki, A.; Szmyd, J.S. Influence of multiple oxide (Cr2O3/Nb2O5) addition on the sorption kinetics of MgH2. Int. J. Hydrogen Energy 2009, 34, 3032–3037. [Google Scholar] [CrossRef]
  19. Bhat, V.V.; Rougier, A.; Aymard, L.; Darok, X.; Nazri, G.; Tarascon, J.M. Catalytic activity of oxides and halides on hydrogen storage of MgH2. J. Power Sources 2006, 159, 107–110. [Google Scholar] [CrossRef]
  20. Wang, P.; Tian, Z.; Wang, Z.; Xia, C.; Yang, T.; Ou, X. Improved hydrogen storage properties of MgH2 using transition metal sulfides as catalyst. Int. J. Hydrogen Energy 2021, 46, 27107–27118. [Google Scholar] [CrossRef]
  21. Zhang, W.; Xu, G.; Cheng, Y.; Chen, L.; Huo, Q.; Liu, S. Improved hydrogen storage properties of MgH2 by the addition of FeS2 micro-spheres. Dalton Trans. 2018, 47, 5217–5225. [Google Scholar] [CrossRef]
  22. Plerdsranoy, P.; Thiangviriya, S.; Dansirima, P.; Thongtan, P.; Kaewsuwan, D.; Chanlek, N.; Utke, R. Synergistic effects of transition metal halides and activated carbon nanofibers on kinetics and reversibility of MgH2. J. Phys. Chem. Solids 2019, 124, 81–88. [Google Scholar] [CrossRef]
  23. Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.M.; Hamilton, J.W.J.; Byrne, J.A.; O’Shea, K.; et al. A review on the visible light active titanium dioxide photocatalysts for environmental applications. Appl. Catal. B Environ. 2012, 125, 331–349. [Google Scholar] [CrossRef]
  24. Carp, O.; Huisman, C.L.; Reller, A. Photoinduced reactivity of titanium dioxide. Prog. Solid State Chem. 2004, 32, 33–177. [Google Scholar] [CrossRef]
  25. Haggerty, J.E.S.; Schelhas, L.T.; Kitchaev, D.A.; Mangum, J.S.; Garten, L.M.; Sun, W.; Stone, K.H.; Perkins, J.D.; Toney, M.F.; Ceder, G.; et al. High-fraction brookite films from amorphous precursors. Sci. Rep. 2017, 7, 15232. [Google Scholar] [CrossRef]
  26. Li, G.; Chen, L.; Graham, M.E.; Gray, K.A. A comparison of mixed phase titania photocatalysts prepared by physical and chemical methods: The importance of the solid–solid interface. J. Mol. Catal. Chem. 2007, 275, 30–35. [Google Scholar] [CrossRef]
  27. Zhang, J.; Zhou, P.; Liu, J.; Yu, J. New understanding of the difference of photocatalytic activity among anatase, rutile and brookite TiO2. Phys. Chem. Chem. Phys. 2014, 16, 20382–20386. [Google Scholar] [CrossRef]
  28. Kang, X.; Liu, S.; Dai, Z.; He, Y.; Song, X.; Tan, Z. Titanium Dioxide: From Engineering to Applications. Catalysts 2019, 9, 191. [Google Scholar] [CrossRef]
  29. Chen, X.; Shen, S.; Guo, L.; Mao, S.S. Semiconductor-based Photocatalytic Hydrogen Generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef]
  30. Serafin, J.; Ouzzine, M.; Sreńscek-Nazzal, J.; Llorca, J. Photocatalytic hydrogen production from alcohol aqueous solutions over TiO2-activated carbon composites decorated with Au and Pt. J. Photochem. Photobiol. Chem. 2022, 425, 113726. [Google Scholar] [CrossRef]
