Open access peer-reviewed chapter

Black TiO2 Material for Highly Efficient Green Hydrogen Production Enabled by Advanced Surface Engineering

Written By

Xiaodan Wang, Beibei Wang and Hao Shen

Submitted: 22 February 2024 Reviewed: 10 March 2024 Published: 21 May 2024

DOI: 10.5772/intechopen.1005042

From the Edited Volume

Hydrogen Technologies - Advances, Insights, and Applications

Zak Abdallah and Nada Aldoumani

Chapter metrics overview

31 Chapter Downloads

View Full Metrics

Abstract

Black TiO2 (H-TiO2), as a promising photoanode material, can be used for direct green hydrogen production without emissions to pollute the environment, but the reported surface engineering approaches for the preparation of black TiO2 suffer from high temperatures, long processing time, or chemical residues, limiting its practical application in green hydrogen production. Here, we developed two advanced surface engineering technologies, overcoming the above limitations, to prepare a black TiO2 photoanode that achieved the maximum photocurrent density reported to date. Moreover, we theoretically and experimentally revealed the formation mechanism of black TiO2 and its enhanced photoelectrochemical (PEC) performance. These surface engineering technologies are not only suitable for the preparation of efficient photoanode materials for PEC hydrogen production but also play a beneficial and promoting role in the research and development of new materials for hydrogen fuel cells and hydrogen storage.

Keywords

  • photoelectrochemical green hydrogen production
  • black TiO2
  • surface engineering technologies
  • formation mechanism
  • hydrogen fuel cells
  • hydrogen storage

1. Introduction

Photoelectrochemical (PEC) technology for solar fuel generation has garnered international recognition as a promising solution for addressing both energy and environmental challenges [1, 2]. This technology has the capability to mimic photosynthesis by converting solar energy into hydrogen energy and also has the potential to convert CO2 into hydrocarbon fuel, making it an ideal solution for a sunshine economy.

The technology of PEC hydrogen production was initiated in 1972 by Fujishima and Honda, which discovered that single crystal TiO2 can generate sufficient photovoltage when exposed to light to facilitate the splitting of water into hydrogen and oxygen [3]. This breakthrough demonstrated the potential of utilizing solar energy to directly produce green hydrogen without CO2 emissions, which is regarded as the most promising method for achieving zero-carbon hydrogen in the future. Subsequently, semiconductor photocatalysts have triggered significant interest within the academic community.

TiO2 has emerged as a prominent semiconductor photocatalyst among various materials due to its stable chemical properties, robust oxidation-reduction capabilities, and cost-effectiveness. However, as a wide bandgap semiconductor, TiO2 can only absorb ultraviolet light within the solar spectrum. Enhancing the responsiveness of TiO2 to visible light has become a key focus of research in this field. Recent studies have explored surface engineering technologies via doping of various elements to improve visible light absorption of TiO2, yet the resulting photocurrent remains below 1 mA cm−2 [4]. The field of surface engineering for TiO2 is currently grappling with the obstacle of achieving high solar-to-hydrogen conversion efficiency.

The global interest in the synthesis of black TiO2 through high-pressure hydrogenation of TiO2 has been significant. In 2011, Chen et al. presented an innovative approach to creating black TiO2 in Science [5]. Figure 1 visually represents the unique core-shell structure of black TiO2 and an accompanying energy diagram illustrating a decreased band gap. The researchers verified that black TiO2 displays a wider band for light absorption, allowing it to capture visible, ultraviolet, and infrared light, exhibiting outstanding photocatalytic characteristics. Since its inception, black TiO2 has attracted considerable interest in the scientific community due to its unique combination of high sunlight absorption and exceptionally high photocatalytic abilities. However, its applications are hampered by impractical hydrogenation conditions (20 bar, 200°C, and 5 days). Developing new hydrogenation technologies at mild conditions is highly demanded.

Figure 1.

(a) Core-shell structure and narrowing band gap models; (b) white pristine TiO2 to black TiO2; (c) and (d) high-resolution transmission electron microscope (HR-TEM) images of pristine TiO2 and black TiO2, respectively.

