Next Article in Journal
Role of the Hydroxyl Groups Coordinated toTiO2 Surface on the Photocatalytic Decomposition of Ethylene at Different Ambient Conditions
Next Article in Special Issue
First-Principles Study of Stability and N2 Activation on the Octahedron RuRh Clusters
Previous Article in Journal
Metagenomic Approaches as a Tool to Unravel Promising Biocatalysts from Natural Resources: Soil and Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Density Functional Theory Study on the Influence of Cation and Anion Elements Doping on the Surface of Ti3C2 on the Adsorption Performance of Formaldehyde

1
School of Naval Architecture and Maritime, Zhejiang Ocean University, Zhoushan 316022, China
2
School of Petrochemical Engineering & Environment, Zhejiang Ocean University, Zhoushan 316022, China
3
National Engineering Research Center for Marine Aquaculture, Marine Science and Technology College, Zhejiang Ocean University, Zhoushan 316022, China
*
Author to whom correspondence should be addressed.
Submission received: 1 March 2022 / Revised: 28 March 2022 / Accepted: 29 March 2022 / Published: 30 March 2022
(This article belongs to the Special Issue DFT Study on Electrocatalysis)

Abstract

:
Based on the generalized gradient approximation of density functional theory, the geometric structure and electronic properties of the intrinsic Ti3C2 and Cu-, Pt-, Co-, Si-, F-, Cl- or Br-doped Ti3C2 are optimized, and the adsorption process of HCHO on the surface of the intrinsic Ti3C2 and doped Ti3C2 is calculated. The effects of adsorption energy, stability, DOS and doping on bond length were discussed. The results show that the adsorption energy of the intrinsic Ti3C2 crystal plane at the top site is the strongest, at −7.58 eV. The optimal adsorption sites of HCHO on various doping systems are Cu-Top, Pt-Top, Co-Top, Si-Hollow, Cl-Hollow, F-Bridge and Br-Hollow, respectively. Among the doped elements, anion (F, Cl, Br) doping at each adsorption site generally reduces the formaldehyde adsorption activity of the substrate; cationic doping (Cu, Pt, Co, Si) enhances the adsorption activity of the substrate for formaldehyde at most of the adsorption sites, indicating that the modification effect of anions on Ti3C2 is not as good as that of cations. The adsorption capacity of Si-doped Ti3C2 for formaldehyde was significantly improved. Compared with the intrinsic Ti3C2 crystal plane at the same adsorption site, the adsorption activity of HCHO was improved, and the highest adsorption energy was −8.09 eV.

