Next Article in Journal
A Green Synthesis Strategy for Cobalt Phosphide Deposited on N, P Co-Doped Graphene for Efficient Hydrogen Evolution
Previous Article in Journal
Application of KNN and ANN Metamodeling for RTM Filling Process Prediction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

DFT Calculation of Carbon-Doped TiO2 Nanocomposites

1
Department of Microsystems, University of South-Eastern Norway, 3184 Horten, Norway
2
Institute of Energy Innovation, College of Materials Science and Engineering, Taiyuan University of Technology, Taiyuan 030024, China
3
Department of Inorganic Chemical Technology and Environment Engineering, Faculty of Chemical Technology and Engineering, West Pomeranian University of Technology in Szczecin, 70-322 Szczecin, Poland
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2023, 16(18), 6117; https://doi.org/10.3390/ma16186117
Submission received: 12 August 2023 / Revised: 4 September 2023 / Accepted: 6 September 2023 / Published: 7 September 2023
(This article belongs to the Section Materials Physics)

Abstract

:
Titanium dioxide (TiO2) has been proven to be an excellent material for mitigating the continuous impact of elevated carbon dioxide concentrations. Carbon doping has emerged as a promising strategy to enhance the CO2 reduction performance of TiO2. In this study, we investigated the effects of carbon doping on TiO2 using density functional theory (DFT) calculations. Two carbon doping concentrations were considered (4% and 6%), denoted as TiO2-2C and TiO2-3C, respectively. The results showed that after carbon doping, the band gaps of TiO2-2C and TiO2-3C were reduced to 1.58 eV and 1.47 eV, respectively, which is lower than the band gap of pure TiO2 (2.13 eV). This indicates an effective improvement in the electronic structure of TiO2. Barrier energy calculations revealed that compared to pure TiO2 (0.65 eV), TiO2-2C (0.54 eV) and TiO2-3C (0.59 eV) exhibited lower energy barriers, facilitating the transition to *COOH intermediates. These findings provide valuable insights into the electronic structure changes induced by carbon doping in TiO2, which can contribute to the development of sustainable energy and environmental conservation measures to address global climate challenges.