Advertisement

2. Recent advances in hydrogen technologies promoted by emerging hydrogenation methods

Hydrogenation technology is known as one of the most important material surface engineering technologies today, and here, we highlighted the latest research progress on how hydrogenation technologies can solve the core scientific problems of hydrogen-related technologies like PEC hydrogen production, hydrogen fuel cells and hydrogen storage.

2.1 PEC green hydrogen production

Wang et al. utilized high-temperature H2 hydrogenation to prepare H-TiO2 nanowires with significantly improved PEC performance for solar water splitting [6]. The H-TiO2 nanorods treated at 350°C achieve the best photocurrent density of 2.4 mA cm−2 at 1.23 VRHE compared to untreated nanorods, indicating efficient charge separation and transportation (Figure 2a). The incident photon to current efficiency (IPCE) analyses confirm that the enhancement of photocurrent density of H-TiO2 is attributed to the enhanced photoactivity in the ultraviolet region (100%) and a significant increase in the donor density (3×), achieved through the creation of a high density of oxygen vacancies that act as electron donors. However, the harsh conditions of the high-temperature H2 approach (400°C, 1 h) limit its practical application in PEC water splitting.

Figure 2.

Photocurrent density versus potential (J-V) measurements of H-TiO2 via (a) high-temperature H2 and (b) Pd-catalyzed H hydrogenations.

Xu et al. explored a Pd-catalyzed H hydrogenation strategy for reducing TiO2, leading to the generation of Ti3+ species in TiO2 and improved photocatalytic activity (Figure 2b) [7]. The hydrogenation process creates intrinsic defects in the materials, narrowing the band gap and enhancing solar energy utilization. The unique crystalline core/disordered shell structure of H-TiO2 formed during the process contributes to improved photocatalytic activity and long-term stability. However, the high cost of noble metal and chemical residues in this hydrogenation method hinder its practical application in PEC hydrogen production.

2.2 Hydrogen fuel cells

The hydrogen fuel cell is divided into two chambers, which pass into H2 and O2, respectively; in the left chamber, under the catalytic Pt, H2 dissociates to protons and electrons (2H2 → 4H+ + 4e), protons reach the counter electrode of the right chamber through the solid electrolyte, and at the same time, the electrons pass through the peripheral circuit to the counter electrode, so as to cooperate with the reduction reaction of protons to O2 to generate H2O (O2 + 4H+ + 4e → 2H2O), which completes the transfer of chemical energy to electrical energy (Figure 3a). It is evident that conversion efficiency depends strongly on the proton conductivity of solid electrolytes.

Figure 3.

(a) Scheme of hydrogen fuel cells. (b) Temperature-dependent proton conductivity: HSrCoO2.5 vs. state-of-the art. (c) Scheme of noble metal catalyzed H hydrogenation and resulting organized oxygen vacancy channels in HSrCoO2.5.

Solid oxide ionic conductors are utilized to serve as solid electrolytes in hydrogen fuel cells. Meanwhile, traditional low ionic conductors based on metal oxides typically necessitate temperatures exceeding 500°C to facilitate ionic transport (Figure 3b). Very recently, Lu et al. reported a new solid oxide proton conductor, HSrCoO2.5, prepared by noble metal catalyzed H hydrogenation, exhibiting remarkably high proton conductivity within the temperature range of 40–140°C [8]. The proton conductivity in the specified temperature range ranged from 0.028 to 0.33 S cm−1 (Figure 3b), which is ascribed to the elevated proton concentration and the organized oxygen vacancy channels facilitated by the noble metal catalyzed H hydrogenation (Figure 3c). However, the prohibitive expense of noble metals and the presence of chemical residues in the hydrogenation process pose significant obstacles to its feasibility for use in practical hydrogen fuel cells.

2.3 Hydrogen storage

Solid-state hydrogen storage materials, facilitating the storage and controlled release of hydrogen, have been a focal point of extensive research for numerous decades, establishing ambitious goals for their advancement in recent years. Developing such materials that can store an appreciable amount of hydrogen reversibly without having to use high pressure and/or low temperature is a significant challenge.

