Next Article in Journal
Evaluation of the Bioactive Compounds of Apis mellifera Honey Obtained from the Açai (Euterpe oleracea) Floral Nectar
Next Article in Special Issue
A New Insight into the Molecular Mechanism of the Reaction between 2-Methoxyfuran and Ethyl (Z)-3-phenyl-2-nitroprop-2-enoate: An Molecular Electron Density Theory (MEDT) Computational Study
Previous Article in Journal
Development of a Fast and Efficient Strategy Based on Nanomagnetic Materials to Remove Polystyrene Spheres from the Aquatic Environment
Previous Article in Special Issue
Computational Exploration of the Mechanism of Action of a Sorafenib-Containing Ruthenium Complex as an Anticancer Agent for Photoactivated Chemotherapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Surface Anions Adsorbed by Rutile TiO2 (001) on Photocatalytic Nitrogen Reduction Reaction: A Density Functional Theory Calculation

1
Suzhou Institute for Advanced Research, University of Science and Technology of China, Suzhou 215123, China
2
School of Materials Science and Engineering, Henan Normal University, Xinxiang 453007, China
3
School of Chemistry and Chemical Engineering, Hubei Normal University, Huangshi 435002, China
*
Author to whom correspondence should be addressed.
This author contributed equally to the work.
Molecules 2024, 29(19), 4566; https://doi.org/10.3390/molecules29194566
Submission received: 21 August 2024 / Revised: 21 September 2024 / Accepted: 24 September 2024 / Published: 25 September 2024

Abstract

:
The adsorption of common anions found in water can have a considerable impact on the surface state and optical characteristics of titanium dioxide (TiO2), which has an important impact on the photocatalytic nitrogen reduction reaction (NRR). This work utilizes density functional theory (DFT) computations to examine the electronic and optical characteristics of the TiO2 (001) surface under various anion adsorptions in order to clarify their influence on the photocatalytic NRR of TiO2. The modifications in the structure, optical, and electronic properties of TiO2 before and after anion adsorption are investigated. In addition, the routes of Gibbs free energy for the NRR are also evaluated. The results indicate that the adsorption of anions modifies the surface characteristics of TiO2 to a certain degree, hence impacting the separating and recombining charge carriers by affecting the energy gap of TiO2. More importantly, the adsorption of anions can increase the energy barriers for the NRR, thereby exerting a detrimental effect on its photocatalytic activity. These findings provide a valuable theoretical contribution to understanding the photocatalytic reaction process of TiO2 and its potential application of NRR in the actual complex water phase.

1. Introduction

Due to constant population growth and the increasing tempo of industrialization, the problems concerned with energy and environmental deterioration have become more severe. Conventional resources of energy are scarce, and many of them are dwindling in their capacity to meet the growing demand [1]. Therefore, the challenge of constant energy consumption has led to research and development on renewable, clean, and efficient energy technologies. Photocatalysis, as one of the key approaches, has shown great promise in addressing energy limitations [2,3].
Of all the photocatalytic materials, TiO2 stands out due to its favorable electronic structure; it can generate photo-induced electron–hole pairs under sunlight [4,5]. Due to its high chemical resistance and cheap price, it has become the object of research focus in the photocatalysis area [6]. Commercially, TiO2 has been widely used in various photocatalytic reactions, includes hydrogen/oxygen production by water splitting [7,8], degradation of organic pollutants [9,10], CO2 reduction [11], nitrogen reduction reaction (NRR) for ammonia synthesis, and so on [12]. Among them, the eco-friendly photocatalytic NRR is particularly prominent; it can convert N2 into NH3 at normal temperature and pressure, and the product, NH3, is one of the key chemicals for maintaining the continuation of life on earth and meeting the needs of human society in the energy and chemical industries.
Rutile and anatase are two types of crystal phases of TiO2 recognized for their superior electrical properties and photocatalytic efficiency [13]. Rutile TiO2 is renowned for its exceptional chemical stability, particularly in demanding industrial settings characterized by high temperatures and severe pH levels [14]. Rutile TiO2 has a wider bandgap; the material has a high photocatalytic activity mainly under UV light, and hence is ideal for photocatalytic processes that require effective separation of photogenerated electron–hole pairs [15]. A natural TiO2 crystal predominantly exposes the (110), (111), and (001) facets. Among them, the (001) facet exhibits significantly higher normalized surface photo reactivity [16].
Photocatalytic processes with TiO2 as the catalyst begin with the absorption of light. TiO2 effectively produces electron–hole pairs when exposed to UV light. These pairs can then engage in redox processes [17]. The reactions mostly take place on the surface of TiO2, where the reactants are adsorbed and interact. The surface characteristics of TiO2 play a crucial role in determining its total photocatalytic effectiveness [18,19]. Surface adsorption capacity, defect states, and surface charge distribution have a direct impact on the activation of reactants and the routes of reactions [20].
TiO2 is extensively employed in aqueous photocatalytic reactions [21]. Nevertheless, real water environments consist of a multitude of anions, including fluoride (F), chloride (Cl), bromide (Br), iodide (I), sulfate (SO42−), nitrate (NO3), and carbonate (CO32−) [22]. These anions readily adsorb onto the TiO2 surface, potentially introducing surface states, altering the positions of the valence and conduction bands, and thereby affecting the bandgap. This in turn impacts the photocatalytic performance of TiO2 [23]. Therefore, it is of great practical interest to examine the adsorption effects of these anions on the surface of TiO2 and their impact on photocatalytic performance.
We conducted a comprehensive study on how the adsorption of common aqueous anions, including F, Cl, Br, I, NO3, CO32−, and SO42−, affects the electrical and optical characteristics of the TiO2 (001) surface. Using density functional theory (DFT) calculations, we analyzed changes in the bandgap, density of states (DOSs), optical absorption spectra, and Mulliken charge distribution before and after anion adsorption. Moreover, we examined the effects of anion adsorption on the Gibbs free energy of the NRR on TiO2. This proves that anion adsorption does indeed affect the photocatalytic behavior of TiO2 in a negative manner. These findings not only provide a significant theoretical framework and guidance for using TiO2 photocatalysts in multicomponent aqueous systems, but are also applicable to electrocatalytic nitrogen fixation, further expanding its potential applications in clean energy technologies [24,25].