  31. Dang, H.; Cheng, Z.; Yang, W.; Chen, W.; Huang, W.; Li, B.; Shi, Z.; Qiu, Y.; Dong, X.; Fan, H. Room-temperature synthesis of CuxS (x = 1 or 2) co-modified TiO2 nanocomposite and its highly efficient photocatalytic H2 production activity. J. Alloys Compd. 2017, 709, 422–430. [Google Scholar] [CrossRef]
  32. Jovic, V.; Al-Azri, Z.H.N.; Chen, W.-T.; Sun-Waterhouse, D.; Idriss, H.; Waterhouse, G.I.N. Photocatalytic H2 Production from Ethanol–Water Mixtures Over Pt/TiO2 and Au/TiO2 Photocatalysts: A Comparative Study. Top. Catal. 2013, 56, 1139–1151. [Google Scholar] [CrossRef]
  33. Jovic, V.; Chen, W.-T.; Sun-Waterhouse, D.; Blackford, M.G.; Idriss, H.; Waterhouse, G.I.N. Effect of gold loading and TiO2 support composition on the activity of Au/TiO2 photocatalysts for H2 production from ethanol–water mixtures. J. Catal. 2013, 305, 307–317. [Google Scholar] [CrossRef]
  34. Serafin, J.; Soler, L.; Vega, D.; Rodríguez, Á.; Llorca, J. Macroporous silicon coated with M/TiO2 (M = Au, Pt) as a highly efficient photoreactor for hydrogen production. Chem. Eng. J. 2020, 393, 124701. [Google Scholar] [CrossRef]
  35. Rusinque, B.; Escobedo, S.; de Lasa, H. Photoreduction of a Pd-Doped Mesoporous TiO2 Photocatalyst for Hydrogen Production under Visible Light. Catalysts 2020, 10, 74. [Google Scholar] [CrossRef]
  36. Sadanandam, G.; Lalitha, K.; Kumari, V.D.; Shankar, M.V.; Subrahmanyam, M. Cobalt doped TiO2: A stable and efficient photocatalyst for continuous hydrogen production from glycerol: Water mixtures under solar light irradiation. Int. J. Hydrogen Energy 2013, 38, 9655–9664. [Google Scholar] [CrossRef]
  37. Yan, Z.; Yu, X.; Zhang, Y.; Jia, H.; Sun, Z.; Du, P. Enhanced visible light-driven hydrogen production from water by a noble-metal-free system containing organic dye-sensitized titanium dioxide loaded with nickel hydroxide as the cocatalyst. Appl. Catal. B Environ. 2014, 160–161, 173–178. [Google Scholar] [CrossRef]
  38. Subha, N.; Mahalakshmi, M.; Myilsamy, M.; Neppolian, B.; Murugesan, V. Direct Z-scheme heterojunction nanocomposite for the enhanced solar H2 production. Appl. Catal. Gen. 2018, 553, 43–51. [Google Scholar] [CrossRef]
  39. Kotesh Kumar, M.; Naresh, G.; Vijay Kumar, V.; Sai Vasista, B.; Sasikumar, B.; Venugopal, A. Improved H2 yields over Cu-Ni-TiO2 under solar light irradiation: Behaviour of alloy nano particles on photocatalytic H2O splitting. Appl. Catal. B Environ. 2021, 299, 120654. [Google Scholar] [CrossRef]
  40. Reddy, J.K.; Lalitha, K.; Reddy, P.V.L.; Sadanandam, G.; Subrahmanyam, M.; Kumari, V.D. Fe/TiO2: A Visible Light Active Photocatalyst for the Continuous Production of Hydrogen from Water Splitting Under Solar Irradiation. Catal. Lett. 2014, 144, 340–346. [Google Scholar] [CrossRef]
  41. Subha, N.; Mahalakshmi, M.; Myilsamy, M.; Lakshmana Reddy, N.; Shankar, M.V.; Neppolian, B.; Murugesan, V. Effective excitons separation on graphene supported ZrO2-TiO2 heterojunction for enhanced H2 production under solar light. Int. J. Hydrogen Energy 2018, 43, 3905–3919. [Google Scholar] [CrossRef]