Izumi et al. prepared crystalline samples of the material La1−xSrxH3−x−2yOy (0.1 ≤ x ≤ 0.6, y < 0.171) through ball-milling, followed by annealing under high-pressure H2 hydrogenation [9]. They conducted an examination of the samples at ambient temperature, determining their ability to facilitate the conduction of hydride ions at a notable rate. Subsequently, they assessed the efficacy of the material in an all-solid-state cell composed of Ti| La1−xSrxH3−x−2yOy|LaH3−δ, manipulating the quantities of strontium and oxygen within the composition. Upon identifying an optimal strontium content of at least 0.2, they observed the complete conversion of titanium to titanium hydride (TiH2) at a rate of 100%, indicating minimal wastage of hydride ions. Based on the results, they focus on enhancing performance and developing electrode materials with the capability to reversibly absorb and release hydrogen. This advancement is crucial for enabling the recharging of batteries and facilitating the storage and controlled release of hydrogen, thereby meeting the demands of hydrogen-based energy applications. While the above findings have made important advances in hydrogen storage, the high-temperature and high-pressure hydrogen hydrogenation method (400°C, 0.5 MPa, 12 h) they employed will prevent the development of its practical applications.

Advertisement

3. Advanced hydrogenation technologies for PEC green hydrogen production

The above-hydrogenated surface engineering technologies have shown great potential in PEC green hydrogen production, fuel cells, and hydrogen storage applications; however, the efficacy of their applications is hindered by impractical hydrogenation conditions and ambiguous hydrogenation mechanisms. Here, two advanced methods, low-temperature H [10] and room-temperature H+ hydrogenations [11], are developed to address the challenging conditions associated with traditional high-temperature H2 reductions and the constraints posed by chemical residues resulting from chemical reductions.

3.1 Low-temperature H hydrogenation

3.1.1 Experimental set-up

The concept of low-temperature hot wire H hydrogenation draws inspiration from previous research. Langmuir’s study in 1912 investigated the efficiency of a tungsten filament in facilitating the dissociation of H2 molecules into H atoms [12]. Experimental results demonstrated that when a tungsten filament is heated to temperatures exceeding 1000°C in a hydrogen-rich environment, H2 molecules can be catalytically dissociated upon contact with the hot tungsten filament. In contrast to the H+ hydrogenation approach, this method exclusively involves only hydrogen atoms rather than hydrogen ions.

Figure 4a depicts the experimental configuration utilized for the low-temperature H hydrogenation. Within a vacuum chamber, a parallel array of 10× tungsten wires, each measuring 600 mm in length, 0.53 mm in diameter, and 50 mm in space, was positioned, resulting in a total activated area of 600 × 450 mm2. The samples were subjected to treatment at varying wire temperatures (1600, 1700, and 1800°C), with a separation of 75 mm between the wires and the samples.

Figure 4.

(a) Experimental set-up of low-temperature H hydrogenation and (b) plot of the flux of atomic hydrogen vs. wire temperature.

The regulation of the flux rate of atomic hydrogen is contingent upon the wire temperature and pressure, with the pressure typically preset at 1 Pa. Figure 4b shows the plot of the flux of atomic hydrogen vs. wire temperature. It is evident that the flux of atomic hydrogen increases exponentially with increased wire temperature. This means that the low-temperature H hydrogenation [10] can achieve precise H-doping by tuning the wire temperature, demonstrating the superior features over the uncontrollable traditional high-temperature or high-pressure H2 hydrogenations.

3.1.2 Hydrogenation-PEC performance relation

In this section, we examine in detail the effects of wire temperature on the structural, optical, electrical, and PEC properties of black TiO2.