1. Introduction

Energy shortage and environmental pollution are two major global challenges facing mankind today [1]. Air pollution is closely related to human health and has attracted considerable attention in the past few years. Formaldehyde has many uses in industry. Commonly used boards, paints, carpets and wallpapers in interior decoration release formaldehyde. The incomplete combustion of fuel and tobacco leaves also releases formaldehyde. In medicine, formaldehyde is often used as an antiseptic and disinfectant. The main ways humans are exposed to formaldehyde are inhalation through the respiratory tract, ingestion through the mouth and contact through the skin. Formaldehyde poisoning can cause congestion, inflammation of the conjunctiva, skin allergies, nasopharyngeal discomfort, cough, acute and chronic bronchitis and other respiratory diseases. It can also cause nausea, vomiting and gastrointestinal disorders. Therefore, removing formaldehyde is a necessary measure to reduce air pollution and protect human health.
At present, to deal with toxic and harmful gases, the adsorption method and the catalytic oxidation method are mainly used. Su Yuetan et al. used density functional theory chemical calculation methods to study the adsorption performance of formaldehyde molecules on C2N and Al-modified C2N. The results show that the modification of Al atoms changes the nearby electronic structure, thereby changing the chemical and physical behaviors of the modified Al atoms, making them act as a bridge connecting the formaldehyde molecules and the C2N layer, which enhances the adsorption capacity [2]. Hong-ping Zhang et al. studied the influence of doped Ti or N atoms on the interaction of these gases with graphene through density functional theory calculations. The analysis show that doped Ti atoms can greatly improve the interaction between gas molecules and graphene [3]. Zhijian Liu et al. found that the doping of N atoms into the graphene plane can significantly increase the adsorption strength of formaldehyde by adjusting the charge of the metal atoms. According to the analysis of geometric structure, electron transfer and density of states (DOS), the adsorption of formaldehyde belongs to a stable chemisorption, which is a combined action of electron transfer and hybridization effect [4]. Zhenzhong Zhang et al. used density functional theory to calculate the influence of Pd modification on the sensitivity of ZnO nanotubes to formaldehyde (HCHO) gas, and the results indicated that electron donation and back-donation processes between the reactants and the ZnO surface were the main reasons for the enhanced adsorption to formaldehyde [5]. Xi Zhou et al. studied the adsorption of formaldehyde (HCHO) molecules on pure nanotube and Pd-doped, Si-doped single-walled carbon nanotubes (SWCNT) by density functional theory (DFT) method; the results show that conductivities of CNTs with HCHO molecule adsorption according to their energy gaps between HOMO and LUMO in frontier molecular orbital are, in decreasing order, Pd-doped CNT, Si-doped CNT and pure CNT [6]. Navaratnarajah Kuganathan et al. studied the effect of graphene and graphene doped with B, Si and N surfaces to remove Pb atoms by using density functional theory calculations, and the results showed that the bonding of Si-doped graphene surfaces was significantly enhanced [7]. Hao Luo et al. used first-principles calculations to study the adsorption of NO2 and NH3 gas molecules on Al, Si and P-doped single-layer MoS2. The results showed that Al-, Si- and P-doped single-layer MoS2 increased the NO2 and NH3 structural stability [8]. B. Zhao et al. used first-principles calculations to explore the interaction between H2O molecules and a single-layer MoS2 surface doped with B and Si atoms, and the results showed that the introduction of impurity, in the form of B/Si atoms, can destroy the single-layer MoS2 (001) surface. The chemical insensitivity of this product promotes the capture of H2O molecules [9]. Xiaoli Jiang et al. stated in the article that, due to the synergy of different dopants, proper anion–cation double doping can greatly promote the performance of OER electrocatalysts [10]. Wenpei Kang et al. found that, for the possible kinetic mechanism, the ultra-fast pseudo capacitance contribution and higher adsorption energy caused by anion doping may promote its high-rate sodium storage capacity [11]. Yinlong Zhu et al. revealed the effect of Cl-anion doping in perovskite on the improvement of OER performance; the results demonstrate that proper Cl doping at the oxygen site of LaFeO3 (LFO) perovskite can induce multiple favorable characteristics for catalyzing the OER, including rich oxygen vacancies, increased electrical conductivity and enhanced Fe-O covalency. [12]. Zaheer Ahmed Ujjan et al. found that nonmetal doping via S and Cl significantly enhanced the photocatalytic properties of ZnO towards methylene blue (MB), due to the high density of active sites and the decreased charge recombination rate of electron–hole pairs [13]. These fruitful studies showed that doping with various elements can change the overall electronic structure of the material, thereby achieving the purpose of improving the material properties [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19].
Zhou Junhui et al. used the first-principles calculation method to study the adsorption and catalytic oxidation performance of formaldehyde molecules on the surface of single-atom catalysts (monolayer MXene-Ti3C2 modified with Ti atoms). Ti3C2 can automatically dissociate into CO molecules, and two H atoms and activated O atoms form two *OH groups [20]. Yang Jianhui et al.’s research on the surface adsorption activity of Ti3C2 is helpful in understanding their surface characteristics. First-principles calculation studies have shown that Ti2C and Ti3C2 have strong adsorption activities for O, OH and F [21]. Studies by Yujuan Zhang et al. have shown that Ti3C2 nanosheets can provide ideal recyclable materials for indoor formaldehyde removal and, at the same time, promote formaldehyde desorption at higher temperatures [22]. Qui Thanh Hoai Ta et al. used density functional theory calculations to show that Si atoms may promote the NO2 adsorption process. The results of this work may highlight the potential of Si@TiO2/Ti3C2TX heterostructures as multifunctional nanomaterials [23].
The research results of many scholars have shown that Ti3C2 MXene material is a potential adsorption material for harmful gases, and the performance of detecting, capturing and catalyzing organic vapor pollutants can be improved by doping it with various anions or cations [20,21,22,23,24,25,26,27,28,29,30,31].