  42. Praveen Kumar, D.; Shankar, M.V.; Mamatha Kumari, M.; Sadanandam, G.; Srinivas, B.; Durgakumari, V. Nano-size effects on CuO/TiO2 catalysts for highly efficient H2 production under solar light irradiation. Chem. Commun. 2013, 49, 9443. [Google Scholar] [CrossRef] [PubMed]
  43. Subha, N.; Mahalakshmi, M.; Monika, S.; Neppolian, B. Ni(OH)2-CuxO-TiO2 nanocomposite for the enhanced H2 production under solar light: The mechanistic pathway. Int. J. Hydrogen Energy 2020, 45, 7552–7561. [Google Scholar] [CrossRef]
  44. Gao, B.; Yuan, X.; Lu, P.; Lin, B.; Chen, Y. Enhanced visible-light-driven photocatalytic H2-production activity of CdS-loaded TiO2 microspheres with exposed (001) facets. J. Phys. Chem. Solids 2015, 87, 171–176. [Google Scholar] [CrossRef]
  45. Qian, Y.; Yang, M.; Zhang, F.; Du, J.; Li, K.; Lin, X.; Zhu, X.; Lu, Y.; Wang, W.; Kang, D.J. A stable and highly efficient visible-light-driven hydrogen evolution porous CdS/WO3/TiO2 photocatalysts. Mater. Charact. 2018, 142, 43–49. [Google Scholar] [CrossRef]
  46. Luo, X.; Ke, Y.; Yu, L.; Wang, Y.; Homewood, K.P.; Chen, X.; Gao, Y. Tandem CdS/TiO2(B) nanosheet photocatalysts for enhanced H2 evolution. Appl. Surf. Sci. 2020, 515, 145970. [Google Scholar] [CrossRef]
  47. Zhu, Y.; Wang, Y.; Chen, Z.; Qin, L.; Yang, L.; Zhu, L.; Tang, P.; Gao, T.; Huang, Y.; Sha, Z.; et al. Visible light induced photocatalysis on CdS quantum dots decorated TiO2 nanotube arrays. Appl. Catal. Gen. 2015, 498, 159–166. [Google Scholar] [CrossRef]
  48. Shah, S.A.; Khan, I.; Yuan, A. MoS2 as a Co-Catalyst for Photocatalytic Hydrogen Production: A Mini Review. Molecules 2022, 27, 3289. [Google Scholar] [CrossRef]
  49. Yuan, Y.-J.; Fang, G.; Chen, D.; Huang, Y.; Yang, L.-X.; Cao, D.-P.; Wang, J.; Yu, Z.-T.; Zou, Z.-G. High light harvesting efficiency CuInS2 quantum dots/TiO2/MoS2 photocatalysts for enhanced visible light photocatalytic H2 production. Dalton Trans. 2018, 47, 5652–5659. [Google Scholar] [CrossRef]
  50. Shen, M.; Yan, Z.; Yang, L.; Du, P.; Zhang, J.; Xiang, B. MoS2 nanosheet/TiO2 nanowire hybrid nanostructures for enhanced visible-light photocatalytic activities. Chem. Commun 2014, 50, 15447–15449. [Google Scholar] [CrossRef]
  51. Ma, B.; Guan, P.-Y.; Li, Q.-Y.; Zhang, M.; Zang, S.-Q. MOF-Derived Flower-like MoS2 @TiO2 Nanohybrids with Enhanced Activity for Hydrogen Evolution. ACS Appl. Mater. Interfaces 2016, 8, 26794–26800. [Google Scholar] [CrossRef] [PubMed]
  52. Du, J.; Wang, H.; Yang, M.; Zhang, F.; Wu, H.; Cheng, X.; Yuan, S.; Zhang, B.; Li, K.; Wang, Y.; et al. Highly efficient hydrogen evolution catalysis based on MoS2/CdS/TiO2 porous composites. Int. J. Hydrogen Energy 2018, 43, 9307–9315. [Google Scholar] [CrossRef]
  53. Qin, N.; Xiong, J.; Liang, R.; Liu, Y.; Zhang, S.; Li, Y.; Li, Z.; Wu, L. Highly efficient photocatalytic H2 evolution over MoS2/CdS-TiO2 nanofibers prepared by an electrospinning mediated photodeposition method. Appl. Catal. B Environ. 2017, 202, 374–380. [Google Scholar] [CrossRef]