3.1.2.1 Structure

Scanning electron microscopy (SEM) and HR-TEM were used to investigate the microstructure of H-TiO2 nanorods treated at different wire temperatures. H-TiO2 treated at Twire = 1700°C closely resembles pristine TiO2 in morphology. HR-TEM images show that pristine TiO2 has a single crystalline structure (Figure 5ac), while H-TiO2 at Twire = 1700°C has a core-shell structure with a 2 nm thick disordered shell (Figure 5df). Electron energy loss spectroscopy (EELS) analysis showed that the [O]/[Ti] ratio in pristine TiO2 remains constant at about 2 throughout the nanorod, while in H-TiO2 at Twire = 1700°C, the ratio decreases gradually in the disordered region.

Figure 5.

TEM, HR-TEM images, and fast Fourier transform (FFT) patterns of (a–c) pristine TiO2 and (d–f) H-TiO2 nanorods treated at Twire = 1700°C.

3.1.2.2 Optical properties

The optical absorption spectra of pristine TiO2 and H-TiO2 were analyzed to study the improved optical absorption properties of H-TiO2. Results show a red shift in the band edge of H-TiO2 with increasing wire temperatures and increased absorption in the visible spectrum (Figure 6a). Tauc plots indicate a band edge of 2.9 eV for H-TiO2 compared to 3.0 eV for pristine TiO2 (Figure 6b).

Figure 6.

(a) Optical absorption and (b) Tauc plots for pristine TiO2 and H-TiO2.

3.1.2.3 PEC properties

J-V measurements were conducted to investigate the improved PEC properties of H-TiO2 nanorods (Figure 7a). The comparison of J-V curves of pristine TiO2 and H-TiO2 samples showed an increase in photocurrent density after treatment at Twire = 1600°C, peaking at Twire = 1700°C, and decreasing at Twire = 1800°C. The effect of wire temperature on the conductivity of H-TiO2 was studied using Mott-Schottky measurements (Figure 7b). Higher wire temperatures led to a decrease in the slope of H-TiO2, indicating an increase in donor density and conductivity. The donor densities of H-TiO2 were found to be 3× higher than those of pristine TiO2, showing a trend of increased conductivity in H-TiO2 samples with higher temperatures.

Figure 7.

(a) J-V curves of pristine TiO2 and H-TiO2 nanorods in dark and under solar illumination. (b) Mott-Schottky plots of pristine TiO2 and H-TiO2 nanorods. (c) O1s and (d) Ti2p XPS spectra of pristine and H-TiO2 samples.

3.1.2.4 Ti3+ identification

X-ray photoelectron spectra (XPS) analysis of H-TiO2 samples at different wire temperatures showed that the photocurrent density was highest at Twire = 1700°C. The O1s spectra revealed peaks at 530 and 531.5 eV, corresponding to TiO2 and Ti-OH species (Figure 7c). The intensity of the Ti∙OH peak increased with increasing wire temperature up to 1700°C but decreased at 1800°C. The reduced Ti-OH peak in H-TiO2 at 1800°C suggests less Ti3+ presence. XPS Ti2p spectra for both samples show similar characteristics, with binding energies of 458.5 eV for Ti2p3/2 and 464.3 eV for Ti2p1/2, typical of TiO2 (Figure 7d). A Ti-H peak at 456.5 eV, rather than metallic Ti at 453.8 eV, is observed in the sample treated at Twire = 1800oC. The formation of Ti‒H bonds in H-TiO2 at 1800°C reduces Ti∙OH bonds and decreases Ti3+ concentration, confirming the insight into improved photocurrent density of H-TiO2 at Twire = 1700°C.

The results suggest that the low-temperature H hydrogenation process effectively facilitates the formation of a disordered shell on the H-TiO2 nanorods, resulting in a decrease in band gap (3.0 → 2.9 eV), an increase in donor density (3×), and Ti3+ concentration, leading to the highest photocurrent density of 2.5 mA cm−2 at 1.23 VRHE of H-TiO2 at Twire = 1700°C reported to date.

3.1.3 Formation mechanism

The low-temperature H hydrogenation process is delineated in the subsequent three stages: (a) hydroxylation of the TiO2(110) surface through hydrogen adsorption; (b) subsurface diffusion of hydrogen; (c) subsurface hydrogenation.

