Next Article in Journal
Effect of Temperature-Dependent Low Oxygen Partial Pressure Annealing on SiC MOS
Next Article in Special Issue
Boosting Piezocatalytic Performance of BaTiO3 by Tuning Defects at Room Temperature
Previous Article in Journal
Integrated Ozonation Ni-NiO/Carbon/g-C3N4 Nanocomposite-Mediated Catalytic Decomposition of Organic Contaminants in Wastewater under Visible Light
Previous Article in Special Issue
Enhanced Catalytic Performance of Ag NP/0.95AgNbO3-0.05LiTaO3 Heterojunction from the Combination of Surface Plasma Resonance Effect and Piezoelectric Effect Using Facile Mechanical Milling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Geometrical Stabilities and Electronic Structures of Rh5 Nanoclusters on Rutile TiO2 (110) for Green Hydrogen Production

Department of Physics, College of Science and Humanities in Al-Kharj, Prince Sattam Bin Abdulaziz University, Al-Kharj 11942, Saudi Arabia
Nanomaterials 2024, 14(2), 191; https://doi.org/10.3390/nano14020191
Submission received: 28 December 2023 / Revised: 10 January 2024 / Accepted: 13 January 2024 / Published: 15 January 2024

Abstract

:
Addressing the urgent need for sustainable energy sources, this study investigates the intricate relationship between rhodium (Rh5) nanoclusters and TiO2 rutile (110) surfaces, aiming to advance photocatalytic water splitting for green hydrogen production. Motivated by the imperative to transition from conventional fossil fuels, this study employs density functional theory (DFT) with DFT-D3 and HSE06 hybrid functionals to analyse the geometrical stabilities and electronic structures of Rh5 nanoclusters on TiO2 rutile (110). TiO2, a prominent photocatalyst, faces challenges such as limited visible light absorption, leading researchers to explore noble metals like Rh as cocatalysts. Our results show that bipyramidal Rh5 nanoclusters exhibit enhanced stability and charge transfer when adsorbed on TiO2 rutile (110) compared to trapezoidal configurations. The most stable adsorption induces the oxidation of the nanocluster, altering the electronic structure of TiO2. Extending the analysis to defective TiO2 surfaces, this study explores the impact of Rh5 nanoclusters on oxygen vacancy formation, revealing the stabilisation of TiO2 and increased oxygen vacancy formation energy. This theoretical exploration contributes insights into the potential of Rh5 nanoclusters as efficient cocatalysts for TiO2-based photocatalytic systems, laying the foundation for experimental validations and the rational design of highly efficient photocatalysts for sustainable hydrogen production. The observed effects on electronic structures and oxygen vacancy formation emphasize the complex interactions between Rh5 nanoclusters and the TiO2 surface, guiding future research in the quest for clean energy alternatives.