1. Introduction

The increasing concentration of CO2 in the Earth’s atmosphere is one of the most pressing challenges facing humanity today, with profound implications for climate change and global sustainability. Earth’ temperature could be 33 °C lower without the greenhouse effect, and carbon dioxide makes up about 76% of the greenhouse gases [1,2,3,4,5]. To mitigate the impact of CO2 emissions, there is a growing interest in exploring novel materials for efficient CO2 capture, storage, and utilization (CCU). Among these materials, titanium dioxide (TiO2) has shown great promise due to its abundance, low cost, and environmental friendliness. However, the practical application of TiO2 is limited by its moderate CO2 adsorption capacity and relatively large energy band gap. In recent years, efforts have been made to improve the CO2 adsorption properties of titanium dioxide through various methods, such as surface functionalization and morphological engineering [6]. Among these, doping TiO2 with foreign elements has emerged as a promising strategy. Such a doping can cause a creation of Ti3+ sites, improving the photocatalytic performance of titania, as it was proven by Wei et al. in doping with hydrogen and nitrogen [7].
Carbon materials are known as excellent CO2 sorbents [8,9,10,11,12,13,14,15,16,17,18] because of their high surface area, microporosity, and mechanical and chemical stability. Among carbon materials, remarkable adsorption properties towards carbon dioxide have been reported for carbon spheres [19,20,21,22].
A high adsorption selectivity to CO2 over N2 was demonstrated for titania/carbon spheres composites [23], which is of crucial importance, because nitrogen is the main component of flue gases. These nanocomposites also exhibited excellent cyclic stability checked over 10 consecutive adsorption–desorption cycles.
As is well known, non-metallic doping offers significant advantages over metallic doping in terms of higher photostability, lower cost, and reduced secondary environmental pollution. The use of non-metallic elements such as B, N, C and S, either individually or in combination, to narrow the bandgap and surface state modification of pristine titanium dioxide enhances its photocatalytic performance. For instance, pioneering research by Liu et al. demonstrates that the co-doping of B and N not only increases visible light absorption but also reduces the recombination of photogenerated electrons and holes. This is achieved by forming Ti-O-B-N molecules, which enhance the efficiency of electron transfer, thereby improving photocatalytic degradation under visible light irradiation [24]. They also propose that the introduction of interstitial boron in the TiO2 lattice effectively weakens nearby Ti–O bonds, making it easier for nitrogen to substitute for oxygen, thereby increasing visible light absorption and enhancing the chemical stability of doped TiO2 [25]. Additionally, Li and colleagues have prepared B and N co-doped black TiO2 using a simple sol–gel method combined with magnesium thermal reduction, resulting in stronger light absorption and higher hydrogen production in photocatalysis [26]. However, these studies typically involve a two-step doping process using costly and toxic precursors, inevitably leading to complexity, high costs, and poor production safety.
Among the different dopants to titania, carbon stands out as a particularly attractive candidate due to its unique electronic structure and potential to introduce defect sites within the TiO2 lattice. The role of carbon in titania–carbon composites is not only to enhance carbon dioxide adsorption, but also to improve prevention of electron-hole pairs recombination, thanks to the existence of electron scavenger carbon doped into TiO2. The enhancement of photocatalytic activity can be attributed to either the band gap narrowing [27] or the formation of localized mid-gap state [28] in TiO2 band gap. Table 1 shows recent advances in photocatalytic CO2 reduction catalysts.
Janus et al. prepared C-doped titania via ethanol carbonization at different temperature [34] from 150 to 400 °C or at 180 °C under the pressure of 10 bar [35]. With increasing carbon content in TiO2 photocatalysts, the activity for phenol decomposition under UV light decreased, but that under visible light was stable. The photocatalytic activity of the commercial titania P25 modified under pressure with ethanol, tested during three azo dyes decomposition under UV light irradiation, was two times higher than for unmodified titania.
Kang et al. [36] described visible sensitive carbon-doped TiO2 photocatalyst via grinding titania with ethanol and heating treatment. The formation of Ti–C and C–O bindings in the ground product was determined by heating at 200 °C. Further heating the product at 400 °C caused a decrease in the photocatalytic activity because of the dissociation of Ti–C bindings in the product, whilst the C–O bindings were still maintained.
The composites of titania with carbon spheres used for the reduction in CO2 has been studied within the project PhotoRed [37,38].
This article employs the Density Functional Theory (DFT) to investigate the impact of carbon (C) doping on titanium dioxide (TiO2). The created TiO2 structures consist of 48 atoms (32 oxygen atoms and 16 titanium atoms). To simulate the influence of different carbon concentrations on the crystal structure, researchers have created two different carbon doping levels (4% and 6%), referred to as TiO2-2C and TiO2-3C, respectively. The models for all structures are depicted in Figure 1. Subsequently, the electronic density of states (DOS) distribution for these different structures is studied to explore the interactions between individual atoms. The analysis of the DOS sheds light on the electronic structure modifications induced by C doping. Furthermore, the free energies of each structure were evaluated, and the energy requirements at each stage of CO2 adsorption were analyzed. This enabled a deeper understanding of the role of TiO2 in the CO2 adsorption process, elucidating its impact on CO2 adsorption capacity.

2. Materials and Methods

All the computational simulations were conducted within the framework of Materials Studio, utilizing the CASTEP and DMol3 modules to explore a wide range of surface models of titanium dioxide. These simulations aimed to investigate the structural and electronic properties of TiO2 surfaces under various conditions. To avoid unwanted periodic interactions, a 20 Å vacuum layer was carefully introduced around the surface models. In the calculation of the band gap, the Perdew–Burke–Ernzerhof (PBE) approximation within the Generalized Gradient Approximation (GGA) method was employed [39]. However, in the case of transition metal oxides, the GGA approach may introduce self-interaction errors, affecting the electronic structure. To mitigate this error and alleviate the underestimation of the band gap and the delocalization of electrons, the GGA+U method was employed. Specifically, the 3D orbitals of Ti were treated with an additional Coulomb potential (UTi = 8) to account for the strong on-site electron–electron interactions [40]. This enabled a more accurate representation of the TiO2 electronic structure and its response to different surface modifications. In terms of computational parameters, a plane wave energy cutoff of 400 eV was chosen to ensure a well-converged solution for the electronic states. Additionally, norm-conserving pseudopotentials were employed to efficiently describe the interaction between valence electrons and atomic cores. The Brillouin zone integration was performed using a 1 × 2 × 1 Monkhorst-Pack grid, which provided an adequate sampling of the reciprocal space. To obtain reliable structural and energetic properties, extensive structural optimizations were conducted until the forces acting on each atom were minimized to a convergence criterion of less than 10−3 eV Å−1. This ensured that the simulated systems reached their most stable configurations and allowed for accurate comparisons between different surface models.
After structural optimization, the lattice parameters of pure rutile-type TiO2 were calculated to be a = b = 7.5520 Å and c = 18.972 Å, which closely match experimental results [41]. This demonstrates the reliability of the computational methods and results presented in this study.
For pure TiO2, the calculated average Ti–O bond length was found to be 2.055 Å. The results are as expected [42]. The doped TiO2, following geometric optimization, is summarized in Table 2. It can be observed from the table that the average C-O-Ti bond length in the doped superlattice shows significant variations. This indicates that the introduction of C elements induces pronounced lattice distortion, facilitating easier electron transitions from the valence band to the conduction band.