Density functional theory (DFT) simulations suggest that H treatment facilitates the occupation of O3c surface sites by H, subsurface diffusion, and subsurface hydrogenation through the O3c → Osub low-energy-barrier pathway (0.87 eV), leading to the thermodynamically favorable formation of a disordered shell. Conversely, H2 treatment typically results in only submonolayer degrees of hydroxylation, with the occupation of O3c sites being thermodynamically unfavorable. Consequently, subsurface diffusion from O2c to Osub is kinetically hindered during H2 treatment; only H2O desorption and the formation of oxygen vacancies are deemed thermodynamically feasible, necessitating high pressure and/or high temperatures for the process to occur (Table 1).

[H]H2
HydroxylationAlmost no barrier for adsorption of atomic hydrogen → low temperature is sufficient H2 desorption is suppressed due to high-energy barrierHigh barrier for dissociative adsorption molecular hydrogen → high temperature and flux are needed. High surface coverages are not thermodynamically favorable
Subsurface diffusionLower barrier for subsurface diffusion pathway is O3c → Osub subsurface diffusion is faster than H2 and H2O desorption at low temperaturesHigh barrier for subsurface diffusion pathway is O2c → O3c → Osub subsurface diffusion is very slow
Subsurface H-dopingIncorporation of H in TiO2 lattice is thermodynamically favorable up to high concentrationsIncorporation of H in TiO2 lattice is not thermodynamically favorable. At high temperatures, incorporation becomes even less favorable
Formation of disordered shellFormation of disordered shell is thermodynamically favorable through O3c → Osub → H-doping in TiO2 lattice with relatively low barriers. Formation pathway: low temperatures are enoughO2c → O3c → Osub → H-doping in TiO2 is not thermodynamically favorable and kinetically suppressed. Different formation pathway: only H2O desorption and formation of oxygen vacancies are thermodynamically favorable → high temperatures are needed

Table 1.

Different formation pathways of disordered shell: H vs. H2.

In conclusion, we have experimentally and theoretically confirmed that our low-temperature H hydrogenation method is indeed superior to the traditional H2 hydrogenation method: the highest photocurrent density and highly efficient hydrogenation are achieved.

3.2 Room-temperature H+ hydrogenation

3.2.1 Experimental set-up

Recent interest in H+ hydrogenation using kinetic hydrogen ion species has increased because it does not require thermal activation. Traditional methods of H+ hydrogenation are limited by the need for high temperatures or high power output from plasma equipment to generate high-energy hydrogen ions, leading to inefficient and uncontrollable processes as well as limited photocurrent density of <1 mA cm−2 at 1.23 VRHE [13].

The magnetic field from an electromagnetic coil affects hydrogen ions and free electrons differently. Electrons change direction, but ions are not affected due to their greater mass (1000×). The magnetic field causes electrons to orbit in a circular path, increasing the chance of ionization and creating more hydrogen ions.

Inspired by the physical picture above, unlike traditional plasma hydrogenation methods, we used a circular electromagnetic coil in a radio frequency (RF) plasma physical vapor deposition (PVD) system to create a magnetic field (Figure 8). This field increases hydrogen ion concentration and decreases RF power and self-bias voltage, allowing for controlled low-energy hydrogen ion production for room-temperature H+ hydrogenation [11].

Figure 8.

Experimental set-up of room-temperature H+ hydrogenation. The role of a circular electromagnetic coil in the H+ hydrogenation process.

A programmable recipe controlled the RF power, magnetic field strength, hydrogen flow rate, processing pressure, and treatment duration for the H+ hydrogenation process. This included using RF power of 20 W, a magnetic field strength of 30 mT, a hydrogen flow rate of 50 sccm, and a processing pressure of 1.5 Pa to achieve a low self-bias voltage of −250V at the substrate electrode.

3.2.2 Hydrogenation-PEC performance relation

H-TiO2 nanorods were synthesized at different time intervals (2.5, 5, 10, 20, and 40 mins) to investigate their structural, optical, electrical, and PEC properties.