1. Introduction

Contemporary chemistry is characterized by a heightened focus on addressing global challenges related to energy production and environmental remediation. The pursuit of innovative energy alternatives, driven by the pressing need to move beyond conventional fossil fuels, has become a central theme in scientific research. This shift towards sustainable practices is not only motivated by environmental consciousness but also by stringent environmental regulations that highlight the necessity of adopting eco-friendly approaches. A pivotal contribution to this field is exemplified by the groundbreaking work of Fujishima and Honda, who showed the viability of photoelectrochemical water splitting using rutile TiO2 as an anode and a Pt wire cathode under a chemical bias [1]. Beyond its role in water splitting, TiO2, a prominent photocatalyst, has garnered significant attention for its applications in the decomposition of harmful organic materials [2,3,4,5]. This multifaceted functionality extends TiO2 photocatalysis to diverse environmental challenges, showcasing their effectiveness in both gaseous environments and solutions. This versatility aligns with the broader trend in contemporary chemistry, emphasizing interdisciplinary research approaches aimed at developing sustainable solutions for the interconnected issues of energy and the environment. The collaborative efforts of scientists and researchers across disciplines underscore the significance of integrated approaches in meeting the complex challenges of the modern era.
In photocatalytic water splitting [6], the role of TiO2 is pivotal but not without its challenges. TiO2 is commonly employed as a photocatalyst, and to intensify the efficiency of the hydrogen evolution reaction (HER), it is a prevalent practice to load suitable cocatalysts, often metallic in nature [7,8]. Upon photon excitation, TiO2 generates photoelectrons in its bulk region, underscoring the importance of cocatalysts that excel in two key aspects: (i) facilitating improved charge transfer over the interface from catalyst to the metal and (ii) ensuring fast H2 generation on the surface of metal [9,10]. However, TiO2 faces intrinsic challenges [11], including a band gap energy (Eg) of around 3.2 eV [12,13], that confines its light absorption primarily to ultraviolet (UV) wavelengths. Additionally, a notable concern arises from the significant degree of charge recombination for photo-generated charges in TiO2. In response to these challenges, researchers have implemented various strategies to broaden the application of TiO2 in photocatalysis. For instance, efforts have been made to integrate TiO2 with other semiconductor materials [14,15], a technique aimed at mitigating limitations associated with its band gap and extending its responsiveness to a broader spectrum of light. Furthermore, innovative approaches such as dye sensitization in solar cells have been explored [16], providing alternative avenues to enhance the performance of TiO2 in capturing and utilizing solar energy, fabricating with both metallic and non-metallic ions [17,18] and depositing noble metals [19]. These strategies exemplify the ongoing endeavours within the scientific community to overcome the inherent limitations of TiO2 and advance the field of photocatalysis for sustainable energy production.
Noble metals, such as Pd (palladium), Pt (platinum), and Rh (rhodium), have emerged as valuable catalysts in the realm of photocatalytic hydrogen evolution reaction (HER) due to their notable work functions and favourable Gibbs adsorption energies for hydrogen atoms [20,21]. These metals, with their distinctive properties, play a significant role in enhancing the efficiency of hydrogen evolution during photocatalysis. For example, the incorporation of Rh into TiO2 has been explored as a strategy to boost photoreactivity. Rh-doped TiO2 has demonstrated enhanced performance, a phenomenon attributed to the facilitated electron transfer between Rh and the TiO2 conduction band (CB) or valence band (VB). Studies have shown that TiO2 samples decorated with Rh exhibit superior activity compared to alternative modifications, emphasizing the effectiveness of noble metal doping in optimizing the photocatalytic properties of TiO2 [22,23]. This underscores the significance of exploring and understanding the synergistic effects between noble metals and semiconductor materials to further advance the development of effective photocatalysts for applications in sustainable energy.
Xing et al. [24] made notable contributions to the field of photocatalysis by employing a single-step approach to synthesise isolated metal atoms stably loaded on the TiO2 anatase (101) surface, thereby extending the concept of single atom catalysts to the domain of photocatalytic hydrogen production. This groundbreaking approach involves the deposition of single atoms of noble metals, including Pd, Pt, Rh, and Ru, uniformly on the TiO2 anatase (101) surface. The resulting catalysts exhibit remarkably improved photocatalytic performance, particularly in the context of hydrogen evolution. This innovative methodology opens new avenues for the design of highly efficient and stable photocatalysts. The work by Xing et al. underscores the importance of exploring diverse approaches to advance the understanding and application of single atom catalysts in photocatalysis. While Pt, Pd, Rh, and Ru nanoclusters have found practical applications [25,26], it is noteworthy that a comprehensive theoretical analysis of Rh nanocluster-loaded TiO2 rutile (110)-based photocatalytic systems is currently lacking. Further theoretical investigations into the unique properties and behaviours of these systems could provide valuable insights into their photocatalytic mechanisms and guide the development of advanced materials for sustainable energy applications.
In the current study, an examination of the pristine and reduced TiO2 rutile (110) surface loaded with Rh5 nanoclusters is conducted using density functional theory (DFT). Upon a reduction in the size of the Rh cluster, it exhibits a more substantial HOMO-LUMO gap relative to larger clusters of Rh [27]. The phenomenon of energy band discretisation in metal clusters is notably pronounced and size-dependent, resulting in the emergence of substantial band gaps in proximity to the Fermi level. These gaps often surpass one electronvolt (eV) and play a pivotal role in dictating the excitation–emission characteristics within clusters. Such a feature paves the way for the strategic design of metal clusters, capitalising on their distinctive luminescence properties across the ultraviolet-visible–infrared (UV-Vis-IR) spectrum for tailored optical applications [28]. Furthermore, an investigation into the low-index faces of rutile revealed that, in accordance with the composition found in natural rutile powder, the (110) face exhibits the highest stability, followed by the (100) and (101) faces in terms of their relative stability [29,30]. Therefore, the TiO2 rutile (110) surface was chosen in this study. The choice of employing the DFT-D3 method [31] is rooted in its effectiveness in characterising the adsorption behaviour of Rh nanoclusters on rutile TiO2. To further elucidate the electronic structure relevant to polaron formation on TiO2 surfaces [32,33], this study utilises the HSE06 hybrid functional, a theoretical framework developed by Heyd, Scuseria, and Ernzerhof [34,35]. This hybrid functional incorporates a fraction of exact exchange, improving the description of electronic properties compared to standard DFT methods. This article is organised as follows: Section 2 provides comprehensive details on the simulation methodologies employed, offering transparency and reproducibility in the research process. Section 3 is dedicated to presenting and discussing the outcomes of the simulations, shedding light on the interactions and behaviours observed in the system under investigation. It also delves into the concept of polaron and compares the obtained results with previous findings, contributing to the theoretical understanding of charge carriers on TiO2 surfaces. Finally, Section 4 outlines the principal results, providing a pragmatic perspective on their implications for the broader scientific community and applications in renewable energy.

2. Computational Details

To explore the electronic properties and charge density of Rh5 nanoclusters and understand their photon absorption capabilities, we employed the Vienna Ab initio Simulation Package (VASP 5.4.4) [36,37,38]. This allowed us to derive optimised geometries and electronic structures for both bare Rh5 nanoclusters and Rh5@TiO2. The simulations utilised the HSE06 hybrid exchange–correlation functional with periodic boundary conditions, encompassing short-range and long-range elements of the Perdew–Burke–Ernzerhof (PBE) exchange functional. The exchange–correlation component incorporates short-range Hartree–Fock (HF) exchange and a PBE correlation functional [34].
The interplay between valence electrons and the ion core was clarified through the utilisation of the projector-augmented wave (PAW) approach. [39,40], employing PAW-PBE [41] pseudopotentials. Valence electrons from Ti (3s, 3p, 4s, 3d), O (2s, 2p), and Rh (4d, 5s) atomic orbitals were considered. To rectify the self-interaction error and ensure accurate predictions of polaronic states and the band gap of TiO2, we incorporated the generalised-gradient approximation (GGA) with a Hubbard parameter (U) [42]. The assigned U value for the titanium (3d) state study was 4.2 eV, as documented in the literature [43,44]. Spin-polarized Perdew–Burke–Ernzerhof (PBE) with the Becke–Jonson (BJ) damping function, following Grimmme’s technique [31], was applied for van der Waals (vdW) corrections, selected for its accuracy in predicting the adsorption energy of metal oxide materials [45].
To model the pristine rutile TiO2 (110) surface, we constructed unit cells with dimensions of 12 Å × 13 Å, consisting of four O-Ti-O trilayers. A 20 Å vacuum layer was added above the surface. For individual Rh5 nanoclusters without any interaction with periodic images, we utilised large supercells (30 × 30 × 30 Å3). A k-point mesh was employed, adhering to the Monkhorst–Pack scheme [46], wherein all simulations were conducted utilising a singular k-point value, employing a fixed planewaves basis set of 500 eV. A Gaussian smearing parameter of 0.05 eV was applied for band occupation due to the substantial supercell used in the tetrahedron. To achieve self-consistent electronic minimisation, we set a convergence threshold of 104 eV. The convergence criterion of 104 eV was chosen based on common practices in computational studies [47,48,49,50,51]. This criterion ensures adequate precision in the calculations while balancing computational resources, and all modelled structures underwent relaxation with a force threshold value of 0.02 eV/Å.
The stability of catalysts during chemical reactions poses a crucial challenge for practical applications. Consequently, we computed the adsorption energy ( E a d s ) of the Rh5 nanocluster to assess its stability during adsorption, as per the following formula:
E a d s = E t o t E T i O 2 E R h 5
Here, E t o t is the total energy of the whole system, E T i O 2 is the total energy of TiO2, and E R h 5 is the total energy of Rh5 nanoclusters. In the context of defects, the oxygen vacancy ( E V o ) formation energy was determined through the following formula:
E V o = E s u r f a c e + V o + 1 2 E O 2 E s u r f a c e
This involves E s u r f a c e + V o , representing the total energy of the reduced TiO2, E O 2 denoting the total energy of free oxygen in the gas phase, and E s u r f a c e representing the total energy of the perfect TiO2. Both adsorption and formation energies were assessed using GGA + U calculations. The construction and visualisation of all structures presented in this study were performed using (VESTA 3.5.8) software [52].