3. Results and Discussion

The electronic band density has multiple impacts on the CO2 reduction reaction. Firstly, the material’s electronic band structure determines its ability to effectively supply the necessary electrons to CO2 molecules adsorbed on the surface, thus influencing reduction activity. Secondly, a higher electronic band density aids in enhancing material conductivity, facilitating faster conduction of electrons from external sources to the reduction reaction sites, thereby promoting reaction rates. Additionally, different electronic band structures may lead to distinct product selectivities, and tuning the band structure allows for control over product types and proportions. Figure 2 displays the Density of States (DOS) for pure titanium dioxide and carbon-doped titanium dioxide. It is worth noting that the actual band gap of rutile titanium dioxide is 3.23 eV [43]. However, in this simulation, the band gap calculated using the Generalized Gradient Approximation (GGA) method is 2.13 eV. This is due to the lower calculated position of the valence band, resulting in a smaller band gap. Upon carbon doping, the band gaps of TiO2-2C and TiO2-3C are reduced to 1.58 eV and 1.47 eV, respectively. These findings underscore the significance of optimizing catalytic materials by adjusting electronic band density to promote CO2 reduction reactions. To gain a deeper understanding of band gap variations with carbon doping, Figure 3 illustrates the Projected Density of States (PDOS). In the case of rutile-phase titanium dioxide, the valence band is primarily composed of O-2p states, while the conduction band is dominated by Ti-3d states. Following carbon doping, due to lattice distortion, impurity bands above the valence band maximum (VBM) are mainly composed of O-2p states, while the conduction band minimum (CBM) shifts to lower energy levels, facilitating hybridization between C-2p and Ti-3d states [44]. It can be observed that as elemental doping reduces crystal symmetry, the doped energy levels of TiO2 with C merge into the band gap above VBM, reducing the energy required for charge carrier transitions and favoring the progressive excitation of electrons from the valence band to the conduction band, thereby facilitating electron transfer.
Subsequently, the adsorption process of carbon dioxide on the surface of titanium dioxide was systematically investigated, and the formation and transformation of intermediates were thoroughly studied. This research holds significant importance for understanding CO2 catalytic reduction and the performance of TiO2 as a catalyst. Free energy is an important concept used to characterize the stability of a system and the direction of a reaction, representing the energy available to the system. When the system is under constant temperature and pressure, if the Gibbs free energy of the system decreases, it means that the reaction is reversible and proceeds in a more stable direction, while an increase in Gibbs free energy means that the reaction is irreversible and proceeds in an unstable direction [42,45].
The free energy (∆G) was computed using the following formula:
G = E + Z P E T S
where E means total energy and ZPE was the zero-point energy; the entropy (∆S) of each adsorbed state were obtained from DFT calculation, whereas the thermodynamic corrections for gas molecules were from standard tables.
Figure 4, exemplifying the TiO2-2C configuration, provides a representation of the atomic adsorption process occurring at each stage of CO2 reduction. Through the simulation of intermediate structures, it becomes possible to derive the respective free energy changes associated with each of these processes. This comprehensive analysis is of paramount importance in advancing our understanding of CO2 reduction mechanisms. Initially, during the carbon dioxide adsorption process, CO2 molecules chemically adsorb on the TiO2 surface, forming *CO2 intermediate species. The activated molecules are denoted by asterisks. This adsorption process releases heat, indicating favorable energy for carbon dioxide adsorption on TiO2. This initial adsorption serves as a crucial starting step for subsequent intermediate formation. The results reveal that TiO2-2C and TiO2-3C require energy of −0.29 eV and −0.36 eV, respectively, for CO2 adsorption, both of which are lower than pure TiO2 (−0.18 eV). This suggests that TiO2 doped with carbon is more prone to CO2 adsorption, attributed to its structural bandgap. Enhanced conductivity can provide electrons to CO2, allowing it to closely bind to the crystal plane.
Subsequently, hydrogen atoms (H) interact with *CO2, forming *COOH intermediates. However, our calculations indicate that the formation of *COOH is endothermic, requiring energy absorption. Thus, *COOH formation might be the energy-limiting step of the entire reaction, influencing the rate and efficiency of the overall process. Additionally, the *COOH intermediate is transient, existing in a higher-energy state during the reaction and likely acting as a temporary reaction intermediate. This process entails overcoming a certain energy barrier and is influenced by reaction temperature, environmental conditions, and catalyst properties. Significantly, the energy barriers for doped TiO2 are lower compared to pristine TiO2, measuring 0.54 eV and 0.59 eV for TiO2-2C and TiO2-3C, respectively, compared to the original TiO2’s 0.65 eV. This implies that carbon-doped TiO2 more readily forms *COOH intermediates, consequently accelerating the reaction process. The introduction of carbon leads to changes in the lattice structure, influencing the adsorption energy and surface diffusion behavior of reactants. These changes likely result in the formation of more stable *COOH intermediates on the TiO2 surface, creating favorable conditions for subsequent catalytic steps.
In the subsequent reaction process, *COOH gradually loses its secondary hydroxyl (OH) groups. After dissociation, these OH groups are released from the *COOH intermediate, becoming free hydroxide ions. This state is transient. As *OH further reacts with the catalyst surface, the OH groups on *COOH progressively detach, releasing hydroxyls and forming CO.
In summary, our research provides valuable insights into the complex processes of carbon dioxide adsorption, intermediate formation, and conversion on the surface of TiO2. These findings are of significant importance for optimizing the design of titanium dioxide catalysts and enhancing their catalytic ability for carbon dioxide reduction. Understanding these intricate mechanisms will aid in the development and production of efficient catalysts for carbon dioxide reduction. Furthermore, this knowledge can be extended to other catalytic systems, contributing to sustainable energy conversion and environmental preservation. By harnessing the catalytic potential of titanium dioxide and its interactions with carbon dioxide, we can more effectively address global energy and climate challenges.