  54. Chen, M.; Xiao, X.; Zhang, M.; Zheng, J.; Liu, M.; Wang, X.; Jiang, L.; Chen, L. Highly dispersed metal nanoparticles on TiO2 acted as nano redox reactor and its synergistic catalysis on the hydrogen storage properties of magnesium hydride. Int. J. Hydrogen Energy 2019, 44, 15100–15109. [Google Scholar] [CrossRef]
  55. Gao, H.; Liu, Y.; Zhu, Y.; Zhang, J.; Li, L. Catalytic effect of sandwich-like Ti3C2/TiO2(A)-C on hydrogen storage performance of MgH2. Nanotechnology 2020, 31, 115404. [Google Scholar] [CrossRef] [PubMed]
  56. Shao, Y.; Gao, H.; Tang, Q.; Liu, Y.; Liu, J.; Zhu, Y.; Zhang, J.; Li, L.; Hu, X.; Ba, Z. Ultra-fine TiO2 nanoparticles supported on three-dimensionally ordered macroporous structure for improving the hydrogen storage performance of MgH2. Appl. Surf. Sci. 2022, 585, 152561. [Google Scholar] [CrossRef]
  57. Huang, Y.; Shao, H.; Zhang, Q.; Zang, L.; Guo, H.; Liu, Y.; Jiao, L.; Yuan, H.; Wang, Y. Layer-by-layer uniformly confined Graphene-NaAlH4 composites and hydrogen storage performance. Int. J. Hydrogen Energy 2020, 45, 28116–28122. [Google Scholar] [CrossRef]
  58. Ma, Z.; Liu, J.; Zhao, Y.; Zhang, J.; Zhu, Y.; Zhang, Y.; Liu, Y.; Li, L. Enhanced dehydrogenation properties of LiAlH4–Mg2NiH 4 nanocomposites via doping Ti-based catalysts. Mater. Res. Express 2019, 6, 075067. [Google Scholar] [CrossRef]
  59. Saldan, I.; Campesi, R.; Zavorotynska, O.; Spoto, G.; Baricco, M.; Arendarska, A.; Taube, K.; Dornheim, M. Enhanced hydrogen uptake/release in 2LiH–MgB2 composite with titanium additives. Int. J. Hydrogen Energy 2012, 37, 1604–1612. [Google Scholar] [CrossRef]
  60. Xian, K.; Nie, B.; Li, Z.; Gao, M.; Li, Z.; Shang, C.; Liu, Y.; Guo, Z.; Pan, H. TiO2 decorated porous carbonaceous network structures offer confinement, catalysis and thermal conductivity for effective hydrogen storage of LiBH4. Chem. Eng. J. 2021, 407, 127156. [Google Scholar] [CrossRef]
  61. Rafi-ud-din; Xuanhui, Q.; Zahid, G.H.; Asghar, Z.; Shahzad, M.; Iqbal, M.; Ahmad, E. Improved hydrogen storage performances of MgH2–NaAlH4 system catalyzed by TiO2 nanoparticles. J. Alloys Compd. 2014, 604, 317–324. [Google Scholar] [CrossRef]
  62. Guo, L.; Jiao, L.; Li, L.; Wang, Q.; Liu, G.; Du, H.; Wu, Q.; Du, J.; Yang, J.; Yan, C.; et al. Enhanced desorption properties of LiBH4 incorporated into mesoporous TiO2. Int. J. Hydrogen Energy 2013, 38, 162–168. [Google Scholar] [CrossRef]
  63. Zhang, X.; Leng, Z.; Gao, M.; Hu, J.; Du, F.; Yao, J.; Pan, H.; Liu, Y. Enhanced hydrogen storage properties of MgH2 catalyzed with carbon-supported nanocrystalline TiO2. J. Power Sources 2018, 398, 183–192. [Google Scholar] [CrossRef]
  64. Jardim, P.M.; da Conceição, M.O.T.; Brum, M.C.; dos Santos, D.S. Hydrogen sorption kinetics of ball-milled MgH2–TiO2 based 1D nanomaterials with different morphologies. Int. J. Hydrogen Energy 2015, 40, 17110–17117. [Google Scholar] [CrossRef]