2. Results and Discussion

2.1. Adsorption Models and Adsorption Energies

As shown in Figure 1 and Figure 2, the study focuses on the adsorption of halide anions on two specific spots of the rutile TiO2 (001) surface, namely, directly above the O atom and directly above the Ti atom. These sites are denoted as TiO2-O-X and TiO2-Ti-X (X = F, Cl, Br, I), respectively. For the adsorption of NO3, CO32−, and SO42− anions, only the adsorption site directly above the Ti atom was considered, denoted as TiO2-NO3, TiO2-CO32−, and TiO2-SO42−, respectively.
Table 1 provides a thorough examination of the changes in bond length and adsorption energy on the rutile TiO2 (001) surface when different anions are adsorbed. We labeled the four edges surrounding the Ti atom beneath the adsorbed ion as a, b, c, and d, and measured the average lengths of these edges for the eight Ti-O bonds directly below. The results show that the adsorption of all anions causes an elongation of these O-Ti bonds. Among the halide anions adsorbed at the Ti site (F, Cl, Br, I), F induces the most significant bond elongation, with the bond length reaching 1.950 Å. This is complemented by the most negative adsorption energy of −3.505 eV. As the atomic radius of the halogen increases from Cl to I, both the adsorption energy and the extent of bond elongation decrease accordingly. When comparing the adsorption of halide anions at the Ti and O sites, it is discovered that adsorption at the O site leads to a somewhat lower elongation of the O-Ti bond compared to the Ti site. Likewise, the adsorption energies at the O site regularly exhibit less negative values compared to those at the Ti site. Therefore, all the following computations were conducted with the Ti site.
Besides halide anions, the study of complex anion adsorption (NO3, CO32−, SO42−) on the surface structure of TiO2 is also examined in this work. It is found that all three anions lead to good chemisorption, with the SO42− anion having the highest effect because of its large multi-atomic anion. This includes an adsorption energy of −3.025 eV and the largest degree of bond extension by 1.950 Å. These facts show that the interaction of complex anions in the solution with the TiO2 surface may lead to impressive distortions of the crystal lattice.

2.2. Electronic Structure

The changes in the band structure are of critical importance to the photocatalytic activity of TiO2 as they determine the material’s optical response range and electron transport characteristics [26]. In general, a narrower bandgap enables the material to capture a wider variety of light wavelengths, which in turn increases its ability to catalyze chemical reactions when exposed to visible light [27]. Nonetheless, a desirable bandgap is when it is not too small since it may lead to early recombination of electron–hole pairs, thus reducing the photocatalysis effect [28]. In this study, we calculated the interfacial bandgap and band alignments of TiO2 before and after anion adsorption, as illustrated in Figure 3 and Figure 4, respectively. The pristine rutile TiO2 (001) surface exhibits a bandgap of 2.45 eV. Significant alterations in the bandgap were detected with the adsorption of anions. For halide anions adsorbed at the Ti site, the bandgaps of the TiO2-Ti-F, TiO2-Ti-Cl, and TiO2-Ti-Br systems increased to 3.39 eV, 3.23 eV, and 3.14 eV, respectively, indicating a widening of the bandgap, with fluorine having the most pronounced effect. In contrast, the TiO2-Ti-I system, following I adsorption, likely formed localized states, leading to a substantial reduction in the bandgap to 1.59 eV.
The bandgaps of the TiO2-NO3, TiO2-CO32−, and TiO2-SO42− systems were calculated to be 3.27 eV, 3.29 eV, and 2.40 eV, respectively, for the adsorption of complex anions. The substantial increase in bandgap resulting from the adsorption of NO3 and CO32− indicates a robust electrical interaction with the surface of TiO2. In contrast, the little reduction in the bandgap after the adsorption of SO42− might be ascribed to the bulky molecular structure of the sulfate ion, which could lead to electronic rearrangement on the surface.
The density of states (DOSs) analysis is a powerful tool for elucidating the distribution of electrons across various energy levels and for identifying the contributions of specific atoms and orbitals to the overall electronic structure. In Figure 5, the DOSs of the TiO2 show that the valence band (VB) is mainly composed of the oxygen 2p orbitals, while the conduction band (CB) is mainly composed of titanium 3D orbitals. The analysis of the electronic structure of TiO2 revealed that the adsorption of NO3, CO32−, and SO42− organic anions has a weak impact on the material. This suggests that the N, C, and S orbitals of these complex anions have a negligible impact on the valence and conduction bands of TiO2.
Nevertheless, a noticeable change in the electronic configuration is noticed when halogen anions are absorbed into the surface of TiO2. The outer orbitals of the halogen atoms not only contribute to the modification of the valence band but also lead to the formation of localized states, particularly in the case of I. The emergence of these localized states around the Fermi level leads to a substantial decrease in the band gap of TiO2. In contrast, the adsorption of F and Cl induces a relatively broad energy dispersion, whereas Br exhibits the most pronounced energy localization. This localization may give rise to deep-level defect states, potentially impacting the photoelectric properties of the material.
The Mulliken charge analysis data, as shown in Table 2, provide a detailed view of the electron transfer dynamics within the TiO2 (001) surface following the adsorption of various complex anions. Among the complex anions studied, the TiO2-NO3 complex exhibits the most pronounced electron transfer, indicating a strong interaction between the NO3 anion and the TiO2 surface. This is followed by TiO2-CO32− and TiO2-SO42−, both of which also show significant, albeit varying, degrees of electron transfer.
In systems involving halogen anion adsorption, the TiO2-Ti-I complex demonstrates the highest level of electron transfer, followed by TiO2-Ti-F, TiO2-Ti-Cl, and TiO2-Ti-Br. The adsorption of all these anions results in a redistribution of electrons across the TiO2 surface.
After adsorption, the electron density is mostly concentrated around the adsorbed anions, resulting in a decrease in electron density on the TiO2 (001) surface. This redistribution of electrons alters the electronic environment of the TiO2 surface, thereby modifying its surface properties and potentially influencing its catalytic behavior.