4. Conclusions

This study delves into the influence of C element doping on the CO2 adsorption performance of TiO2 through DFT calculations. Using TiO2-2C as an example, we conducted simulations and found that the inclusion of C significantly enhances CO2 reduction capacity of TiO2.
By analyzing the DOS distributions of each structure, we have unveiled the intricate and complex interactions between C elements and TiO2, resulting in profound changes in the electronic structure. The introduction of C elements triggers a significant distortion in the TiO2 lattice. Notably, the band gaps of TiO2-2C and TiO2-3C have been dramatically reduced from their initial values of 2.13 eV (for pure TiO2) to a mere 1.58 eV and 1.47 eV, respectively. This reduction in band gaps holds the potential to substantially enhance the efficiency of TiO2 as a catalyst for carbon dioxide conversion. Furthermore, the outcomes of our free energy analysis prominently highlight the paramount impact of carbon doping on titanium dioxide. The formation of the *COOH intermediate, known to be an endothermic process, plays a pivotal role in dictating the reaction rate. In comparison, carbon-doped titanium dioxide experiences a significant reduction in energy barriers, measuring only 0.54 eV and 0.59 eV, respectively, in stark contrast to the pristine material. This substantial decrease in energy barriers greatly promotes the formation of *COOH intermediates during the carbon dioxide conversion process. As a result, the reaction rate is accelerated, potentially leading to a significantly improved efficiency in the carbon dioxide conversion of carbon-doped titanium dioxide.
The findings of this research offer crucial guidance for designing optimized CO2 capture materials. C element doping not only improves TiO2’s adsorption capability but also opens new possibilities for future carbon capture technologies. These achievements are of vital significance in addressing global carbon emissions and promoting sustainable development. Future research could further explore the CO2 adsorption mechanisms of TiO2 under various doping conditions, expanding its application range and achieving more efficient CO2 capture materials.