3. Results and Discussion

3.1. Isolated Rh5 Nanoclusters

Figure 1a and Figure 1b illustrate the optimised structures of Rh5 nanoclusters in the gas phase, showcasing bipyramidal and trapezoidal shapes, respectively. Both structures, representing the doublet state, indicate that the bipyramidal configuration (Figure 1a) is more stable in the gas phase compared to the trapezoidal shape (Figure 1b), with an energy difference of 1.38 eV. The bipyramidal Rh5 nanoclusters in Figure 1a deviate from perfect D3h symmetry, featuring equatorial Rh atoms forming a triangular ring and axial Rh atoms located above and below the ring. Notably, the equatorial atoms exhibit inequivalence, with three bonds shaping the triangle (d1), measuring 2.48 Å, while the remaining bond (d4) is 2.50 Å. The bonds d2 and d3, formed by axial Ag with the three equatorial sites measure 2.47 Å for both (see Table 1 for detailed information). The unpaired electron in the Rh5 doublet state (with S = ½), as depicted in Figure 1c, primarily localises on the two axial Rh atoms, resembling the charge distribution observed in Cu5 [53] and Ag5 nanoclusters [33]. According to the density of states analysis in Figure 1, the calculated band gaps of bipyramidal and trapezoidal Rh5 nanoclusters are 1.15 eV and 0.57 eV, respectively.

3.2. Bipyramidal Rh5 Nanocluster Loaded on TiO2

As part of a benchmark analysis, we computed the electronic density of states for the pristine TiO2 rutile (110), as shown in Figure S1. Our calculations yielded an estimated band gap value of approximately 3.2 eV, aligning well with experimental findings [12]. After examining the geometrical and electronic characteristics of bare Rh5 nanoclusters, our focus shifted to an investigation of the corresponding attributes in Rh5 nanoclusters adsorbed into both pristine and reduced TiO2 (110) surfaces. Three distinct adsorption sites of the bipyramidal Rh5 nanocluster on TiO2 are scrutinized, as depicted in Figure 2. Notably, the configuration illustrated in Figure 2a demonstrates superior stability, evidenced by an adsorption energy of −5.28 eV in comparison to the other two configurations, with an average Rh-O bond length of 2.10 Å (see Table 2). Conversely, the structure presented in Figure 2b exhibits the least stability, marked by an adsorption energy of −4.78 eV. The configuration in Figure 2c manifests a metastable state with an adsorption energy of −4.84 eV. The discerned disparities in stability may be ascribed to the nature of the Rh-O bonds; specifically, in the most stable configuration (Figure 2a), four Rh atoms are bonded to four O atoms, whereas in the remaining two structures, only three Rh atoms form bonds with three O atoms.
Additionally, to enhance our understanding of the observed adsorption patterns, we focus on the analysis of charge transfer. In the case of the most stable configuration, the Rh5 nanocluster exhibits a charge transfer of approximately +0.6 e to TiO2. This observed electron transfer implies an oxidation state for the Rh5 nanocluster, corroborating findings from previous studies [51,54,55]. To evaluate the influence of the Rh5 nanocluster on the electronic structures of the TiO2 rutile (110) surface, we conducted density of states calculations applying the HSE06 functional and wavefunction computations for the most stable configuration, as depicted in Figure 2a and presented in Figure 3. Our findings indicate that incorporating a bipyramidal Rh5 nanocluster into the TiO2 rutile (110) surface results in the creation of mid-gap states in the band gap.
As an illustration, the highest occupied molecular orbital (HOMO) exhibits a high-energy state situated at −0.23 eV, roughly 0.72 eV below the CB edge. The introduction of mid-gap states is a result of charge transfer from the Rh5 nanocluster to the TiO2 surface. These intermediary states play a pivotal role in absorbing photons within the visible and UV regions. Furthermore, the deposition of the Rh5 nanocluster on TiO2 results in the repopulation of the CB, initiating a manifestation of metallic characteristics within the system. Similar findings have been documented for a TiO2 system when exposed to Ag3 and Ag5 clusters [56]. In the context of the visible-light spectrum, it becomes apparent that mid-gap states can accept electrons from the VB. The energetic nature of visible-light irradiation facilitates electron transfer due to the diminished energy separation between intra-gap states and the VB. This electron transition potentially contributes to the augmentation of photocatalytic hydrogen production [57]. For example, Wang et al. [58] experimentally reported that the photocatalytic activity for hydrogen evolution using Rh-doped rutile demonstrated an approximate fiftyfold increase in efficiency compared to that observed with Rh-doped anatase powders.