3.2.2.1 Structure

SEM was used to study morphological changes in TiO2 before and after H+ treatment. No significant changes were observed in SEM images of pristine TiO2 and H-TiO2 treated at 5 and 40 mins. X-ray diffraction (XRD) patterns showed only a rutile phase with (101) and (002) peaks, indicating no phase change in either sample. HR-TEM images show that H-TiO2 samples have a composite structure with a crystalline core and a disordered shell. The core remains single crystalline, with the shell thickness increasing from 2 to 5 nm as treatment time increases from 5 to 40 mins (Figure 9). The line-scan EELS data of [O]/[Ti] ratio changes in pristine TiO2 and H-TiO2 samples. Oxygen vacancies form in the shell of H-TiO2 over time but not in the core of H-TiO2 treated for 5 mins.

Figure 9.

TEM, HR-TEM, and FFT images of H-TiO2 at (a–c) 5 and (d–f) 40 mins, respectively.

3.2.2.2 Optical properties

We recorded optical absorption spectra to compare the optical properties of pristine TiO2 and H-TiO2. The absorption band edges of H-TiO2 shift toward red and absorb more in the visible to infrared range with longer treatment time. The band gap decreases from 3.0 eV in pristine TiO2 to 2.91 eV in H-TiO2 after 5 mins of treatment.

3.2.2.3 PEC properties

Figure 10a shows that the photocurrent density of H-TiO2 increases up to 5 mins of treatment time, then decreases. At 5 mins, a photocurrent density of 2.55 mA cm−2 at 1.23 VRHE is achieved, the highest reported for H-TiO2 studies. Increasing treatment time increases donor density and decreases depletion region width. The negative shift in flat band potential of H-TiO2 at 2.5 and 5 mins is due to a higher donor density, shifting the Fermi level toward the conduction band. The largest band bending of 1.1 V is observed for H-TiO2 at 5 mins, determined from M-S plots (Figure 10b).

Figure 10.

(a) J-V curves of pristine TiO2 and H-TiO2 at different treatment times in dark and under solar illumination. (b) M-S plots of pristine TiO2 and H-TiO2 at different treatment times. (c) O1s and (d) Ti2p XPS spectra of pristine TiO2 and H-TiO2 at 5 and 40 mins, respectively.

3.2.2.4 Ti3+ identification

XPS analysis of the disordered shell showed the presence of Ti‒O and Ti‒OH in H-TiO2. Ti3+ was detected in H-TiO2 treated at 5 mins, with a strong reduction to Ti2+ observed in H-TiO2 treated at 40 mins (Figure 10c and d).

Our experiment results confirmed that the highest photocurrent density of 2.55 mA cm−2 at 1.23 VRHE of H-TiO2 at 5 mins is due to a smaller band gap (2.91 eV), increased band bending (1.1 eV) and bulk defect suppression.

3.2.3 Formation mechanism

We studied how the penetration ratio of vertically impinging H+ changes with initial kinetic energy Ekin,0. At the lowest Ekin,0 of 0.1 eV, all 15 H+ were reflected, consistent with the above DFT study of low-temperature H hydrogenation predicting a minimal energy barrier of 0.87 eV for subsurface diffusion. At Ekin,0 = 1 eV, 1 out of 15 H+ bonded to a surface O ion. At Ekin,0 = 10 eV, 8 out of 15 H+ penetrated the TiO2 surface. At the maximum velocity of 80 eV, 10 out of 15 H+ penetrated the TiO2 surface, resulting in a 67% penetration ratio. H+ with kinetic energies of at least 1 eV can penetrate the TiO2 surface without needing thermal activation. The penetration ratio for H+ with kinetic energies of at least 10 eV can reach approximately 0.5.