2. Calculation Method and Structural Model

The calculation is done using the Vienna Ab-initio Simulation Package (VASP 5.5.4) software based on density functional theory [32]. Among them, the energy exchange correlation energy function uses the PBE (Perdew–Burke–Ernzerh) function of generalized gradient approximation (GGA) [33], and the interaction potential between valence electrons and ions uses the conjugated plane wave pseudopotential method to describe it.
The model used in the study is the Ti3C2 (001) surface, with a 3 × 3 × 1 expanded cell as the substrate for doping and adsorption [20,21]. The surface was simulated using multiple atomic layers plus a 15 Å vacuum layer. The 15 Å vacuum layer ensures that there is no interaction between adjacent plates in the z-axis direction. The lattice parameter is 3a0 × 3a0 × 20, where a0 is the lattice length obtained by energy optimization. The calculated a0 is 0.311 nm, which is in good agreement with the results of previous studies [34,35].
In the selection of the doping position, the doping of Cu atoms is used as the verification, because the Ti3C2 (001) of the 5-layer structure is symmetrical on the upper and lower sides, and only the upper 3 layers are required to be verified. On the Ti3C2 (001) surface of the 5-layer structure, one Ti atom was replaced by Cu atom in the first and third layers, and one C atom was replaced by Cu atom in the second layer. After structure optimization, it was found that the energies of doping in layers 1 to 3 were −387.96 eV, −257.50 eV, and −383.81 eV, respectively. When the first layer was doped, the energy was lower and the structure was more stable. Furthermore, based on many studies, the first layer of the Ti3C2 (001) plane is generally selected for doping.
The cutoff energy used for structural optimization and calculations was 500 eV. When the structure is optimized, the next three layers of atoms are fixed. The atomic relaxation uses a conjugate gradient algorithm to optimize atomic positions with a maximum relaxation step of 500. In terms of the iterative loop convergence criterion for electrons, EDIFF is set to 1 × 10−5 eV. The value of K-point is 3 × 3 × 1 [21]. After the structure optimization is completed, the electronic structure information of each system is calculated to discuss the adsorption stability and adsorption mechanism of each adsorption system.
The adsorption energy of HCHO molecule is calculated by the following formula:
E ads = E total E doping   atom / slab E HCHO
In the formula: E total represents the total energy of the system after adsorption; E doping   atom / slab represents the energy of the intrinsic Ti3C2 surface before adsorption or the Ti3C2 surface after doping, and E HCHO represents the energy of the free gas molecule HCHO. E ads is a negative value, indicating that energy is released during the adsorption process, and the more negative its value, the stronger the adsorption.

3. Results and Discussion

3.1. The Adsorption Characteristics of HCHO on the Surface of Ti3C2

In order to find the best adsorption site on the Ti3C2 surface, three possible preliminary adsorption methods were designed in the study, as shown in Figure 1. The first method of adsorption is to adsorb formaldehyde molecules perpendicularly to the titanium atoms in the center of the Ti3C2 surface; the second method is to place formaldehyde molecules in parallel on the gaps between the three titanium atoms; the third method of adsorption is to adsorb formaldehyde molecules on the gap between the three titanium atoms. The plane is vertically adsorbed on the bridge between two titanium atoms.
It is found from Table 1 that, when formaldehyde is adsorbed on the top site on the Ti3C2 surface, the two C–H bonds change from 1.12 Å to 1.41 Å and 1.41 Å after adsorption, which is larger than the sum of the covalent radii of C and H (1.09 Å). There is a tendency to dissociate, and its adsorption energy is −7.58 eV. When formaldehyde is adsorbed at the hollow site, the C–O bond length of formaldehyde is reduced by 0.12 Å compared to the bond length when it is not adsorbed, while the two C–H bonds increase by 0.30 Å, indicating that the C–O bond is strengthened after formaldehyde is adsorbed on the Ti3C2 surface, while the C–H bond is activated and its adsorption energy is −7.54 eV. When formaldehyde is adsorbed on the bridge site, the formaldehyde is decomposed into oxygen atoms and methylene groups (–CH2), and the oxygen atoms are adsorbed between the three titanium atoms of the substrate. The two C–H bonds increase by 0.06 Å, and the two C–H bonds do not change a lot, and the adsorption energy is −5.42 eV. By comparing the three adsorption methods, it is found that the system energy is the lowest when formaldehyde is adsorbed at the top site, and the adsorption energy produced is also the largest, which is the best adsorption method for formaldehyde on the intrinsic Ti3C2 crystal surface.
In order to further understand the mechanism of formaldehyde adsorption and dissociation on Ti3C2 surface, this paper analyzes the electronic structure of the most stable adsorption position from the charge density distribution [36]. As shown in Figure 2, the yellow area around the atom indicates that the charge density of the bond area increases, and the blue area indicates that the charge density of the bond area decreases. In addition, the differential charge analysis method is used to calculate the charge of each atom in the adsorption system. It can be seen from Figure 2 that, when formaldehyde is stably adsorbed at each adsorption site on the Ti3C2 surface, electron transfer occurs between the oxygen atom and the hydrogen atom. The oxygen atom loses electrons, while the two hydrogen atoms gain electrons, causing the two hydrogen atoms to have a tendency to be catalytically decomposed. When formaldehyde is adsorbed on the top and hollow sites, the C atom of formaldehyde and the Ti atom of the substrate generate Ti–C bonds during the electron transfer process. When formaldehyde is stably adsorbed at the bridge site, the carbon and oxygen atoms in the adsorption system lose part of their electrons, the charge density of the Ti–O bond area decreases, and the Ti–O bond is formed.