2.3. Optical Absorption

The optical absorption spectrum can be obtained through linear response theory. First, the real and imaginary parts of the dielectric function (ε(ω)) are calculated, and then the optical absorption spectrum is derived by analyzing the imaginary part of the dielectric function (Im ε(ω)).
The UV-Vis absorption spectra of TiO2 samples before and after anion adsorption were measured, as depicted in Figure 6. It is evident that anion adsorption significantly alters the optical properties of TiO2. The pristine TiO2 sample exhibits a strong absorption peak around 480 nm. However, following anion adsorption, the samples display varying degrees of peak shifts and changes in absorption intensity.
For TiO2 adsorbing NO3, CO32−, and SO42−, while the primary absorption characteristics remain similar to those of pristine TiO2, the intensity changes are more pronounced, particularly with a noticeable decrease in the absorption intensity near 480 nm. On the other hand, the adsorption of halide anions (F, Cl, Br, and I) leads to significant red shifts and blue shifts in the TiO2 absorption peaks. The absorption intensity is notably higher compared to the samples adsorbing NO3, CO32−, and SO42−. Specifically, the adsorption of F and Cl results in a marked red-shift of the TiO2 absorption peak, extending the absorption range to longer wavelengths. In comparison, the adsorption of Br and I causes smaller peak shifts but still affects the absorption characteristics.
Overall, the adsorption of anions induces substantial modifications in the optical absorption properties of TiO2, particularly affecting its visible light absorption characteristics.