Author Contributions

K.R.G., T.F. and H.H. performed the DFT calculation, data analysis and manuscript writing. U.N. provided funding support and reviewed the manuscript. G.L. provided the software. K.W. reviewed the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The research leading to these results received funding from the Norway Grants 2014–2021 via the National Centre for Research and Development under the grant number NOR/POLNORCCS/PhotoRed/0007/2019-00. H.H. acknowledged Horizon MSCA-IF-2020-Individual Fellowships (Grant No. 101024758).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Siqueira, R.M.; Freitas, G.R.; Peixoto, H.R.; Nascimento, J.F.d.; Musse, A.P.S.; Torres, A.E.B.; Azevedo, D.C.S.; Bastos-Neto, M. Carbon Dioxide Capture by Pressure Swing Adsorption. Energy Procedia 2017, 114, 2182–2192. [Google Scholar] [CrossRef]
  2. Reddy, N.R.; Reddy, P.M.; Jyothi, N.; Kumar, A.S.; Jung, J.H.; Joo, S.W. Versatile TiO2 bandgap modification with metal, non-metal, noble metal, carbon material, and semiconductor for the photoelectrochemical water splitting and photocatalytic dye degradation performance. J. Alloy Compd. 2023, 935, 167713. [Google Scholar] [CrossRef]
  3. Kokulnathan, T.; Vishnuraj, R.; Chen, S.-M.; Pullithadathil, B.; Ahmed, F.; Hasan, P.M.Z.; Bilgrami, A.L.; Kumar, S. Tailored construction of one-dimensional TiO2/Au nanofibers: Validation of an analytical assay for detection of diphenylamine in food samples. Food Chem. 2022, 380, 132052. [Google Scholar] [CrossRef]
  4. Camposeco, R.; Torres, A.E.; Zanella, R. Influence of the Preparation Method of Au, Pd, Pt, and Rh/TiO2 Nanostructures and Their Catalytic Activity on the CO Oxidation at Low Temperature. Top. Catal. 2022, 65, 798–816. [Google Scholar] [CrossRef]
  5. Nur, A.S.M.; Sultana, M.; Mondal, A.; Islam, S.; Robel, F.N.; Islam, A.; Sumi, M.S.A. A review on the development of elemental and codoped TiO2 photocatalysts for enhanced dye degradation under UV–vis irradiation. J. Water Process Eng. 2022, 47, 102728. [Google Scholar] [CrossRef]
  6. 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]
  7. Wei, S.; Wu, R.; Jian, J.; Chen, F.; Sun, Y. Black and yellow anatase titania formed by (H,N)-doping: Strong visible-light absorption and enhanced visible-light photocatalysis. Dalton Trans. 2015, 44, 1534–1538. [Google Scholar] [CrossRef] [PubMed]
  8. Sharma, A.; Jindal, J.; Mittal, A.; Kumari, K.; Maken, S.; Kumar, N. Carbon materials as CO2 adsorbents: A review. Environ. Chem. Lett. 2021, 19, 875–910. [Google Scholar] [CrossRef]
  9. Gao, X.; Yang, S.; Hu, L.; Cai, S.; Wu, L.; Kawi, S. Carbonaceous materials as adsorbents for CO2 capture: Synthesis and modification. Carbon Capture Sci. Technol. 2022, 3, 100039. [Google Scholar] [CrossRef]
  10. Ogungbenro, A.E.; Quang, D.V.; Al-Ali, K.; Abu-Zahra, M.R.M. Activated Carbon from Date Seeds for CO2 Capture Applications. Energy Procedia 2017, 114, 2313–2321. [Google Scholar] [CrossRef]
  11. Rashidi, N.A.; Yusup, S. An overview of activated carbons utilization for the post-combustion carbon dioxide capture. J. CO2 Util. 2016, 13, 1–16. [Google Scholar] [CrossRef]
  12. Mukherjee, A.; Okolie, J.A.; Abdelrasoul, A.; Niu, C.; Dalai, A.K. Review of post-combustion carbon dioxide capture technologies using activated carbon. J. Environ. Sci. 2019, 83, 46–63. [Google Scholar] [CrossRef] [PubMed]