3.2. The Adsorption of Formaldehyde by Element-Doped Ti3C2

3.2.1. Doping System Structure Optimization

Replacing the central titanium atom on the Ti3C2 surface with Cu, Pt, Co, Si, F, Cl or Br, obtained after structure optimization: Cu-Ti3C2 surface, Pt-Ti3C2 surface, Co-Ti3C2 surface, Si-Ti3C2 surface, Cl-Ti3C2 surface, F-Ti3C2 surface and Br-Ti3C2 surface, as shown in Figure 3.
After the structures of Cu-Ti3C2 surface, Pt-Ti3C2 surface and Co-Ti3C2 surface are optimized, and the three neighboring carbon atoms around the top doping atom all move to the center of the doping atom. For example, Cu–C bond’s length is shortened from 2.08 Å to 2.02 Å; the surface dopant insertion energy of the three is −386.14 eV, −389.03 eV and −383.61 eV, respectively. The Pt-doped surface has the lowest energy and is the most stable, followed by the Cu-doped. The Cl-Ti3C2 surface, the F-Ti3C2 surface and the Br-Ti3C2 surface doped with non-metal anions, after the structure optimization, the top dopant atoms protrude above the surface, and the surface dopant insertion energy is −385.11 eV, −385.52 eV, and −384.91 eV; the energy of the three as well as the optimized structure is similar, and the change trend is the same, indicating that non-metal anion doping may have a consistent aspect for the modification of Ti3C2 substrate. The surface of Si-Ti3C2 doped with non-metallic cations is relatively special, and the structural change is consistent with the former cation doping. The three C atoms around Si atoms on the surface of Si-Ti3C2 move to the center of the Si atom, and the Si–C bond length is 2.07 Å is shortened to 1.94 Å, and its surface dopant insertion energy is −388.34 eV.