It is intriguing to compare how energetic H+ particles interact with the TiO2 surface versus thermal H species like H2 and H in hydrogenation processes. Impinging thermal H2 or H cannot directly penetrate; instead, hydrogen uptake in the subsurface region involves chemisorption and subsurface diffusion. Atomic H is more efficient than molecular H2 for subsurface diffusion due to its ability to adsorb at higher energy surface sites, resulting in a lower energy barrier. However, efficient hydrogenation only occurs at high substrate temperatures. In our previous studies, we found that the lowest energy barrier for the movement of a chemisorbed H atom into the subsurface area is 0.87 eV and occurs near an O3c site. In the H+ case, as the initial kinetic energy increases, additional penetration pathways with higher energy barriers surrounding O3c → Osub are available. This phenomenon results in a rapid increase in the penetration probability with increasing initial kinetic energy, ultimately contributing to the high efficiency of room-temperature H+ hydrogenation.

In summary, our study has experimentally and theoretically validated the superiority of our room-temperature H+ hydrogenation over the thermal hydrogen hydrogenation approaches, as evidenced by the attainment of the highest photocurrent density and exceptional efficiency in hydrogenation processes.

Advertisement

4. Comparison of low-temperature H and room-temperature H+ methods with state-of-the-art hydrogenations

Through a comparative analysis of the synthesis of black TiO2 using the traditional hydrogenation techniques and the subsequent evaluation of photocurrent density, our two advanced hydrogenation methods, low-temperature H and room-temperature H+, exhibit the highest reported photocurrent density (Table 2), while simultaneously effectively addressing the challenges associated with traditional H2 hydrogenation, such as high temperature, high pressure, and prolonged treatment, as well as chemical residues from chemical reductions.

Hydrogenation methodTemperatureRF powerTreatment time, pressurePhotocurrent density of H-TiO2 at 1.23 VRHEFeatures and references
High-pressure H2200°CN/A5 days, 20 barN/AX High pressure; prolonged treatment [5]
High-temperature H2400°CN/A1 h, 1 bar2.4 mA cm−2X High temperature; prolonged treatment [6]
Low-temperature H265°CN/A20 mins, 1 Pa2.5 mA cm−2✓ Low temperature; short treatment [10]
Noble metal-catalyzed H250°CN/A10 mins, 1 bar1.5 mA cm−2X High cost: noble metals; chemical residues [7]
High-temperature H+425°C200 W1 h0.9 mA cm−2X High temperature and power; prolonged treatment [13]
Room temperature H+25°C20 W5 mins, 1.5 Pa2.55 mA cm−2✓ Room temperature; short treatment [11]

Table 2.

A comparison of the low-temperature H hydrogenation and room-temperature H+ hydrogenation with state-of-the-art hydrogenations for PEC applications.

Table 3 shows that the hydrogenation conditions of the hydrogenation technologies recently applied to hydrogen fuel cells and hydrogen storage are not suitable for practical applications, and it is evident that the two advanced hydrogenation technologies can compensate for their shortcomings.

Hydrogenation methodTemperatureTreatment time, pressureApplicationsFeatures and references
Noble metal-catalyzed H250°C10 minsHydrogen fuel cellsX High cost: noble metals; chemical residues [8]
High-pressure H2400°C12 h, 0.5 MPaHydrogen storageX High pressure; prolonged treatment [9]

Table 3.

Reported hydrogenations for hydrogen fuel cells and storage applications.

Advertisement

5. Conclusions

In summary, we highlighted the advancements in surface engineering-hydrogenation technology for use in PEC hydrogen production, hydrogen fuel cells, and hydrogen storage. We presented the operational principles and experimental set-up of our invented hydrogenation technologies, and subsequently investigated the correlation between hydrogenation parameters and the structural, optical, electrical, and PEC properties of materials through combined experimental and theoretical studies. Finally, we conducted a comparative analysis of the benefits and drawbacks of the two hydrogenation technologies in relation to reported hydrogenation technologies. Our findings suggest that these two advanced technologies hold significant potential for widespread applications in various hydrogen-related technologies, such as PEC hydrogen production, hydrogen fuel cells, and hydrogen storage.