2.4. NRR Gibbs Free Energy Pathway

The nitrogen reduction reaction (NRR) shows great potential in the fields of environmental protection and energy production, particularly in the realms of sustainable agriculture and green chemistry [29]. The synthesis of ammonia (NH3) is a critical process in modern industry, serving as a key intermediate in fertilizer production and, potentially, as a hydrogen carrier in the future, offering a clean and efficient energy solution. The nitrogen molecule (N2) is characterized by an exceptionally high bond dissociation energy (approximately 941 kJ/mol) and the extraordinary stability of the N≡N triple bond [30]. Therefore, a catalyst capable of effectively activating and reducing N2 must possess robust electron transfer capabilities and stable reactive sites. These sites need to effectively adsorb and activate N2 molecules, as well as support the consecutive electron transfer and hydrogenation processes necessary for the reduction of N2 to NH3. Therefore, it is necessary to have catalyst materials with outstanding surface electronic properties that can be consistently maintained to facilitate uninterrupted photocatalytic reactions. TiO2 with its efficient charge separation capabilities and chemical stability, has emerged as a focal point in photocatalytic NRR research [31]. Anions present in water include F, Cl, Br, I, NO3, CO32−, and SO42− adsorb on the surface of TiO2 and modify the electronic structure, and therefore, the activity of the catalyst in NRR. Consequently, among the various photocatalytic reactions, we have prioritized NRR for evaluating catalyst efficiency. Investigating the catalytic performance of TiO2 before and after anion adsorption will be useful for the improvement of TiO2-based NRR catalysts.
The photocatalytic NRR commonly follows the alternating hydrogenation mechanism as the predominant reaction pathway [32,33]. In order to investigate the impact of anion adsorption on the photocatalytic performance of TiO2, we examined the changes in Gibbs free energy for all materials involved in this process, as depicted in Figure 7. The initial hydrogenation step (Equation (2)) within the alternating hydrogenation mechanism, which marks the onset of the NRR process, is recognized as the rate-determining step, meaning it determines the entire reaction rate and acts as a bottleneck [34].
Our observations indicate that the Gibbs free energy change for the initial hydrogenation step on pure TiO2 remains relatively small, both before and after anion adsorption. The impact of anion adsorption on the TiO2 catalyst varies significantly depending on the type of anion. The Gibbs free energy change for the rate-determining step on pure TiO2 is 1.81 eV; for polyatomic anions such as NO3, CO32−, and SO42−, the ΔG increases to 2.35 eV, 2.4 eV, and 2.25 eV, respectively, which is notably higher than the 1.81 eV of pure TiO₂. The notable impact may be ascribed to the intricate molecular compositions of these substances, which cause substantial deformation of the crystal structure and reorganization of the electrical configuration on the surface of TiO2. In contrast, the impact of monoatomic anions is relatively moderate. For instance, the ΔG after I adsorption is 2.16 eV, the highest among monoatomic anions; Br shows a ΔG of 2.02 eV, Cl is 1.94 eV, and F also exhibits a ΔG of 2.02 eV. Despite F having a stronger electron-withdrawing capability that could potentially lead to lattice distortion and changes in electronic structure similar to polyatomic anions, the larger atomic radii of Cl, Br, and I result in greater energy barriers. As for I, although it has the largest atomic radius and the lowest electronegativity among the halide anions, its larger size likely causes more severe lattice distortion, significantly raising the energy barrier in the hydrogenation step of NRR, thereby impacting the efficiency.
In a typical photocatalytic process, the photo-induced electrons (e) can be transferred to the conduction band (CB) from the valence band (VB), and the holes ( h + ) can be formed in the VB under visible light irradiation (Equation (1)). Subsequently, the photo-induced e would react with N2 to form NH3 directly (Equation (2)), and h+ in VB will react with H2O molecules to generate O2 during the reaction (Equation (3)).
T i O 2 h ν T i O 2 ( e , h + ) #
N 2 + 6 e + 6 H + 2 N H 3 #
3 H 2 O + 6 h + 3 2 O 2 + 6 H +
N 2 + 3 H 2 O h ν 2 N H 3 + 3 2 O 2
Aside from this rate-controlling step, the remaining hydrogenation processes proceed along a spontaneous downhill pathway [35]. Overall, the adsorption of various anions on TiO2 hinders the effective activation of the N≡N bond, thereby impeding the initiation of NRR to varying degrees. The effects of polyatomic anions are generally more pronounced, while the influence of monoatomic anions, although milder, still increases the energy barrier and weakens the catalytic activity of TiO2.
The individual steps of the alternating mechanism can be expressed as the following reaction equations:
N 2 +   N N
  N N + H   N N H
  N N H + H   N H N H
  N H N H + H   N H N H 2
  N H N H 2 + H   N H 2 N H 2
  N H 2 N H 2 + H   N H 2 + N H 3
  N H 2 + H   N H 3
  N H 3   + N H 3
The corresponding changes in Gibbs free energy (ΔG) can be calculated as follows:
G 1 = G   N N G G N 2
G 2 = G   N N H G   N N G H
G 3 = G   N H N H G   N N H G H
G 4 = G   N H N H 2 G   N H N H G H
G 5 = G   N H 2 N H 2 G   N H N H 2 G H
G 6 = G   N H 2 + G ( N H 3 ) G   N H 2 N H 2 G H
G 7 = G   N H 3 G   N H 2 G H
G 8 = G + G ( N H 3 ) G   N H 3

3. Computational Methods and Models

To study the changes in the electrical and optical properties of TiO2 after the adsorption of anions, we employ the Vienna ab initio simulation package (VASP) to perform first-principles calculations based on density functional theory (DFT). The projected augmented wave (PAW) method employs the Perdew–Burke–Ernzerhof (PBE) generalized gradient approximation (GGA) to characterize the exchange–correlation interactions and core electrons [36]. To solve the problem of underestimation of the band gap that is observed if standard DFT methods are used, the HSE06 hybrid functional was used [37]. By including this, it allowed for a thorough analysis of the electrical characteristics, leading to a more precise representation of the band structure. To avoid spurious interactions arising from the periodic boundary conditions used in the simulations, a vacuum layer of 20 Å thickness was inserted [38]. A plane-wave basis set with an energy cutoff of 550 eV was implemented to ensure computational precision [39]. Moreover, the convergence criterion for energy was set at 10−5 eV and for force was set at 0.01 eV/Å so as to allow full relaxation of all geometric structures [40].
The research calculates the adsorption energies using the following formula:
E a d s = E t o t a l ( E s u b s t r a t e + E a b s o r b a t e )
The overall energy of the system with the adsorbed species is denoted as E total , the energy of the clean substrate is denoted as E substrate , and the energy of the isolated adsorbate molecule is denoted as E absorbate .
The calculation of the Gibbs free energy change (ΔG) for the stages of the NRR is performed using the following formula:
G = Δ E + Δ Z P E T Δ S
The reaction energy derived from DFT calculations is denoted as Δ E , the entropy contribution at 298.15 K is denoted as T Δ S , and the zero-point energy difference is denoted as Δ ZPE [41].
This work utilized a highly refined model of the rutile TiO2 (001) surface, which was particularly developed to examine surface interactions and reactions. The lattice constants of the pristine rutile TiO2, as reported in the literature, and the relative errors from GGA calculations are summarized in Table 3 [42], demonstrating excellent agreement. The discrepancies in the a and b lattice constants are nearly 0%, while the difference along the c-axis is only 1.26%. The model was expanded to include four distinct atomic layers, comprising a total of 96 atoms, with 32 titanium (Ti) atoms and 64 oxygen (O) atoms.