3.3. Trapezoidal Rh5 Nanocluster Loaded on TiO2

This investigation involves the computational simulation of three distinct adsorption configurations of trapezoidal Rh5 nanoclusters on the TiO2 rutile (110) surface, specifically adopting upstanding, tilted, and lying-down orientations, as illustrated in Figure 4. Analysis of the simulation reveals notable distortions in the upstanding and tilted Rh5 nanoclusters (see Figure 4a,d and Figure 4b,e) upon their adsorption onto the TiO2 surface, leading to diminished stability, characterised by an average Rh-O bond length of approximately 2.02 Å. A higher adsorption energy is discerned in the case of a slight tilt in the Rh5 nanocluster towards the TiO2 surface, amounting to approximately 0.01 eV. Conversely, when the Rh5 nanocluster assumes a parallel orientation to the TiO2 surface (depicted in Figure 4c,f), a considerably higher adsorption energy of approximately −6.46 eV is observed, indicative of enhanced stability in comparison to the upstanding and tilted configurations. This trend mirrors findings from prior DFT studies on trapezoidal Ag5 and Cu5 adsorbed on TiO2 rutile (110) [33,53]. Additionally, a pronounced distortion is evident on the TiO2 surface directly beneath the loaded Rh5 nanocluster. Table 3 provides a comparative analysis of the adsorption energies and charges associated with the various Rh5 nanocluster structures depicted in Figure 4.
The electronic characteristics of the most stable structure of trapezoidal Rh5@TiO2 are subjected to an in-depth analysis through the density of states and analysis of Bader charge, as presented in Figure 4 and Table 3, respectively. Bader charge analysis reveals that all trapezoidal Rh5 nanoclusters exhibit electron donation to the TiO2 surface, inducing oxidation. Interestingly, a noticeable correlation is observed, where less charge transfer from the Rh5 nanocluster to the catalyst correlates with higher stability, while augmented charge transfer corresponds to less stability. This trend contrasts with the behaviour noted in the loading of Ag5 clusters on the TiO2 rutile (110) surface [33]. Furthermore, the adsorption of the trapezoidal Rh5 nanocluster induces notable alterations in the electronic characteristics of the pristine TiO2, generating mid-gap states within the band gap, as illustrated in Figure 5. The density of states analysis delineates that the HOMO state of the Rh5 nanocluster is situated approximately 1.2 eV below the CB. To conclude this section, the simulation results show that the bipyramidal Rh5 nanocluster exhibits superior efficacy in enhancing the photocatalytic activity of TiO2 rutile (110) compared to the trapezoidal Rh5 nanocluster. This is substantiated by the energy difference, with the most stable configuration of the bipyramidal Rh5 nanocluster registering a –0.2 eV reduction compared to the most stable structure of the trapezoidal Rh5 nanocluster.

3.4. Bipyramidal Rh5 Nanocluster Loaded on Defective TiO2

To investigate the influence of the Rh5 nanocluster on the generation of an oxygen vacancy on the TiO2 rutile (110) surface, we initially present results related to defective TiO2 rutile (110). In our prior DFT calculations [33], it has been demonstrated that the formation energy of a surface oxygen vacancy on pristine TiO2 rutile (110) is lower than that of the subsurface by approximately 0.6 eV (see Figure S2 and Table S1), aligning with previous studies [59,60]. Subsequently, with reference to the most stable configuration of Rh5 loaded on TiO2, as depicted in Figure 2a, an exploration into the impact on the photocatalytic activity concerning surface oxygen vacancy was conducted. The investigation reveals that the introduction of the Rh5 nanocluster stabilises TiO2 rutile (110), leading to an elevation in the formation energy of both surface and subsurface oxygen vacancies by 0.44 eV and 0.17 eV, respectively (see Figure S3 and Table S1 for further comparative analysis).
To explore the electronic characteristics of a Rh5 nanocluster adsorbed onto reduced TiO2 rutile (110), we conducted density of states and wavefunction calculations, as shown in Figure 6. The figure indicates that the combined presence of the Rh5 nanocluster and an oxygen vacancy introduces additional gap states. Notably, the CB edge undergoes a significant downward shift towards lower energy levels, resulting in an elevated energy HOMO state of the Rh5 nanocluster by approximately 0.3 eV from its position. Remarkably, the state appearing at –0.72 eV corresponds to the singly occupied molecular orbital (SOMO), situated on a Ti61 atom on the surface of TiO2, exhibiting an electron gain of approximately 0.3 e and giving rise to a polaronic state (as depicted by SOMO in Figure 6). Furthermore, our investigation shows an electron donation from the Rh5 nanocluster to the material amounting to +0.3 e, representing a reduction of half compared to the configuration without the oxygen vacancy (i.e., the configuration illustrated in Figure 2a). The formation of the polaronic state is identified as a significant factor contributing to the absorption of visible-light photons [61]. In summary, the reciprocal presence of the Rh5 nanocluster and the oxygen vacancy synergistically enhances the photocatalytic activity of the substrate. Consequently, both the Rh5 nanocluster and the oxygen vacancy emerge as potential catalysts for water splitting, offering promising insights for the systematic design of highly efficient photocatalysts dedicated to photocatalytic hydrogen generation.

4. Concluding Remarks

This article delves into the investigation of geometrical stabilities and electronic characteristics of Rh5 nanoclusters on the TiO2 rutile (110) surface, aiming for potential applications in green hydrogen production through photocatalytic water splitting. The research is motivated by the need for efficient and sustainable energy sources, particularly focusing on enhancing the photocatalytic performance of TiO2, a well-established photocatalyst. By employing DFT with the DFT-D3 technique and the HSE06 hybrid functional, we conducted a comprehensive examination of the adsorption behaviour, electronic structures, and charge transfer dynamics of Rh5 nanoclusters on TiO2 surfaces. The computational analysis explored the stability of Rh5 nanoclusters on both pristine and reduced TiO2 surfaces, along with their influence on oxygen vacancy formation. The results indicate that trapezoidal Rh5 nanoclusters exhibit superior stability and adsorption energy compared to bipyramidal structures when deposited on TiO2 rutile (110). The most stable adsorption structures of the bipyramidal Rh5 nanocluster resulted in a charge transfer of approximately +0.6 e to TiO2, inducing oxidation of the nanocluster. Electronic structure analysis reveals the generation of intra-gap states in the band gap of TiO2 upon Rh5 nanocluster deposition, suggesting potential implications for visible and ultraviolet photon absorption.
Additionally, this study investigates the interaction of Rh5 nanoclusters with defective TiO2 surfaces, with a specific focus on the oxygen vacancies’ evolution. The results indicate that the existence of Rh5 nanoclusters stabilizes TiO2 and increases the oxygen vacancies’ formation energy, hinting at a potential role in enhancing photocatalytic activity. To summarise, this theoretical investigation offers valuable insights into the potential of Rh5 nanoclusters as efficient cocatalysts for TiO2-based photocatalytic systems. The results propose that the bipyramidal configuration of Rh5 nanoclusters, when appropriately adsorbed on TiO2 rutile (110), may contribute to enhanced photocatalytic performance, providing opportunities for the rational design of highly efficient photocatalysts for green hydrogen production. The HOMO state being located at 0.72 eV below the CB edge (see Figure 3), which possesses high energy, can significantly benefit photocatalytic water splitting for green hydrogen production. This positioning of the HOMO level enhances the ability of the photocatalyst to transfer electrons effectively. During water splitting, electrons in the HOMO can be excited to the CB, leaving holes in the HOMO. These holes can then participate in the oxidation of water to produce oxygen. The excited electrons in the CB can reduce protons in water, generating hydrogen. Therefore, the position of the HOMO level is crucial for efficient photocatalytic activity, influencing the HER in water splitting processes. The observed effects on oxygen vacancy formation further underscore the intricate interplay between metal nanoclusters and the semiconductor surface, paving the way for future experimental validations and practical applications in sustainable energy production.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano14020191/s1, Figure S1: Density of states of pristine rutile TiO2 (110). The green and red curves show the electronic density of states on titanium and oxygen atoms, respectively. The black vertical dashed line shows the Fermi energy level. Reproduced from our previous calculations [33]; Figure S2: Oxygen vacancy formation at (a) surface and (b) subsurface locations of TiO2 rutile (110). The black circles represent the oxygen vacancy position. Reproduced from our previous calculations [33]; Figure S3: Oxygen vacancy formation at (a) surface and (b) subsurface locations of Rh5@TiO2 rutile (110). The black circles represent the oxygen vacancy position; Table S1: Comparisons of formation energies of oxygen vacancy.