Advertisement

Acknowledgments

X. Wang acknowledges the support of the research fund by the Institute of Carbon Neutrality, ShanghaiTech University. H. Shen and B. Wang thank the support of the research fund by Yiwu Research Institute, Fudan University.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Yang W, Prabhakar RR, Tan J, Tilley SD, Moon J. Strategies for enhancing the photocurrent, photovoltage, and stability of photoelectrodes for photoelectrochemical water splitting. Chemical Society Reviews. 2019;48(19):4979-5015. DOI: 10.1039/C8CS00997J
  2. 2. El Ouardi M, El Idrissi A, Arab M, Zbair M, Haspel H, Saadi M, et al. Review of photoelectrochemical water splitting: From quantitative approaches to effect of sacrificial agents, oxygen vacancies, thermal and magnetic field on (photo)electrolysis. International Journal of Hydrogen Energy. 2024;51:1044-1067. DOI: 10.1016/j.ijhydene.2023.09.111
  3. 3. Fujishima A, Honda K. Electrochemical photolysis of water at a semiconductor electrode. Nature. 1972;238(5358):37-38. DOI: 10.1038/238037a0
  4. 4. Zhang X, Zhang S, Cui X, Zhou W, Cao W, Cheng D, et al. Recent advances in TiO2-based photoanodes for photoelectrochemical water splitting. Chemistry, an Asian Journal. 2022;17(20):e202200668. DOI: 10.1002/asia.202200668
  5. 5. Chen XB, Liu L, Yu PY, Mao SS. Increasing solar absorption for photocatalysis with black hydrogenated titanium dioxide nanocrystals. Science. 2011;331(6018):746-750. DOI: 10.1126/science.1200448
  6. 6. Wang GM, Wang HY, Ling YC, Tang YC, Yang XY, Fitzmorris RC, et al. Hydrogen-treated TiO2 nanowire arrays for photoelectrochemical water splitting. Nano Letters. 2011;11(7):3026-3033. DOI: 10.1021/nl201766h
  7. 7. Xu Y, Zhang C, Zhang L, Zhang X, Yao H, Shi J. Pd-catalyzed instant hydrogenation of TiO2 with enhanced photocatalytic performance. Energy & Environmental Science. 2016;9(7):2410-2417. DOI: 10.1039/C6EE00830E
  8. 8. Lu N, Zhang Z, Wang Y, Li H-B, Qiao S, Zhao B, et al. Enhanced low-temperature proton conductivity in hydrogen-intercalated brownmillerite oxide. Nature Energy. 2022;7(12):1208-1216. DOI: 10.1038/s41560-022-01166-8
  9. 9. Izumi Y, Takeiri F, Okamoto K, Saito T, Kamiyama T, Kuwabara A, et al. Electropositive metal doping into lanthanum hydride for H−conducting solid electrolyte use at room temperature. Advanced Energy Materials. 2023;13(43):2301993. DOI: 10.1002/aenm.202301993
  10. 10. Wang X, Mayrhofer L, Hoefer M, Estrade S, Lopez-Conesa L, Zhou H, et al. Facile and efficient atomic hydrogenation enabled black TiO2 with enhanced photo-electrochemical activity via a favorably low-energy-barrier pathway. Advanced Energy Materials. 2019;9(33):1900725. DOI: 10.1002/aenm.201900725
  11. 11. Wang X, Mayrhofer L, Keunecke M, Estrade S, Lopez-Conesa L, Moseler M, et al. Low-energy hydrogen ions enable efficient room-temperature and rapid plasma hydrogenation of TiO2 nanorods for enhanced photoelectrochemical activity. Small. 2022;18(46):2204136. DOI: 10.1002/smll.202204136
  12. 12. Langmuir I. A chemically active modification of hydrogen 5. Journal of the American Chemical Society. 1912;34(10):1310-1325. DOI: 10.1021/ja02211a004
  13. 13. Tian Z, Cui H, Zhu G, Zhao W, Xu J, Shao F, et al. Hydrogen plasma reduced black TiO2-B nanowires for enhanced photoelectrochemical water-splitting. Journal of Power Sources. 2016;325:697-705. DOI: 10.1016/j.jpowsour.2016.06.074

Written By

Xiaodan Wang, Beibei Wang and Hao Shen

Submitted: 22 February 2024 Reviewed: 10 March 2024 Published: 21 May 2024