4. Conclusions

This work utilized density functional theory (DFT) to examine the influence of anion adsorption on the electrical and optical characteristics of the rutile TiO2 (001) surface. From the results of the present work, it can be concluded that typical anions inherent in water influence the photocatalytic activity of TiO2. Indeed, the presence of anions modifies the energy gap between the valence and conduction bands, the distribution of energy levels, and the ability of light absorption. These changes directly influence the ability to inhibit charge carrier separation and recombination. Additionally, complex anions such as NO3, CO32−, and SO42− significantly increase the energy barrier for the NRR process, thereby reducing the photocatalytic activity of TiO2. These findings would help build and refer to a significant theoretical framework and guide to work with TiO2 photocatalysts in multicomponent aqueous systems.

Author Contributions

X.J.: Conceptualization, Methodology, Data curation, Writing—original draft; M.G.: Validation, Formal analysis; H.L.: Writing—review and editing, Conceptualization, Methodology, Project administration, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Key Research and Development Program (No. 2022YFEO134600).

Institutional Review Board Statement

This study did not require ethical approval.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article material, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no competing financial interests.

References

  1. Bai, W.; Shi, L.; Li, Z.; Liu, D.; Liang, Y.; Han, B.; Qi, J.; Li, Y. Recent progress on the preparation and application in photocatalysis of 2D MXene-based materials. Mater. Today Energy 2024, 41, 101547. [Google Scholar] [CrossRef]
  2. Aldosari, O.F.; Hussain, I. Unlocking the potential of TiO2-based photocatalysts for green hydrogen energy through water-splitting: Recent advances, future perspectives and techno feasibility assessment. Int. J. Hydrogen Energy 2024, 59, 958–981. [Google Scholar] [CrossRef]
  3. Meng, D.; Ruan, X.; Xu, M.; Jiao, D.; Fang, G.; Qiu, Y.; Zhang, Y.; Zhang, H.; Ravi, S.K.; Cui, X. An S-scheme artificial photosynthetic system with H-TiO2/g-C3N4 heterojunction coupled with MXene boosts solar H2 evolution. J. Mater. Sci. Technol. 2025, 211, 22–29. [Google Scholar] [CrossRef]
  4. Li, Z.; Wang, S.; Wu, J.; Zhou, W. Recent progress in defective TiO2 photocatalysts for energy and environmental applications. Renew. Sustain. Energy Rev. 2022, 156, 111980. [Google Scholar] [CrossRef]
  5. Imai, K.; Fukushima, T.; Kobayashi, H.; Higashimoto, S. Visible-light responsive TiO2 for the complete photocatalytic decomposition of volatile organic compounds (VOCs) and its efficient acceleration by thermal energy. Appl. Catal. B Environ. 2024, 346, 123745. [Google Scholar] [CrossRef]
  6. Ren, Y.; Han, Y.; Li, Z.; Liu, X.; Zhu, S.; Liang, Y.; Yeung, K.W.K.; Wu, S. Ce and Er Co-doped TiO2 for rapid bacteria-killing using visible light. Bioact. Mater. 2020, 5, 201–209. [Google Scholar] [CrossRef]
  7. Wang, H.; Hu, X.; Ma, Y.; Zhu, D.; Li, T.; Wang, J. Nitrate-group-grafting-induced assembly of rutile TiO2 nanobundles for enhanced photocatalytic hydrogen evolution. Chin. J. Catal. 2020, 41, 95–102. [Google Scholar] [CrossRef]
  8. Nakada, A.; Nishioka, S.; Vequizo, J.J.M.; Muraoka, K.; Kanazawa, T.; Yamakata, A.; Nozawa, S.; Kumagai, H.; Adachi, S.; Ishitani, O.; et al. Solar-driven Z-scheme water splitting using tantalum/nitrogen co-doped rutile titania nanorod as an oxygen evolution photocatalyst. J. Mater. Chem. A 2017, 5, 11710–11719. [Google Scholar] [CrossRef]
  9. Marien, C.B.D.; Cottineau, T.; Robert, D.; Drogui, P. TiO2 Nanotube arrays: Influence of tube length on the photocatalytic degradation of Paraquat. Appl. Catal. B Environ. 2016, 194, 1–6. [Google Scholar] [CrossRef]
  10. Wang, J.; Ma, T.; Zhang, Z.; Zhang, X.; Jiang, Y.; Pan, Z.; Wen, F.; Kang, P.; Zhang, P. Investigation on the sonocatalytic degradation of methyl orange in the presence of nanometer anatase and rutile TiO2 powders and comparison of their sonocatalytic activities. Desalination 2006, 195, 294–305. [Google Scholar] [CrossRef]
  11. Xiong, Z.; Lei, Z.; Li, Y.; Dong, L.; Zhao, Y.; Zhang, J. A review on modification of facet-engineered TiO2 for photocatalytic CO2 reduction. J. Photochem. Photobiol. C Photochem. Rev. 2018, 36, 24–47. [Google Scholar] [CrossRef]
  12. Han, Z.; Choi, C.; Hong, S.; Wu, T.-S.; Soo, Y.-L.; Jung, Y.; Qiu, J.; Sun, Z. Activated TiO2 with tuned vacancy for efficient electrochemical nitrogen reduction. Appl. Catal. B Environ. 2019, 257, 117896. [Google Scholar] [CrossRef]
  13. Padayachee, D.; Mahomed, A.S.; Singh, S.; Friedrich, H.B. Effect of the TiO2 Anatase/Rutile Ratio and Interface for the Oxidative Activation of n-Octane. ACS Catal. 2020, 10, 2211–2220. [Google Scholar] [CrossRef]
  14. Mahmood Katun, M.; Kadzutu-Sithole, R.; Moloto, N.; Nyamupangedengu, C.; Gomes, C. Improving Thermal Stability and Hydrophobicity of Rutile-TiO2 Nanoparticles for Oil-Impregnated Paper Application. Energies 2021, 14, 7964. [Google Scholar] [CrossRef]