Funding

This study is supported via funding from Prince Sattam bin Abdulaziz University project number (PSAU/2023/R/1445).

Data Availability Statement

Data are contained within the article.

Acknowledgments

The author extends his appreciation to the Deanship of Scientific Research at Prince Sattam bin Abdulaziz University, Alkharj in Saudi Arabia for funding this research work through project No. (PSAU/2023/R/1445). Moteb Alotaibi is thankful for computer time, this research used the resources of the Supercomputing Laboratory at King Abdullah University of Science & Technology (KAUST) in Thuwal, Saudi Arabia, (https://www.hpc.kaust.edu.sa/content/shaheen-ii accessed on 12 December 2023) and the High End Computing Cluster (HEC) at Lancaster University, UK.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 38–40. [Google Scholar] [CrossRef] [PubMed]
  2. Akira, F.; Rao, T.N.; Tryk, D.A. Titanium dioxide photocatalysis. J. Photochem. Photobiol. C 2000, 1, 1–21. [Google Scholar]
  3. Ochiai, T.; Fujishima, A. Photoelectrochemical properties of TiO2 photocatalyst and its applications for environmental purification. J. Photochem. Photobiol. C Photochem. Rev. 2012, 13, 247–262. [Google Scholar] [CrossRef]
  4. Maeda, K. Z-scheme water splitting using two different semiconductor photocatalysts. ACS Catal. 2013, 3, 1486–1503. [Google Scholar] [CrossRef]
  5. Ma, Y.; Wang, X.; Jia, Y.; Chen, X.; Han, H.; Li, C. Titanium dioxide-based nanomaterials for photocatalytic fuel generations. Chem. Rev. 2014, 114, 9987–10043. [Google Scholar] [CrossRef] [PubMed]
  6. Chen, X.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modifications and applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef]
  7. Yang, J.; Wang, D.; Han, H.; Li, C. Roles of cocatalysts in photocatalysis and photoelectrocatalysis. Acc. Chem. Res. 2013, 46, 1900–1909. [Google Scholar] [CrossRef]
  8. Ran, J.; Zhang, J.; Yu, J.; Jaroniec, M.; Qiao, S.Z. Earth-abundant cocatalysts for semiconductor-based photocatalytic water splitting. Chem. Soc. Rev. 2014, 43, 7787–7812. [Google Scholar] [CrossRef]
  9. Patrocinio, A.O.T.; Schneider, J.; França, M.D.; Santos, L.M.; Caixeta, B.P.; Machado, A.E.H.; Bahnemann, D.W. Charge carrier dynamics and photocatalytic behavior of TiO2 nanopowders submitted to hydrothermal or conventional heat treatment. RSC Adv. 2015, 5, 70536–70545. [Google Scholar] [CrossRef]
  10. Cowan, A.J.; Tang, J.; Leng, W.; Durrant, J.R.; Klug, D.R. Water splitting by nanocrystalline TiO2 in a complete photoelectrochemical cell exhibits efficiencies limited by charge recombination. J. Phys. Chem. C 2010, 114, 4208–4214. [Google Scholar] [CrossRef]
  11. Ni, M.; Leung, M.K.H.; Leung, D.Y.C.; Sumathy, K. A review and recent developments in photocatalytic water-splitting using TiO2 for hydrogen production. Renew. Sustain. Energy Rev. 2007, 11, 401–425. [Google Scholar] [CrossRef]
  12. Tezuka, Y.; Shin, S.; Ishii, T.; Ejima, T.; Suzuki, S.; Sato, S. Photoemission and Bremsstrahlung Isochromat Spectroscopy Studies of TiO2 (Rutile) and SrTiO3. J. Phys. Soc. Jpn. 1994, 63, 347–357. [Google Scholar] [CrossRef]
  13. Kavan, L.; Grätzel, M.; Gilbert, S.E.; Klemenz, C.; Scheel, H.J. Electrochemical and photoelectrochemical investigation of single-crystal anatase. J. Am. Chem. Soc. 1996, 118, 6716–6723. [Google Scholar] [CrossRef]
  14. Chainarong, S.; Sikong, L.; Pavasupree, S.; Niyomwas, S. Synthesis and characterization of nitrogen-doped TiO2 nanomaterials for photocatalytic activities under visible light. Energy Procedia 2011, 9, 418–427. [Google Scholar] [CrossRef]
  15. Fujii, H.; Inata, K.; Ohtaki, M.; Eguchi, K.; Arai, H. Synthesis of TiO2/CdS nanocomposite via TiO2 coating on CdS nanoparticles by compartmentalized hydrolysis of Ti alkoxide. J. Mater. Sci. 2001, 36, 527–532. [Google Scholar] [CrossRef]
  16. Mba, M.; D’Acunzo, M.; Salice, P.; Carofiglio, T.; Maggini, M.; Caramori, S.; Campana, A.; Aliprandi, A.; Argazzi, R.; Carli, S.; et al. Sensitization of nanocrystalline TiO2 with multibranched organic dyes and co(III)/(II) Mediators: Strategies to improve charge collection efficiency. J. Phys. Chem. C 2013, 117, 19885–19896. [Google Scholar] [CrossRef]
  17. Klein, M.; Nadolna, J.; Gołąbiewska, A.; Mazierski, P.; Klimczuk, T.; Remita, H.; Zaleska-Medynska, A. The effect of metal cluster deposition route on structure and photocatalytic activity of mono- and bimetallic nanoparticles supported on TiO2 by radiolytic method. Appl. Surf. Sci. 2016, 378, 37–48. [Google Scholar] [CrossRef]
  18. Zhang, H.; Ming, H.; Lian, S.; Huang, H.; Li, H.; Zhang, L.; Liu, Y.; Kang, Z.; Lee, S.T. Fe2O3/carbon quantum dots complex photocatalysts and their enhanced photocatalytic activity under visible light. Dalton Trans. 2011, 40, 10822–10825. [Google Scholar] [CrossRef]