  13. Zhao, Y.; Liu, X.; Han, Y. Microporous carbonaceous adsorbents for CO2 separation via selective adsorption. RSC Adv. 2015, 5, 30310–30330. [Google Scholar] [CrossRef]
  14. Abd, A.A.; Othman, M.R.; Kim, J. A review on application of activated carbons for carbon dioxide capture: Present performance, preparation, and surface modification for further improvement. Environ. Sci. Pollut. Res. 2021, 28, 43329–43364. [Google Scholar] [CrossRef]
  15. Oschatz, M.; Antonietti, M. A search for selectivity to enable CO2 capture with porous adsorbents. Energy Environ. Sci. 2018, 11, 57–70. [Google Scholar] [CrossRef]
  16. Mehra, P.; Paul, A. Decoding Carbon-Based Materials’ Properties for High CO2 Capture and Selectivity. ACS Omega 2022, 7, 34538–34546. [Google Scholar] [CrossRef]
  17. Liu, Y.; Wilcox, J. Effects of Surface Heterogeneity on the Adsorption of CO2 in Microporous Carbons. Environ. Sci. Technol. 2012, 46, 1940–1947. [Google Scholar] [CrossRef]
  18. Sevilla, M.; Al-Jumialy, A.S.M.; Fuertes, A.B.; Mokaya, R. Optimization of the Pore Structure of Biomass-Based Carbons in Relation to Their Use for CO2 Capture under Low- and High-Pressure Regimes. ACS Appl. Mater. Interfaces 2018, 10, 1623–1633. [Google Scholar] [CrossRef]
  19. Wickramaratne, N.P.; Jaroniec, M. Activated Carbon Spheres for CO2 Adsorption. ACS Appl. Mater. Interfaces 2013, 5, 1849–1855. [Google Scholar] [CrossRef] [PubMed]
  20. Choma, J.; Kloske, M.; Dziura, A.; Stachurska, K.; Jaroniec, M. Preparation and Studies of Adsorption Properties of Microporous Carbon Spheres. Eng. Prot. Environ. 2016, 19, 169–182. [Google Scholar] [CrossRef]
  21. Sibera, D.; Narkiewicz, U.; Kapica, J.; Serafin, J.; Michalkiewicz, B.; Wróbel, R.J.; Morawski, A.W. Preparation and characterisation of carbon spheres for carbon dioxide capture. J. Porous Mater. 2018, 26, 19–27. [Google Scholar] [CrossRef]
  22. Morawski, A.W.; Staciwa, P.; Sibera, D.; Moszyński, D.; Zgrzebnicki, M.; Narkiewicz, U. Nanocomposite Titania–Carbon Spheres as CO2 and CH4 Sorbents. ACS Omega 2020, 5, 1966–1973. [Google Scholar] [CrossRef] [PubMed]
  23. Pełech, I.; Kusiak-Nejman, E.; Staciwa, P.; Sibera, D.; Kapica-Kozar, J.; Wanag, A.; Latzke, F.; Pawłowska, K.; Michalska, A.; Narkiewicz, U.; et al. CO2 Sorbents Based on Spherical Carbon and Photoactive Metal Oxides: Insight into Adsorption Capacity, Selectivity and Regenerability. Molecules 2022, 27, 6802. [Google Scholar] [CrossRef]
  24. Liu, G.; Zhao, Y.; Sun, C.; Li, F.; Lu, G.Q.; Cheng, H.-M. Synergistic Effects of B/N Doping on the Visible-Light Photocatalytic Activity of Mesoporous TiO2. Angew. Chem. Int. Ed. 2008, 47, 4516–4520. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, G.; Yin, L.-C.; Wang, J.; Niu, P.; Zhen, C.; Xie, Y.; Cheng, H.-M. A red anatase TiO2 photocatalyst for solar energy conversion. Energy Environ. Sci. 2012, 5, 9603–9610. [Google Scholar] [CrossRef]
  26. Li, Y.; Fu, R.; Gao, M.; Wang, X. B–N co-doped black TiO2 synthesized via magnesiothermic reduction for enhanced photocatalytic hydrogen production. Int. J. Hydrog. Energy 2019, 44, 28629–28637. [Google Scholar] [CrossRef]