  15. Xu, H.; Shi, J.-L.; Lyu, S.; Lang, X. Visible-light photocatalytic selective aerobic oxidation of thiols to disulfides on anatase TiO2. Chin. J. Catal. 2020, 41, 1468–1473. [Google Scholar] [CrossRef]
  16. Lai, Z.; Peng, F.; Wang, H.; Yu, H.; Zhang, S.; Zhao, H. A new insight into regulating high energy facets of rutile TiO2. J. Mater. Chem. A 2013, 1, 4182–4185. [Google Scholar] [CrossRef]
  17. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 Photocatalysis: Concepts, Mechanisms, and Challenges. Adv. Mater. 2019, 31, e1901997. [Google Scholar] [CrossRef]
  18. Guo, Q.; Ma, Z.; Zhou, C.; Ren, Z.; Yang, X. Single Molecule Photocatalysis on TiO2 Surfaces. Chem. Rev. 2019, 119, 11020–11041. [Google Scholar] [CrossRef]
  19. Low, J.; Cheng, B.; Yu, J. Surface modification and enhanced photocatalytic CO2 reduction performance of TiO2: A review. Appl. Surf. Sci. 2017, 392, 658–686. [Google Scholar] [CrossRef]
  20. Cai, Q.; Wang, F.; He, J.; Dan, M.; Cao, Y.; Yu, S.; Zhou, Y. Oxygen defect boosted photocatalytic hydrogen evolution from hydrogen sulfide over active {0 0 1} facet in anatase TiO2. Appl. Surf. Sci. 2020, 517, 146198. [Google Scholar] [CrossRef]
  21. Pekakis, P.A.; Xekoukoulotakis, N.P.; Mantzavinos, D. Treatment of textile dyehouse wastewater by TiO2 photocatalysis. Water Res. 2006, 40, 1276–1286. [Google Scholar] [CrossRef]
  22. Saleh, R.; Taufik, A.; Prakoso, S.P. Fabrication of Ag2O/TiO2 composites on nanographene platelets for the removal of organic pollutants: Influence of oxidants and inorganic anions. Appl. Surf. Sci. 2019, 480, 697–708. [Google Scholar] [CrossRef]
  23. Habiba, U.; Islam, M.S.; Siddique, T.A.; Afifi, A.M.; Ang, B.C. Adsorption and photocatalytic degradation of anionic dyes on Chitosan/PVA/Na-Titanate/TiO2 composites synthesized by solution casting method. Carbohydr. Polym. 2016, 149, 317–331. [Google Scholar] [CrossRef] [PubMed]
  24. Guan, D.; Xu, H.; Zhang, Q.; Huang, Y.-C.; Shi, C.; Chang, Y.-C.; Xu, X.; Tang, J.; Gu, Y.; Pao, C.-W. Identifying a Universal Activity Descriptor and a Unifying Mechanism Concept on Perovskite Oxides for Green Hydrogen Production. Adv. Mater. 2023, 35, 2305074. [Google Scholar] [CrossRef]
  25. Tan, Y.; Shao, M.; Legesse, M.; Mellouhi, F.E.; Bentria, E.T.; Madjet, M.E.; Fisher, T.S.; Kais, S.; Alharbi, F.H. A Universal Chemical-Induced Tensile Strain Tuning Strategy to Boost Oxygen-Evolving Electrocatalysis on Perovskite Oxides. Appl. Phys. Rev. 2022, 9, 011422. [Google Scholar]
  26. Wang, Y.; Huang, L.; Zhang, T.C.; Wang, Y.; Yuan, S. Visible-Light-Induced photocatalytic oxidation of gaseous ammonia on Mo, c-codoped TiO2: Synthesis, performance and mechanism. Chem. Eng. J. 2024, 482, 148811. [Google Scholar] [CrossRef]
  27. Ma, X.; Wu, X.; Wang, H.; Wang, Y. A Janus MoSSe monolayer: A potential wide solar-spectrum water-splitting photocatalyst with a low carrier recombination rate. J. Mater. Chem. A 2018, 6, 2295–2301. [Google Scholar] [CrossRef]
  28. Huang, L.; Cui, H.; Zhang, W.; Pu, D.; Zeng, G.; Liu, Y.; Zhou, S.; Wang, C.; Zhou, J.; Wang, C.; et al. Efficient Narrow-Bandgap Mixed Tin-Lead Perovskite Solar Cells via Natural Tin Oxide Doping. Adv. Mater. 2023, 35, e2301125. [Google Scholar] [CrossRef]
  29. Zhao, X.; Hu, G.; Chen, G.F.; Zhang, H.; Zhang, S.; Wang, H. Comprehensive Understanding of the Thriving Ambient Electrochemical Nitrogen Reduction Reaction. Adv. Mater. 2021, 33, e2007650. [Google Scholar] [CrossRef]
  30. Trangwachirachai, K.; Kao, I.T.; Huang, W.-H.; Chen, C.-L.; Lin, Y.-C. Co-activation of methane and nitrogen to acetonitrile over MoCx/Al2O3 catalysts. Catal. Sci. Technol. 2023, 13, 5248–5258. [Google Scholar] [CrossRef]
  31. Zhao, R.; Wang, G.; Mao, Y.; Bao, X.; Wang, Z.; Wang, P.; Liu, Y.; Zheng, Z.; Dai, Y.; Cheng, H.; et al. Li-intercalation boosted oxygen vacancies enable efficient electrochemical nitrogen reduction on ultrathin TiO2 nanosheets. Chem. Eng. J. 2022, 430, 133085. [Google Scholar] [CrossRef]
  32. Zhang, L.; Ding, L.X.; Chen, G.F.; Yang, X.; Wang, H. Ammonia Synthesis Under Ambient Conditions: Selective Electroreduction of Dinitrogen to Ammonia on Black Phosphorus Nanosheets. Angew. Chem. Int. Ed. 2019, 58, 2612–2616. [Google Scholar] [CrossRef] [PubMed]
  33. Li, H.; Gu, S.; Sun, Z.; Guo, F.; Xie, Y.; Tao, B.; He, X.; Zhang, W.; Chang, H. The in-built bionic “MoFe cofactor” in Fe-doped two-dimensional MoTe2 nanosheets for boosting the photocatalytic nitrogen reduction performance. J. Mater. Chem. A 2020, 8, 13038–13048. [Google Scholar] [CrossRef]
  34. Vedhanarayanan, B.; Chiu, C.-C.; Regner, J.; Sofer, Z.; Seetha Lakshmi, K.C.; Lin, J.-Y.; Lin, T.-W. Highly exfoliated NiPS3 nanosheets as efficient electrocatalyst for high yield ammonia production. Chem. Eng. J. 2022, 430, 132649. [Google Scholar] [CrossRef]