  19. Liang, C.H.; Hou, M.F.; Zhou, S.G.; Li, F.B.; Liu, C.S.; Liu, T.X.; Gao, Y.X.; Wang, X.G.; Lü, J.L. The effect of erbium on the adsorption and photodegradation of orange I in aqueous Er3+-TiO2 suspension. J. Hazard. Mater. 2006, 138, 471–478. [Google Scholar] [CrossRef]
  20. Greeley, J.; Jaramillo, T.F.; Bonde, J.; Chorkendorff, I.; Nørskov, J.K. Computational high-throughput screening of electrocatalytic materials for hydrogen evolution. Nat. Mater. 2006, 5, 909–913. [Google Scholar] [CrossRef]
  21. Nørskov, J.K.; Bligaard, T.; Logadottir, A.; Kitchin, J.R.; Chen, J.G.; Pandelov, S.; Stimming, U. Trends in the Exchange Current for Hydrogen Evolution. J. Electrochem. Soc. 2005, 152, J23. [Google Scholar] [CrossRef]
  22. Choi, W.; Termin, A.; Hoffmann, M.R. The Role of Metal Ion Dopants in Quantum-Sized TiO2: Correlation between Photoreactivity and Charge Carrier Recombination Dynamics. J. Phys. Chem. 2002, 98, 13669–13679. [Google Scholar] [CrossRef]
  23. Kitano, S.; Hashimoto, K.; Kominami, H. Photocatalytic degradation of 2-propanol over metal-ion-loaded titanium(IV) oxide under visible light irradiation: Effect of physical properties of nano-crystalline titanium(IV) oxide. Appl. Catal. B 2011, 101, 206–211. [Google Scholar] [CrossRef]
  24. Xing, J.; Chen, J.F.; Li, Y.H.; Yuan, W.T.; Zhou, Y.; Zheng, L.R.; Wang, H.F.; Hu, P.; Wang, Y.; Zhao, H.J.; et al. Stable isolated metal atoms as active sites for photocatalytic hydrogen evolution. Chem.-A Eur. J. 2014, 20, 2138–2144. [Google Scholar] [CrossRef]
  25. Kaise, M.; Nagai, H.; Tokuhashi, K.; Kondo, S.; Nimura, S.; Kikuchi, O. Electron Spin Resonance Studies of Photocatalytic Interface Reactions of Suspended M/TiO2 (M = Pt, Pd, Ir, Rh, Os, or Ru) with Alcohol and Acetic Acid in Aqueous Media. Langmuir 1994, 10, 1345–1347. [Google Scholar] [CrossRef]
  26. Sathish, M.; Viswanathan, B.; Viswanath, R.P. Alternate synthetic strategy for the preparation of CdS nanoparticles and its exploitation for water splitting. Int. J. Hydrogen Energy 2006, 31, 891–898. [Google Scholar] [CrossRef]
  27. Futschek, T.; Marsman, M.; Hafner, J. Structural and magnetic isomers of small Pd and Rh clusters: An ab initio density functional study. J. Phys. Condens. Matter 2005, 17, 5927–5963. [Google Scholar] [CrossRef]
  28. Buceta, D.; Piñeiro, Y.; Vázquez-Vázquez, C.; Rivas, J.; López-Quintela, M.A. Metallic clusters: Theoretical background, properties and synthesis in microemulsions. Catalysts 2014, 4, 356–374. [Google Scholar] [CrossRef]
  29. Perron, H.; Domain, C.; Roques, J.; Drot, R.; Simoni, E.; Catalette, H. Optimisation of accurate rutile TiO2 (110), (100), (101) and (001) surface models from periodic DFT calculations. Theor. Chem. Acc. 2007, 117, 565–574. [Google Scholar] [CrossRef]
  30. Ramamoorthy, M.; Vanderbilt, D.; King-Smith, R.D. First-principles calculations of the energetics of stoichiometric TiO2 surfaces. Phys. Rev. B 1994, 49, 16721–16727. [Google Scholar] [CrossRef]
  31. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  32. López-Caballero, P.; Ramallo-López, J.M.; Giovanetti, L.J.; Buceta, D.; Miret-Artés, S.; López-Quintela, M.A.; Requejo, F.G.; de Lara-Castells, M.P. Exploring the properties of Ag5-TiO2 interfaces: Stable surface polaron formation, UV-Vis optical response, and CO2 photoactivation. J. Mater. Chem. A Mater. 2020, 8, 6842–6853. [Google Scholar] [CrossRef]
  33. Alotaibi, M.; Wu, Q.; Lambert, C. Computational Studies of Ag 5 Atomic Quantum Clusters Deposited on Anatase and Rutile TiO2 Surfaces. Appl. Surf. Sci. 2022, 613, 156054. [Google Scholar] [CrossRef]
  34. Heyd, J.; Scuseria, G.E. Influence of the exchange screening parameter on the performance of screened hybrid functionals. J. Chem. Phys. 2006, 125, 224106. [Google Scholar] [CrossRef]
  35. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef]
  36. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  37. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  38. Kresse, G.; Hafner, J. Ab initio molecular-dynamics simulation of the liquid-metalamorphous- semiconductor transition in germanium. Phys. Rev. B 1994, 49, 14251–14269. [Google Scholar] [CrossRef]
  39. Kresse, D.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  40. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  41. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  42. Seriani, N.; Pinilla, C.; Crespo, Y. Presence of gap states at Cu/TiO2 anatase surfaces: Consequences for the photocatalytic activity. J. Phys. Chem. C 2015, 119, 6696–6702. [Google Scholar] [CrossRef]
  43. Morgan, B.J.; Watson, G.W. A DFT + U description of oxygen vacancies at the TiO2 rutile (1 1 0) surface. Surf. Sci. 2007, 601, 5034–5041. [Google Scholar] [CrossRef]
  44. Morgan, B.J.; Watson, G.W. A density functional theory + u study of Oxygen vacancy formation at the (110), (100), (101), and (001) surfaces of rutile TiO2. J. Phys. Chem. C 2009, 113, 7322–7328. [Google Scholar] [CrossRef]