  27. Khan, S.U.M.; Al-Shahry, M.; Ingler, W.B., Jr. Efficient Photochemical Water Splitting by a Chemically Modified n-TiO2. Science 2002, 297, 2243–2245. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, X.; Meng, S.; Zhang, X.; Wang, H.; Zhong, W.; Du, Q. Multi-type carbon doping of TiO2 photocatalyst. Chem. Phys. Lett. 2007, 444, 292–296. [Google Scholar] [CrossRef]
  29. Hou, W.; Hung, W.H.; Pavaskar, P.; Goeppert, A.; Aykol, M.; Cronin, S.B. Photocatalytic Conversion of CO2 to Hydrocarbon Fuels via Plasmon-Enhanced Absorption and Metallic Interband Transitions. ACS Catal. 2011, 1, 929–936. [Google Scholar] [CrossRef]
  30. Chen, P.; Lei, B.; Dong, X.a.; Wang, H.; Sheng, J.; Cui, W.; Li, J.; Sun, Y.; Wang, Z.; Dong, F. Rare-Earth Single-Atom La–N Charge-Transfer Bridge on Carbon Nitride for Highly Efficient and Selective Photocatalytic CO2 Reduction. ACS Nano 2020, 14, 15841–15852. [Google Scholar] [CrossRef] [PubMed]
  31. Jiang, Z.; Sun, W.; Miao, W.; Yuan, Z.; Yang, G.; Kong, F.; Yan, T.; Chen, J.; Huang, B.; An, C.; et al. Living Atomically Dispersed Cu Ultrathin TiO2 Nanosheet CO2 Reduction Photocatalyst. Adv. Sci. 2019, 6, 1900289. [Google Scholar] [CrossRef] [PubMed]
  32. Pan, H.; Wang, X.; Xiong, Z.; Sun, M.; Murugananthan, M.; Zhang, Y. Enhanced photocatalytic CO2 reduction with defective TiO2 nanotubes modified by single-atom binary metal components. Environ. Res. 2021, 198, 111176. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, H.; Wang, Y.; Zuo, S.; Zhou, W.; Zhang, J.; Lou, X.W.D. Isolated Cobalt Centers on W18O49 Nanowires Perform as a Reaction Switch for Efficient CO2 Photoreduction. J. Am. Chem. Soc. 2021, 143, 2173–2177. [Google Scholar] [CrossRef] [PubMed]
  34. Janus, M.; Inagaki, M.; Tryba, B.; Toyoda, M.; Morawski, A.W. Carbon-modified TiO2 photocatalyst by ethanol carbonisation. Appl. Catal. B Environ. 2006, 63, 272–276. [Google Scholar] [CrossRef]
  35. Janus, M.; Morawski, A. New method of improving photocatalytic activity of commercial Degussa P25 for azo dyes decomposition. Appl. Catal. B Environ. 2007, 75, 118–123. [Google Scholar] [CrossRef]
  36. Kang, I.-C.; Zhang, Q.; Yin, S.; Sato, T.; Saito, F. Preparation of a visible sensitive carbon doped TiO2 photo-catalyst by grinding TiO2 with ethanol and heating treatment. Appl. Catal. B Environ. 2008, 80, 81–87. [Google Scholar] [CrossRef]
  37. Morawski, A.W.; Kusiak-Nejman, E.; Wanag, A.; Narkiewicz, U.; Edelmannová, M.; Reli, M.; Kočí, K. Influence of the calcination of TiO2-reduced graphite hybrid for the photocatalytic reduction of carbon dioxide. Catal. Today 2021, 380, 32–40. [Google Scholar] [CrossRef]
  38. Morawski, A.; Ćmielewska, K.; Witkowski, K.; Kusiak-Nejman, E.; Pełech, I.; Staciwa, P.; Ekiert, E.; Sibera, D.; Wanag, A.; Gano, M.; et al. CO2 Reduction to Valuable Chemicals on TiO2-Carbon Photocatalysts Deposited on Silica Cloth. Catalysts 2021, 12, 31. [Google Scholar] [CrossRef]
  39. Hammer, B. Improved adsorption energetics within density-functional theory using revised Perdew-Burke-Ernzerhof functionals. Phys. Rev. B 1999, 59, 7413. [Google Scholar] [CrossRef]
  40. German, E.; Faccio, R.; Mombrú, A.W. Comparison of standard DFT and Hubbard-DFT methods in structural and electronic properties of TiO2 polymorphs and H-titanate ultrathin sheets for DSSC application. Appl. Surf. Sci. 2018, 428, 118–123. [Google Scholar] [CrossRef]
  41. 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]