  65. Zhang, J.; Shi, R.; Zhu, Y.; Liu, Y.; Zhang, Y.; Li, S.; Li, L. Remarkable Synergistic Catalysis of Ni-Doped Ultrafine TiO2 on Hydrogen Sorption Kinetics of MgH2. ACS Appl. Mater. Interfaces 2018, 10, 24975–24980. [Google Scholar] [CrossRef]
  66. Zhang, M.; Xiao, X.; Luo, B.; Liu, M.; Chen, M.; Chen, L. Superior de/hydrogenation performances of MgH2 catalyzed by 3D flower-like TiO2@C nanostructures. J. Energy Chem. 2020, 46, 191–198. [Google Scholar] [CrossRef]
  67. Liu, G.; Wang, L.; Hu, Y.; Sun, C.; Leng, H.; Li, Q.; Wu, C. Enhanced catalytic effect of TiO2@rGO synthesized by one-pot ethylene glycol-assisted solvothermal method for MgH2. J. Alloys Compd. 2021, 881, 160644. [Google Scholar] [CrossRef]
  68. Zhang, M.; Xiao, X.; Wang, X.; Chen, M.; Lu, Y.; Liu, M.; Chen, L. Excellent catalysis of TiO2 nanosheets with high-surface-energy {001} facets on the hydrogen storage properties of MgH2. Nanoscale 2019, 11, 7465–7473. [Google Scholar] [CrossRef]
  69. Xiong, R.; Sang, G.; Yan, X.; Zhang, G.; Ye, X.; Zhu, X. Direct synthesis of nanocrystalline titanium dioxide/carbon composite and its catalytic effect on NaAlH4 for hydrogen storage. Int. J. Hydrogen Energy 2011, 36, 15652–15657. [Google Scholar] [CrossRef]
  70. Xiong, R.; Sang, G.; Yan, X.; Zhang, G.; Ye, X.; Jiang, C.; Luo, L. Improvement of the hydrogen storage kinetics of NaAlH4 with nanocrystalline titanium dioxide loaded carbon spheres (Ti-CSs) by melt infiltration. Int. J. Hydrogen Energy 2012, 37, 10222–10228. [Google Scholar] [CrossRef]
  71. Zhao, Y.; Han, M.; Wang, H.; Chen, C.; Chen, J. LiAlH4 supported on TiO2/hierarchically porous carbon nanocomposites with enhanced hydrogen storage properties. Inorg. Chem. Front. 2016, 3, 1536–1542. [Google Scholar] [CrossRef]
  72. Amama, P.B.; Grant, J.T.; Shamberger, P.J.; Voevodin, A.A.; Fisher, T.S. Improved Dehydrogenation Properties of Ti-Doped LiAlH4: Role of Ti Precursors. J. Phys. Chem. C 2012, 116, 21886–21894. [Google Scholar] [CrossRef]
  73. Mustafa, N.S.; Yahya, M.S.; Itam Sulaiman, N.N.; Abdul Halim Yap, M.F.A.; Ismail, M. Enhanced the hydrogen storage properties and reaction mechanisms of 4MgH2 + LiAlH4 composite system by addition with TiO2. Int. J. Energy Res. 2021, 45, 21365–21374. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of a possible mechanism (Z-scheme mechanism with double charge transfer mechanism) for H2 production by CuZn-TiO2 photocatalyst under solar light.
Figure 1. Schematic representation of a possible mechanism (Z-scheme mechanism with double charge transfer mechanism) for H2 production by CuZn-TiO2 photocatalyst under solar light.
Applsci 13 12521 g001
Figure 2. Schematic representation of a possible mechanism for H2 production by CdS/TiO2(B) photocatalyst under visible light.
Figure 2. Schematic representation of a possible mechanism for H2 production by CdS/TiO2(B) photocatalyst under visible light.
Applsci 13 12521 g002
Figure 3. Schematic representation of a possible mechanism for H2 production using the CdS/WO3/TiO2 photocatalyst under visible light.