  45. Antony, A.; Hakanoglu, C.; Asthagiri, A.; Weaver, J.F. Dispersion-corrected density functional theory calculations of the molecular binding of n-alkanes on Pd(111) and PdO(101). J. Chem. Phys. 2012, 136, 054702. [Google Scholar] [CrossRef]
  46. Monkhorst, H.J.; Pack, J.D. Special points for Brillonin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  47. López-Caballero, P.; Miret-Artés, S.; Mitrushchenkov, A.O.; De Lara-Castells, M.P. Ag5-induced stabilization of multiple surface polarons on perfect and reduced TiO2rutile (110). J. Chem. Phys. 2020, 153, 164702. [Google Scholar] [CrossRef]
  48. López-Caballero, P.; Hauser, A.W.; De Lara-Castells, M.P. Exploring the Catalytic Properties of Unsupported and TiO2-Supported Cu5 Clusters: CO2 Decomposition to CO and CO2 Photoactivation. J. Phys. Chem. C 2019, 123, 23064–23074. [Google Scholar] [CrossRef]
  49. Malik, A.S.; Liu, T.; Rittiruam, M.; Saelee, T.; Da Silva, J.L.; Praserthdam, S.; Praserthdam, P. On a high photocatalytic activity of high-noble alloys Au–Ag/TiO2 catalysts during oxygen evolution reaction of water oxidation. Sci. Rep. 2022, 12, 2604. [Google Scholar] [CrossRef]
  50. Heffner, H.; Faccio, R.; López, I. C–doped TiO2(B): A density functional theory characterization. Appl. Surf. Sci. 2021, 551, 149479. [Google Scholar] [CrossRef]
  51. Jin, C.; Dai, Y.; Wei, W.; Ma, X.; Li, M.; Huang, B. Effects of single metal atom (Pt, Pd, Rh and Ru) adsorption on the photocatalytic properties of anatase TiO2. Appl. Surf. Sci. 2017, 426, 639–646. [Google Scholar] [CrossRef]
  52. Momma, K.; Izumi, F. VESTA: A three-dimensional visualization system for electronic and structural analysis. J. Appl. Crystallogr. 2008, 41, 653–658. [Google Scholar] [CrossRef]
  53. Wu, Q.; Hou, S.; Buceta, D.; Ordoñez, H.J.L.; López-Quintela, M.A.; Lambert, C.J. Tuning the surface states of TiO2 using Cu5 atomic clusters. Appl. Surf. Sci. 2022, 594, 153455. [Google Scholar] [CrossRef]
  54. Ren, Y.; Han, Q.; Su, Q.; Yang, J.; Zhao, Y.; Wen, H.; Jiang, Z. Effects of 4d transition metals doping on the photocatalytic activities of anatase TiO2 (101) surface. Int. J. Quantum Chem. 2021, 121, e26683. [Google Scholar] [CrossRef]
  55. Deraet, X.; Turek, J.; Alonso, M.; Tielens, F.; Weckhuysen, B.M.; Calatayud, M.; De Proft, F. Understanding the Reactivity of Supported Late Transition Metals on a Bare Anatase (101) Surface: A Periodic Conceptual DFT Investigation. ChemPhysChem 2023, 24, e202200785. [Google Scholar] [CrossRef]
  56. De Lara-Castells, M.P.; Cabrillo, C.; Micha, D.A.; Mitrushchenkov, A.O.; Vazhappilly, T. Ab initio design of light absorption through silver atomic cluster decoration of TiO2. Phys. Chem. Chem. Phys. 2018, 20, 19110–19119. [Google Scholar] [CrossRef] [PubMed]
  57. Gomes Silva, C.; Juárez, R.; Marino, T.; Molinari, R.; García, H. Influence of Excitation Wavelength (UV or Visible Light) on the Photocatalytic Activity of Titania Containing Gold Nanoparticles for the Generation of Hydrogen or Oxygen from Water. J. Am. Chem. Soc. 2011, 133, 595–602. [Google Scholar] [CrossRef]
  58. Wang, J.; Liu, K.; Zhang, B.; Qiu, Y.; Xiang, Y.; Lin, W.; Yang, B.; Li, B.; Ma, G. Doping rh into tio2as a visible-light-responsive photocatalyst: The difference between rutile and anatase. Appl. Phys. Lett. 2021, 119, 213901. [Google Scholar] [CrossRef]
  59. Pabisiak, T.; Kiejna, A. Energetics of oxygen vacancies at rutile TiO2(110) surface. Solid State Commun. 2007, 144, 324–328. [Google Scholar] [CrossRef]
  60. Oviedo, J.; Miguel, M.A.S.; Sanz, J.F. Oxygen vacancies on TiO2 (110) from first principles calculations. J. Chem. Phys. 2004, 121, 7427–7433. [Google Scholar] [CrossRef]
  61. Tyo, E.C.; Vajda, S. Catalysis by clusters with precise numbers of atoms. Nat. Nanotechnol. 2015, 10, 577–588. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Bipyramidal and (b) trapezoidal Rh5 nanoclusters. Red numbers show the amount of electron on each atom. d1d5 represent the Rh-Rh bond lengths. The relevant values are provided in Table 1. (c) Density of states of bipyramidal Rh5. (d) Density of states trapezoidal Rh5.
Figure 1. (a) Bipyramidal and (b) trapezoidal Rh5 nanoclusters. Red numbers show the amount of electron on each atom. d1d5 represent the Rh-Rh bond lengths. The relevant values are provided in Table 1. (c) Density of states of bipyramidal Rh5. (d) Density of states trapezoidal Rh5.
Nanomaterials 14 00191 g001
Figure 2. Different adsorption configurations of bipyramidal Rh5 nanoclusters at TiO2 rutile (110) surface; structures presented in (ac) showing the lateral views. While structures presented in (df) are showing the top views. The Rh, Ti, and O atoms are represented by the silver, blue, and red balls, respectively.