  35. Li, H.; Zhao, H.; Li, C.; Li, B.; Tao, B.; Gu, S.; Wang, G.; Chang, H. Redox regulation of photocatalytic nitrogen reduction reaction by gadolinium doping in two-dimensional bismuth molybdate nanosheets. Appl. Surf. Sci. 2022, 600, 154105. [Google Scholar] [CrossRef]
  36. Yalcin, O.; Wachs, I.E.; Onal, I. Role of chromium in Cr–Fe oxide catalysts for high temperature water-gas shift reaction—A DFT study. Int. J. Hydrogen Energy 2021, 46, 17154–17162. [Google Scholar] [CrossRef]
  37. Wang, Z.; Wang, Q.; Han, Y.; Ma, Y.; Zhao, H.; Nowak, A.; Li, J. Deep learning for ultra-fast and high precision screening of energy materials. Energy Storage Mater. 2021, 39, 45–53. [Google Scholar] [CrossRef]
  38. Zhao, P.; Zheng, J.; Guo, P.; Jiang, Z.; Cao, L.; Wan, Y. Electronic and magnetic properties of Re-doped single-layer MoS2: A DFT study. Comput. Mater. Sci. 2017, 128, 287–293. [Google Scholar] [CrossRef]
  39. Pašti, I.A.; Jovanović, A.; Dobrota, A.S.; Mentus, S.V.; Johansson, B.; Skorodumova, N.V. Atomic adsorption on pristine graphene along the Periodic Table of Elements—From PBE to non-local functionals. Appl. Surf. Sci. 2018, 436, 433–440. [Google Scholar] [CrossRef]
  40. Gu, X.; Luo, Y.; Li, Q.; Wang, R.; Fu, S.; Lv, X.; He, Q.; Zhang, Y.; Yan, Q.; Xu, X.; et al. First-Principle Insight Into the Effects of Oxygen Vacancies on the Electronic, Photocatalytic, and Optical Properties of Monoclinic BiVO4(001). Front. Chem. 2020, 8, 601983. [Google Scholar] [CrossRef]
  41. Guo, H.; Li, L.; Wang, X.; Yao, G.; Yu, H.; Tian, Z.; Li, B.; Chen, L. Theoretical Investigation on the Single Transition-Metal Atom-Decorated Defective MoS2 for Electrocatalytic Ammonia Synthesis. ACS Appl. Mater. Interfaces 2019, 11, 36506–36514. [Google Scholar] [CrossRef] [PubMed]
  42. Buchanan, R.C.; Park, T. Materials Crystal Chemistry; Marcel Dekker: New York, NY, USA, 1997. [Google Scholar]
Figure 1. Model of the rutile TiO2 (001) surface. The four Ti-O bonds connected by Ti atoms below the adsorbed ion are a, b, c and d; Gray ball is Ti atom, red balls are O atom.
Figure 1. Model of the rutile TiO2 (001) surface. The four Ti-O bonds connected by Ti atoms below the adsorbed ion are a, b, c and d; Gray ball is Ti atom, red balls are O atom.
Molecules 29 04566 g001
Figure 2. The models before and after anion adsorption: (al) represent TiO2-Ti-F, TiO2-Ti-Cl, TiO2-Ti-Br, TiO2-Ti-I, TiO2-O-F, TiO2-O-Cl, TiO2-O-Br, TiO2-O-I, TiO2-NO3, TiO2-CO32−, TiO2-SO42−, and TiO2, respectively.
Figure 2. The models before and after anion adsorption: (al) represent TiO2-Ti-F, TiO2-Ti-Cl, TiO2-Ti-Br, TiO2-Ti-I, TiO2-O-F, TiO2-O-Cl, TiO2-O-Br, TiO2-O-I, TiO2-NO3, TiO2-CO32−, TiO2-SO42−, and TiO2, respectively.
Molecules 29 04566 g002
Figure 3. The band structures of TiO2, TiO2-NO3, TiO2-CO32−, TiO2-SO42−, TiO2-Ti-F, TiO2-Ti-Cl, TiO2-Ti-Br, and TiO2-Ti-I. The red dotted line represents high symmetry points.
Figure 3. The band structures of TiO2, TiO2-NO3, TiO2-CO32−, TiO2-SO42−, TiO2-Ti-F, TiO2-Ti-Cl, TiO2-Ti-Br, and TiO2-Ti-I. The red dotted line represents high symmetry points.
Molecules 29 04566 g003
Figure 4. The band alignments before and after anion adsorption.
Figure 4. The band alignments before and after anion adsorption.
Molecules 29 04566 g004
Figure 5. (ah) Density of states of TiO2, TiO2-NO3, TiO2-CO32−, TiO2-SO42−, TiO2-Ti-F, TiO2-Ti-Cl, TiO2-Ti-Br, and TiO2-Ti-I.
Figure 5. (ah) Density of states of TiO2, TiO2-NO3, TiO2-CO32−, TiO2-SO42−, TiO2-Ti-F, TiO2-Ti-Cl, TiO2-Ti-Br, and TiO2-Ti-I.
Molecules 29 04566 g005
Figure 6. The optical absorption spectra before and after anion adsorption.
Figure 6. The optical absorption spectra before and after anion adsorption.
Molecules 29 04566 g006
Figure 7. The Gibbs free energy for NRR on TiO2 before and after anion adsorption. “*” represents free electrons.
Figure 7. The Gibbs free energy for NRR on TiO2 before and after anion adsorption. “*” represents free electrons.
Molecules 29 04566 g007
Table 1. Comparison of Ti-O bond length and Eads before and after anion adsorption for rutile TiO2 (001).
Table 1. Comparison of Ti-O bond length and Eads before and after anion adsorption for rutile TiO2 (001).
ConfigurationsAverage Ti-O Bond Length (Å)Eads (eV)
TiO2 (001)1.936/
TiO2-Ti-F1.950−3.505
TiO2-Ti-Cl1.948−0.913
TiO2-Ti-Br1.946−1.075
TiO2-Ti-I1.944−1.504
TiO2-O-F1.942−1.545
TiO2-O-Cl1.940−0.812
TiO2-O-Br1.938−0.631
TiO2-O-I1.937−0.603
TiO2-NO31.945−0.896
TiO2-CO32−1.947−1.278
TiO2-SO42−1.950−3.025
Table 2. The Mulliken charges before and after adsorption.
Table 2. The Mulliken charges before and after adsorption.
ConfigurationsMulliken Charge (e)
TiO2 (001)2.49
TiO2-Ti-F2.28
TiO2-Ti-Cl2.32
TiO2-Ti-Br2.36
TiO2-Ti-I2.24
TiO2-NO32.30
TiO2-CO32−2.34
TiO2-SO42−2.37
Table 3. Comparison of the theoretical model and experimental measurements of the lattice constants for rutile TiO2.
Table 3. Comparison of the theoretical model and experimental measurements of the lattice constants for rutile TiO2.
Lattice Constants (Å)
Categorya/bc
This work3.789.62
Experimental3.789.50
Difference0%1.26%
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