Figure 3. Schematic representation of a possible mechanism for H2 production using the CdS/WO3/TiO2 photocatalyst under visible light.
Applsci 13 12521 g003
Figure 4. Schematic representation of a plausible mechanism for H2 production by CuInS2/TiO2/MoS2 photocatalyst under visible light.
Figure 4. Schematic representation of a plausible mechanism for H2 production by CuInS2/TiO2/MoS2 photocatalyst under visible light.
Applsci 13 12521 g004
Table 1. TiO2-based photocatalysts modified with non-noble metals, metallic oxides, CdS, and MoS2 for H2 production under visible and solar light.
Table 1. TiO2-based photocatalysts modified with non-noble metals, metallic oxides, CdS, and MoS2 for H2 production under visible and solar light.
PhotocatalystCrystalline Phases of TiO2LoadingIrradiationSacrificial ReagentsH2 ProductionRef.
Cu2O/TiO2/Bi2O3anatase2 wt% of Cusolar lightglycerol6727 μmol h−1[1]
2 wt% of Bi
ZnTCPP-MoS2/TiO2anatase/rutile1.00 wt% of MoS2visible lightTEOA10.2 μmol h−1[10]
Cu/TiO2anatase/rutile2 wt% of Cuvisible lightmethanol220 μmol h−1 g−1[11]
Co/TiO2anatase/rutile1 wt% of Cosolar lightglycerol11,021 μmol h−1 g−1[36]
CuZn-TiO2anatase0.5 wt% of Cusolar lightglycerol14,521 μmol h−1 g−1[38]
0.5 wt% of Zn
Cu-Ni/TiO2anatase/rutile2 mol% of Cusolar lightmethanol35.4 mmol h−1 g−1[39]
5 mol% of Ni
Fe/TiO2anatase0.5 wt% of Fe3+solar light-270 μmol h−1[40]
ZrO2-TiO2/rGOanatase1.0 wt% of ZrO2solar lightglycerol7773 μmol h−1 g−1[41]
1.0 wt% of rGO
CuO/TiO2anatase/rutile1.5 wt% of Cusolar lightglycerol99,823 μmol h−1 g−1[42]
Ni(OH)2-CuxO-TiO2anatase1.0 wt% of Ni(OH)2solar lightglycerol7679 μmol h−1 g−1[43]
0.5 wt% of CuxO
CdS/TiO2anataseCdS:TiO2 molar
ratio of 1
visible lightlactic acid76.55 μmol h−1[44]
CdS/WO3/TiO2anatasemolar ratio 8:8:100visible lightNa2S and Na2SO32106 μmol h−1 g−1[45]
CdS/TiO2-1056 μmol h−1 g−1
WO3/TiO2-981 μmol h−1 g−1
CdS/TiO2(B)TiO2(B)5% of CdSvisible lightNa2S and Na2SO31577 μmol h−1 g−1[46]
CdS/TiO2anatase-visible lightNa2S and Na2SO31.89 μmol h−1 cm−2[47]
CuInS2/TiO2/MoS2anatase0.5 wt% of MoS2visible lightNa2S and Na2SO31034 μmol h−1 g−1[49]
MoS2/TiO2
nanowire and
Eosin Y
anatase-visible lightTEOA16.7 mmol h−1 g−1[50]
Flower-like MoS2@TiO2 and fluorescein anatase14.6 wt% of MoS2visible lightTEOA10,046 μmol h−1 g−1[51]
MoS2-CdS-TiO2anatase3% of MoS2 and 10% of CdS molar ratio to TiO2visible lightNa2S and Na2SO34146 μmol h−1 g−1[52]
CdS-TiO2-875 μmol h−1 g−1
MoS2-TiO2-1360 μmol h−1 g−1
MoS2/CdS-TiO2 nanofibersanatase1 wt% of MoS2visible lightlactic acid280.0 μmol h−1[53]
Table 2. Summary of strategies adopted to improve H2 storage performance of MgH2 using TiO2-based catalysts.
Table 2. Summary of strategies adopted to improve H2 storage performance of MgH2 using TiO2-based catalysts.