Figure 2. Different adsorption configurations of bipyramidal Rh5 nanoclusters at TiO2 rutile (110) surface; structures presented in (ac) showing the lateral views. While structures presented in (df) are showing the top views. The Rh, Ti, and O atoms are represented by the silver, blue, and red balls, respectively.
Nanomaterials 14 00191 g002
Figure 3. Density of states and wavefunction of bipyramidal Rh5 nanocluster loaded on perfect TiO2 rutile (110) surface. The states associated with Ti, O, Rh atoms, and Ti27 atom are depicted by the green, red, blue, and pink colours, respectively. The Fermi energy level is indicated by the black vertical line. The reference colours yellow and blue for isosurfaces symbolize the positive and negative stages of wave functions, respectively. It’s important to note that these reference colours are consistently used for all wavefunction plots in the following figures.
Figure 3. Density of states and wavefunction of bipyramidal Rh5 nanocluster loaded on perfect TiO2 rutile (110) surface. The states associated with Ti, O, Rh atoms, and Ti27 atom are depicted by the green, red, blue, and pink colours, respectively. The Fermi energy level is indicated by the black vertical line. The reference colours yellow and blue for isosurfaces symbolize the positive and negative stages of wave functions, respectively. It’s important to note that these reference colours are consistently used for all wavefunction plots in the following figures.
Nanomaterials 14 00191 g003
Figure 4. Three adsorption sites of trapezoidal Rh5 nanoclusters on TiO2 rutile (110) surface; structures presented in (a) upstanding Rh5, (b) titled Rh5, and (c) lying-down Rh5 showing the lateral views. While structures presented in (df) are showing the top views. The Rh, Ti, and O atoms are represented by the silver, blue, and red circles, respectively.
Figure 4. Three adsorption sites of trapezoidal Rh5 nanoclusters on TiO2 rutile (110) surface; structures presented in (a) upstanding Rh5, (b) titled Rh5, and (c) lying-down Rh5 showing the lateral views. While structures presented in (df) are showing the top views. The Rh, Ti, and O atoms are represented by the silver, blue, and red circles, respectively.
Nanomaterials 14 00191 g004
Figure 5. Density of states of the trapezoidal Rh5 nanocluster loaded on perfect TiO2 (110) surface. The states suited on Ti, O, Rh, Ti26, Ti27, and Ti28 atoms are represented by the green, red, blue, pink, cyan, and purple colours.
Figure 5. Density of states of the trapezoidal Rh5 nanocluster loaded on perfect TiO2 (110) surface. The states suited on Ti, O, Rh, Ti26, Ti27, and Ti28 atoms are represented by the green, red, blue, pink, cyan, and purple colours.
Nanomaterials 14 00191 g005
Figure 6. Density of states and wavefunction of the bipyramidal Rh5 nanocluster loaded on reduced TiO2 (110) surface. The green, red, blue, pink, cyan, and purple represent the states located on Ti, O, Rh, Ti27, Ti28, and Ti61 atoms.
Figure 6. Density of states and wavefunction of the bipyramidal Rh5 nanocluster loaded on reduced TiO2 (110) surface. The green, red, blue, pink, cyan, and purple represent the states located on Ti, O, Rh, Ti27, Ti28, and Ti61 atoms.
Nanomaterials 14 00191 g006
Table 1. Bond lengths of the different Rh5 nanoclusters shown in Figure 1a,b.
Table 1. Bond lengths of the different Rh5 nanoclusters shown in Figure 1a,b.
Bond Length (Å)Bipyramidal Rh5Trapezoidal Rh5
d12.482.44
d22.472.40
d32.472.44
d42.502.53
d52.51-
Table 2. DFT + U calculated adsorption energies (Eads) and Bader charge distributions on adsorbed bipyramidal Rh5 nanoclusters shown in Figure 2.
Table 2. DFT + U calculated adsorption energies (Eads) and Bader charge distributions on adsorbed bipyramidal Rh5 nanoclusters shown in Figure 2.
StructureFigure 2aFigure 2bFigure 2c
Eads (eV)−5.28−4.78−4.84
Charge on Rh5 (e)+0.60+0.85+0.80
Table 3. DFT + U calculated adsorption energies (Eads) and Bader charge distributions on adsorbed trapezoidal Rh5 nanoclusters shown in Figure 4.
Table 3. DFT + U calculated adsorption energies (Eads) and Bader charge distributions on adsorbed trapezoidal Rh5 nanoclusters shown in Figure 4.
StructureFigure 4aFigure 4bFigure 4c
Eads (eV)−3.90−3.91−6.46
Charge on Rh5 (e)+0.79+0.80+0.72
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

Alotaibi, M. Geometrical Stabilities and Electronic Structures of Rh5 Nanoclusters on Rutile TiO2 (110) for Green Hydrogen Production. Nanomaterials 2024, 14, 191. https://doi.org/10.3390/nano14020191

AMA Style

Alotaibi M. Geometrical Stabilities and Electronic Structures of Rh5 Nanoclusters on Rutile TiO2 (110) for Green Hydrogen Production. Nanomaterials. 2024; 14(2):191. https://doi.org/10.3390/nano14020191

Chicago/Turabian Style

Alotaibi, Moteb. 2024. "Geometrical Stabilities and Electronic Structures of Rh5 Nanoclusters on Rutile TiO2 (110) for Green Hydrogen Production" Nanomaterials 14, no. 2: 191. https://doi.org/10.3390/nano14020191

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