3.2.2. Adsorption Characteristics of Doped System to HCHO

In order to study the influence of metal atomsand non-metal atom doping on the adsorption of formaldehyde molecule and to compare it with the intrinsic Ti3C2, an adsorption site consistent with Figure 1 was adopted, and three possible adsorption methods were designed for the top site, bridge site and gap site. Ti3C2 was doped with Cu, Pt, Co, Si, F, Cl and Br atoms to optimize the structure. The bond length changes of HCHO in different adsorption systems and the adsorption energy of HCHO are shown in Table 2. The comparison of the adsorption energy of each adsorption system is shown in Figure 4.
As it can be seen from Table 2, among all metal cations (Cu, Pt, Co), copper doping had the best modification effect on Ti3C2, improving the adsorption effect of the Ti3C2 substrate on formaldehyde at the top and bridge sites, and the adsorption energy increased by 0.46 eV and 0.03 eV, respectively. In the hollow position, it decreased by 0.8802 eV. Among the nonmetallic anions (F, Cl, Br), fluorine doping had the best modification effect on Ti3C2 and improved the adsorption capacity of the Ti3C2 substrate to formaldehyde at the top and bridge sites by 0.17 eV and 2.50 eV, respectively, and decreased by 3.11 eV at the hollow site. Among all the doped elements, only silicon doping significantly increased the adsorption energy of formaldehyde at the three adsorption sites on Ti3C2 substrate and the adsorption energy at the top. Hollow and bridge sites increased by 0.17 eV, 0.55 eV and 0.03 eV, respectively. In addition, only chlorine-doped Ti3C2 improved by 1.50 eV at the bridge site, while Ti3C2 doped with other elements had poor effects at all adsorption sites. Based on the judgment of adsorption energy, it can be known that Cu, Si and F have better modification effect on Ti3C2 substrate. Further research can be carried out by doping Ti3C2 with these three elements.
When formaldehyde adsorbed stably on the top position of the Cu-doped Ti3C2 surface, the two H–C bond lengths increased by 0.01 Å and 0.02 Å, respectively, and the C–O bond lengths increased by 0.01 Å. When formaldehyde adsorbed stably on the bridge site of Cu-doped Ti3C2, formaldehyde decomposed into oxygen atoms and methylene group (–CH2). When formaldehyde adsorbed at the hollow position on Cu-doped Ti3C2, the H–C bond length was shortened by 0.12 Å and 0.14 Å, respectively, and the C–O bond length was basically unchanged. When formaldehyde was stably adsorbed on the surface of F-doped Ti3C2, it decomposed into oxygen atoms and methylene group (–CH2) at two positions (top and bridge) where the adsorption energy increased. At the hollow site, the H–C bond lengths of formaldehyde were shortened by 0.32 Å and 0.31 Å, respectively, and the C–O bond lengths increased by 0.45 Å. When formaldehyde was stably adsorbed on the top position of the Si-doped Ti3C2 surface, one H–C bond was shortened by 0.03 Å, and the other H–C bond increased by 0.02 Å, and the C–O bond length was basically unchanged. When formaldehyde was stably adsorbed at hollow sites on the Si-doped Ti3C2 surface, the H–C bond lengths increased by 0.01 Å and 0.02 Å, respectively, and the C–O bond length length was basically unchanged. When formaldehyde adsorbed stably on the bridge site of Si-doped Ti3C2, formaldehyde was catalyzed to decompose into methylene (–CH2).
In order to study the adsorption mechanism of metal atom doping and non-metal atom doping on formaldehyde, the density of states and differential charges of the metal cation Cu atom, non-metal anion F atom and non-metal cation Si atom doping system were calculated.