CompositeHydrogen
Desorption
Hydrogen Desorption Properties
(Time, Temp., Pressure)
Hydrogen
Absorption
Hydrogen Absorption Properties
(Time, Temp., Pressure)
Ref.
MgH2-TiO2/Ti3C2Tx5.98 wt%600 s, 300 °C, 0.05 MPa5.90 wt%1200 s, 175 °C, 3 MPa[13]
MgH2-Co/TiO26.25 wt%10 min, 265 °C, 0.02 bar6.07 wt%10 min, 165 °C, 60 bar[54]
6.2 wt%15 min, 250 °C, 0.02 bar5.56 wt%10 min, 130 °C, 60 bar
5.4 wt%100 min, 220 °C, 0.02 bar4.24 wt%10 min, 100 °C, 60 bar
4.6 wt%100 min, 210 °C, 0.02 bar~ 1 wt%10 min, 50 °C, 60 bar
MgH2-Ti3C2/TiO2(A)-C5 wt%1700 s, 250 °C, 0.05 MPa4 wt%800 s, 125 °C, 3 MPa[55]
MgH2-3DOM TiO25.75 wt%1000 s, 300 °C, 0.005 MPa4.17 wt%1800 s, 100 °C, 3 MPa[56]
5.74 wt%3000 s, 275 °C, 0.005 MPa5.40 wt%1000 s, 175 °C, 3 MPa
MgH2-TiO2@C6.5 wt%7 min, 300 °C, -6.6 wt%10 min, 140 °C, 50 bar[63]
MgH2-TiO2 nanorods~5.5 wt%5 min, 350 °C, 0.1 bar5.5 wt%10 min, 350 °C, 10 bar[64]
MgH2-Ni/TiO25.24 wt%1800 s, 250 °C, 0.005 MPa4.50 wt%120 min, 50 °C, 3.0 MPa[65]
6.08 wt%600 s, 300 °C, 0.005 MPa
MgH2-fl-TiO2@C6.0 wt%7 min, 250 °C, 1 kPa6.3 wt%40 min, 150 °C, 5 MPa[66]
4.9 wt%60 min, 225 °C, 1 kPa5.0 wt%40 min, 100 °C, 5 MPa
3.0 wt%100 min, 200 °C, 1 kPa3.9 wt%40 min, 50 °C, 5 MPa
MgH2-fl-TiO26.0 wt%8 min, 250 °C, 1 kPa6.0 wt%40 min, 150 °C, 5 MPa[66]
3.9 wt%60 min, 225 °C, 1 kPa
0.9 wt%100 min, 200 °C, 1 kPa
MgH2-TiO2@rGO6.0 wt%6 min, 300 °C, 0.0004 MPa5.9 wt%2 min, 200 °C, 3 MPa[67]
MgH2-TiO2 nanosheets6.0 wt%192 s, 260 °C, 1 kPa6.1 wt%10 s, 150 °C, 5 MPa[68]
1.2 wt%300 min, 180 °C, 1 kPa3.3 wt%10 s, 50 °C, 5 MPa
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

De Benedetto, A.; De Luca, A.; Pellegrino, P.; Rinaldi, R.; De Matteis, V.; Cascione, M. The Application of Nano Titanium Dioxide for Hydrogen Production and Storage Enhancement. Appl. Sci. 2023, 13, 12521. https://doi.org/10.3390/app132212521

AMA Style

De Benedetto A, De Luca A, Pellegrino P, Rinaldi R, De Matteis V, Cascione M. The Application of Nano Titanium Dioxide for Hydrogen Production and Storage Enhancement. Applied Sciences. 2023; 13(22):12521. https://doi.org/10.3390/app132212521

Chicago/Turabian Style

De Benedetto, Angelantonio, Agnese De Luca, Paolo Pellegrino, Rosaria Rinaldi, Valeria De Matteis, and Mariafrancesca Cascione. 2023. "The Application of Nano Titanium Dioxide for Hydrogen Production and Storage Enhancement" Applied Sciences 13, no. 22: 12521. https://doi.org/10.3390/app132212521

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