Jiang, X.; Gao, M.; Li, H. Effect of Surface Anions Adsorbed by Rutile TiO2 (001) on Photocatalytic Nitrogen Reduction Reaction: A Density Functional Theory Calculation. Molecules 2024, 29, 4566. https://doi.org/10.3390/molecules29194566

AMA Style

Jiang X, Gao M, Li H. Effect of Surface Anions Adsorbed by Rutile TiO2 (001) on Photocatalytic Nitrogen Reduction Reaction: A Density Functional Theory Calculation. Molecules. 2024; 29(19):4566. https://doi.org/10.3390/molecules29194566

Chicago/Turabian Style

Jiang, Xiaoyu, Mengyuan Gao, and Hongda Li. 2024. "Effect of Surface Anions Adsorbed by Rutile TiO2 (001) on Photocatalytic Nitrogen Reduction Reaction: A Density Functional Theory Calculation" Molecules 29, no. 19: 4566. https://doi.org/10.3390/molecules29194566

APA Style

Jiang, X., Gao, M., & Li, H. (2024). Effect of Surface Anions Adsorbed by Rutile TiO2 (001) on Photocatalytic Nitrogen Reduction Reaction: A Density Functional Theory Calculation. Molecules, 29(19), 4566. https://doi.org/10.3390/molecules29194566

Article Metrics

Back to TopTop