  42. Li, Z.; Xia, W.; Jia, L. Enhanced visible light photocatalytic activity of anatase TiO2 through C, N, and F codoping. Can. J. Phys. 2014, 92, 71–75. [Google Scholar] [CrossRef]
  43. Araujo-Lopez, E.; Varilla, L.A.; Seriani, N.; Montoya, J.A. TiO2 anatase's bulk and (001) surface, structural and electronic properties: A DFT study on the importance of Hubbard and van der Waals contributions. Surf. Sci. 2016, 653, 187–196. [Google Scholar] [CrossRef]
  44. Selli, D.; Fazio, G.; Seifert, G.; Di Valentin, C. Water Multilayers on TiO2 (101) Anatase Surface: Assessment of a DFTB-Based Method. J. Chem. Theory Comput. 2017, 13, 3862–3873. [Google Scholar] [CrossRef] [PubMed]
  45. Di Liberto, G.; Tosoni, S.; Pacchioni, G. Nitrogen doping in coexposed (001)–(101) anatase TiO2 surfaces: A DFT study. Phys. Chem. Chem. Phys. 2019, 21, 21497–21505. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Microstructure of TiO2 and TiO2 doped with different concentrations of C (the gray sphere represents Ti, the red sphere represents O, and the blue sphere represents C).
Figure 1. Microstructure of TiO2 and TiO2 doped with different concentrations of C (the gray sphere represents Ti, the red sphere represents O, and the blue sphere represents C).
Materials 16 06117 g001
Figure 2. Total DOS of (a) TiO2 (b) TiO2-2C and (c) TiO2-3C. The green dashed line represents the Fermi energy level.
Figure 2. Total DOS of (a) TiO2 (b) TiO2-2C and (c) TiO2-3C. The green dashed line represents the Fermi energy level.
Materials 16 06117 g002
Figure 3. PDOS of (a) TiO2 (b) TiO2-2C and (c) TiO2-3C. The green dashed line represents the Fermi energy level.
Figure 3. PDOS of (a) TiO2 (b) TiO2-2C and (c) TiO2-3C. The green dashed line represents the Fermi energy level.
Materials 16 06117 g003
Figure 4. CO2 adsorption by TiO2 (e.g., TiO2-2C) and the free energy required for each structure to undergo the reaction.
Figure 4. CO2 adsorption by TiO2 (e.g., TiO2-2C) and the free energy required for each structure to undergo the reaction.
Materials 16 06117 g004
Table 1. Recent advances in photocatalytic CO2 reduction catalysts.
Table 1. Recent advances in photocatalytic CO2 reduction catalysts.
CatalyzerMain ProductsProduct YieldSelectivenessRef
Au/TiO2CH419.7 µmol/m235[29]
Cu1/TiO2CH41416.9 ppm g−1 h−1/[30]
Cu/TiO2-2/24CO1.5 μmol g−1 h−1/[31]
Pt–Au/R-TNTsCH428.8 μmol g−1 h−1/[32]
W18O49@CoCO21,180 μmol g−1 h−189.5[33]
Table 2. Average bond lengths (Å) of the doped TiO2 after geometry optimization.
Table 2. Average bond lengths (Å) of the doped TiO2 after geometry optimization.
Bond LengthTi–CC–OTi–O
TiO2//2.055
TiO2-2C4.3923.4862.622
TiO2-3C4.0493.1622.731
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

Gustavsen, K.R.; Feng, T.; Huang, H.; Li, G.; Narkiewicz, U.; Wang, K. DFT Calculation of Carbon-Doped TiO2 Nanocomposites. Materials 2023, 16, 6117. https://doi.org/10.3390/ma16186117

AMA Style

Gustavsen KR, Feng T, Huang H, Li G, Narkiewicz U, Wang K. DFT Calculation of Carbon-Doped TiO2 Nanocomposites. Materials. 2023; 16(18):6117. https://doi.org/10.3390/ma16186117

Chicago/Turabian Style

Gustavsen, Kim Robert, Tao Feng, Hao Huang, Gang Li, Urszula Narkiewicz, and Kaiying Wang. 2023. "DFT Calculation of Carbon-Doped TiO2 Nanocomposites" Materials 16, no. 18: 6117. https://doi.org/10.3390/ma16186117

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