Differential charge density can be used to analyze the electron rearrangement of the system after adsorption [37], as shown in Figure 5. Yellow represents electron enrichment, and blue represents electron reduction. It can be seen that, in the Cu-Ti3C2-HCHO system, an oxygen atom loses part of its electrons at three adsorption sites, and a copper atom and two hydrogen atoms gain part of their electrons, and formaldehyde and the substrate produce electron transfer. Formaldehyde is mainly chemisorbed at the top and hollow sites, and decomposed at the bridge sites. In the Si-Ti3C2-HCHO system, oxygen also loses some of its electrons and the two hydrogen atoms gain some of their electrons. Similarly, electron transfer between formaldehyde and the substrate occurs between the top and bridge sites. At the bridge site, formaldehyde is also decomposed and oxygen atoms are adsorbed on the substrate. At each adsorption site of the F-Ti3C2-HCHO system, an oxygen atom loses some electrons and a fluorine atom gains some electrons. The F atom moves from the center of the Ti3C2 surface up to the surface and bonds with the Ti atom.
In order to further understand the change of Ti3C2 doped with Cu, Si and F elements, the density of states of intrinsic Ti3C2 and doped Ti3C2 were compared and analyzed. It can be seen from Figure 6 that Cu-, Si- and F-doped elements significantly changed the electronic structure of the intrinsic Ti3C2 crystal plane. Compared with the intrinsic state density of Ti3C2, the peak of the state density of Cu- and Si-doped Ti3C2-adsorbed formaldehyde shifted to the lower-energy region by 0.2 eV~0.5 eV as a whole and the peak also decreased, indicating that the introduction of Cu or Si elements made the adsorption of Ti3C2 substrate for formaldehyde tend to be more stable. The overall peak of the state density of formaldehyde adsorbed by F-doped Ti3C2 migrated to the high-energy region, about 0.2 eV~0.5 eV, and decreased as well, indicating that the stability of the Ti3C2 substrate was reduced by F. The results of state density analysis showed that, compared with intrinsic Ti3C2, F-doped Ti3C2 adsorbed formaldehyde at the top and bridge sites and its adsorption energy increased, but the overall structure was not stable enough, which was also reflected in the structure optimization and differential charge. The adsorption energy of formaldehyde on the Cu-doped Ti3C2 hollow site decreased compared with the intrinsic Ti3C2, but the structural system tended to be stable. Si-doped Ti3C2 not only enhanced the adsorption energy at the three adsorption sites but also had the largest state density shift to the lower-energy region, so its system structure was the most stable among all of the doping systems studied.
Figure 7 compares electron local density maps of formaldehyde adsorption by intrinsic Ti3C2 and Cu-, Si- or F-doped Ti3C2. Electron localization function (ELF) is often used to determine the chemical bonding properties among atoms, and its value ranges from 0 to 1; when the electrons are completely localized or there are no electrons, the ELF is 0; when the ELF is 0.5, it indicates that a charge distribution similar to an electron gas is formed among atoms, which is a typical metallic bond; when the electrons are completely localized, that is, when the ELF is 1, the typical covalent bonding among atoms is shown [38]. Figure 7 shows that a small amount of Cu, Si or F doping has little effect on the chemical bond environment of Ti3C2, and the electrons of Ti atoms are completely delocalized, which shows that the main effect of doping Cu, Si or F is to improve the adsorption and decomposition of formaldehyde, rather than direct strong adsorption of formaldehyde. Taken together, these theoretical studies suggest that Cu, Si and F additions enhance the electron delocalization of the Ti3C2 catalytic center, thereby enhancing the adsorption and decomposition of formaldehyde.

4. Conclusions

The adsorption of formaldehyde on Ti3C2 and Cu-, Pt-, Co-, Si-, F-, Cl- and Br-doped Ti3C2 was studied by density functional theory. The results show that, in the metal cations, Cu, Pt and Co, only the same Cu element at the two adsorption sites improved the adsorption capacity of the Ti3C2 substrate for formaldehyde. Among the nonmetallic anions Cl, F and Br, only F improved the adsorption capacity of the substrate to formaldehyde at the two adsorption sites. In addition, non-metallic cation Si doping enhanced the adsorption performance of formaldehyde on the substrate at three adsorption sites, and the system structure was the most stable. In general, cationic doping is better than anionic doping. These conclusions provide a theoretical reference for designing new formaldehyde gas-detection materials.

Author Contributions

Software, B.Z. and J.G.; validation, Z.Z. and M.C.; resources, B.Z. and J.G.; writing—original draft preparation, Q.G.; writing—review and editing, Q.G.; supervision, J.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the key research and development program projects in Zhejiang Province, China, in 2021. Project name: Zhejiang Natural Gas Pipeline Damage Intelligent Management and Control Key Technology and System R&D and Application Demonstration. Project number: (2021C03152).

Acknowledgments

I would like to express my sincere thanks to Yingtang Zhou for providing the experimental site and guidance for this research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Le, V.T.; Vasseghian, Y.; Doan, V.D.; Nguyen, T.T.T.; Vo, T.-T.T.; Do, H.H.; Vu, K.B.; Vu, Q.H.; Lam, T.D.; Tran, V.A. Flexible and high-sensitivity sensor based on Ti3C2–MoS2 MXene composite for the detection of toxic gases. Chemosphere 2022, 291, 133025. [Google Scholar] [CrossRef] [PubMed]
  2. Su, Y.; Li, W.; Li, G.; Ao, Z.; An, T. Al-modified C2N adsorption enhancement mechanism for for-maldehyde degradation and a density functional theory study of potential catalytic activity. Chin. J. Catal. 2019, 40, 664–673. [Google Scholar]
  3. Zhang, H.P.; Luo, X.G.; Lin, X.Y.; Lu, X.; Leng, Y.; Song, H.-T. Density functional theory calculations on the adsorption of formaldehyde and other harmful gases on pure, Ti-doped, or N-doped graphene sheets. Appl. Surf. Sci. 2013, 283, 559–565. [Google Scholar] [CrossRef]
  4. Liu, Z.; Zhang, D.; Wei, T.; Wang, L.; Li, X.; Liu, B. Adsorption characteristics of formaldehyde on nitrogen doped graphene-based single atom ad-sorbents: A DFT study. Appl. Surf. Sci. 2019, 493, 1260–1267. [Google Scholar] [CrossRef]
  5. Zhang, Z.; Sun, W.; Hamreh, S. A density functional theory study on the formaldehyde detection mechanism by Pd-decorated ZnO nanotube. J. Phys. Chem. Solids 2020, 144, 109511. [Google Scholar]
  6. Zhou, X.; Zhao, C.; Chen, C.; Chen, J.; Li, Y. DFT study on adsorption of formaldehyde on pure, Pd-doped, Si-doped single-walled carbon nanotube. Appl. Surf. Sci. 2020, 525, 146595. [Google Scholar]
  7. Kuganathan, N.; Anurakavan, S.; Abiman, P.; Iyngaran, P.; Gkanas, E.I.; Chroneos, A. Adsorption of lead on the surfaces of pristine and B, Si and N-doped gra-phene. Phys. B Condens. Matter 2021, 600, 412639. [Google Scholar] [CrossRef]
  8. Luo , H.; Cao, Y.; Zhou, J.; Feng, J.; Cao, J.; Guo, H. Adsorption of NO2, NH3 on monolayer MoS2 doped with Al, Si, and P: A first-principles study. Chem. Phys. Lett. 2016, 643, 27–33. [Google Scholar]
  9. Zhao, B.; Shang, C.; Zhou, B.; Zhang, R.Q.; Wang, J.J.; Chen, Z.Q.; Jiang, M. Adsorption and dissociation of H2O molecule on the doped monolayer MoS2 with B/Si. Appl. Surf. Sci. 2019, 481, 994–1000. [Google Scholar] [CrossRef]
  10. Jiang, X.; Yue, X.; Li, Y.; Wei, X.; Zheng, Q.; Xie, F.; Lin, D.; Qu, G. Anion-Cation-dual doped tremella-like nickel phosphides for electrocatalytic water oxidation. Chem. Eng. J. 2021, 426, 130718. [Google Scholar] [CrossRef]
  11. Kang, W.; Wang, Y.; Wan, Y.; Tuo, Y.; Wang, X.; Sun, D. Embedding anion-doped Fe7S8 in N-doped carbon matrix and shell for fast and stable sodium storage. Mater. Chem. Phys. 2021, 264, 124456. [Google Scholar] [CrossRef]
  12. Zhu, Y.; Lin, Q.; Wang, Z.; Qi, D.; Yin, Y.; Liu, Y.; Zhang, X.; Shao, Z.; Wang, H. Chlorine-anion doping induced multi-factor optimization in perovskties for boosting intrinsic oxygen evolution. J. Energy Chem. 2020, 52, 115–120. [Google Scholar] [CrossRef]
  13. Ujjan, Z.A.; Bhatti, M.A.; Shah, A.A.; Tahira, A.; Shaikh, N.M.; Kumar, S.; Mugheri, A.Q.; Medany, S.S.; Nafady, A.; Alnjiman, F.; et al. Simultaneous doping of sulfur and chloride ions into ZnO nanorods for improved photocatalytic properties towards degradation of methylene blue. Ceram. Int. 2021, 48, 5535–5545. [Google Scholar] [CrossRef]
  14. Thupsuri, S.; Tabtimsai, C.; Ruangpornvisuti, V.; Wanno, B. A study of the transition metal doped boron nitride nanosheets as promising candidates for hydrogen and formaldehyde adsorptions. Phys. E Low-Dimens. Syst. Nanostruct. 2021, 134, 114859. [Google Scholar] [CrossRef]
  15. Fang, Y.; Yang, D.D.; Xiang, C.Y.; Shi, M.; Zhao, H.; Asadi, H. A density functional study on the formaldehyde recognition by Al-doped ZnO nanosheet. J. Mol. Graph. Model. 2020, 99, 107630. [Google Scholar] [CrossRef]
  16. Serinay, N.; Fellah, M.F. Acetaldehyde adsorption and detection: A Density Functional Theory Study on Al-doped Graphene. Vacuum 2020, 175, 109279. [Google Scholar] [CrossRef]
  17. Zhou, Q.; Yuan, L.; Yang, X.; Fu, Z.; Tang, Y.; Wang, C.; Zhang, H. DFT study of formaldehyde adsorption on vacancy defected graphene doped with B, N, and S. Chem. Phys. 2014, 440, 80–86. [Google Scholar] [CrossRef]
  18. Zhang, Y.; Xiao, Y.; Li, L.; Song, K.; Wang, X.; Wang, C.; Jian, X.; Ji, C.; Qian, P. Formaldehyde oxidation on Co-doped reduced CeO2 (111): First-principles calculations. Surf. Sci. 2020, 701, 121693. [Google Scholar] [CrossRef]
  19. Liu, H.; Xia, Q.; Zhang, A.; Luo, W. First-principles study of formaldehyde adsorption on TiO2 surface. Mater. Rev. 2015, 29, 149–154. [Google Scholar]
  20. Zhou, J.; Liu, G.; Jiang, Q.; Zhao, W.; Ao, Z.; An, T. Density functional theory study on single-atom catalyst Ti/Ti3C2O2 for the catalytic oxidation of formaldehyde. Chin. J. Catal. 2020, 41, 1633–1644. [Google Scholar] [CrossRef]
  21. Yang, J.; Zhang, S.; Ji, J.; Wei, S. The adsorption activity of OH, O, F and Au on two-dimensional Ti2C and Ti3C2 surfaces. Acta Phys. Chim. Sin. 2015, 31, 369–376. [Google Scholar]
  22. Zhang, Y.; Zhou, Z.; Lan, J.; Zhang, P. Prediction of Ti3C2O2 MXene as an effective capturer of for-maldehyde. Appl. Surf. Sci. 2019, 469, 770–774. [Google Scholar] [CrossRef]
  23. Ta, Q.T.H.; Tran, N.M.; Tri, N.N.; Sreedhar, A.; Noh, J.-S. Highly surface-active Si-doped TiO2/Ti3C2TX heterostructure for gas sensing and photodegradation of toxic matters. Chem. Eng. J. 2021, 425, 13143. [Google Scholar] [CrossRef]
  24. Obodo, K.O.; Ouma, C.; Modisha, P.M.; Bessarabov, D. Density functional theory calculation of Ti3C2 MXene monolayer as catalytic support for platinum towards the dehydrogenation of methylcyclohexane. Appl. Surf. Sci. 2020, 529, 147186. [Google Scholar] [CrossRef]
  25. Long, R.; Yu, Z.; Tan, Q.; Feng, X.; Zhu, X.; Li, X.; Wang, P. Ti3C2 MXene/NH2-MIL-88B(Fe): Research on the adsorption kinetics and photocatalytic performance of an efficient integrated photocatalytic adsorbent. Appl. Surf. Sci. 2021, 570, 151244. [Google Scholar] [CrossRef]
  26. Yang, S.; Liu, Q.; Wan, X.; Jiang, J.; Ai, L. Edge-rich MoS2 nanosheets anchored on layered Ti3C2 MXene for highly efficient and rapid catalytic reduction of 4-nitrophenol and methylene blue. J. Alloy Compd. 2022, 570, 161900. [Google Scholar] [CrossRef]
  27. Yu, M.; Liang, H.; Zhan, R.; Xu, L.; Niu, J. Sm-doped g-C3N4/Ti3C2 MXene heterojunction for visible-light photocatalytic degradation of ciprofloxacin. Chin. Chem. Lett. 2020, 32, 2155–2158. [Google Scholar] [CrossRef]
  28. Liu, N.; Lu, N.; Yu, H.T.; Chen, S.; Quan, X. Efficient day-night photocatalysis performance of 2D/2D Ti3C2/Porous g-C3N4 nanolayers composite and its application in the degradation of organic pollutants. Chemosphere 2019, 246, 125760. [Google Scholar] [CrossRef]
  29. Bao, X.; Li, H.; Wang, Z.; Tong, F.; Liu, M.; Zheng, Z.; Wang, P.; Cheng, H.; Liu, Y.; Dai, Y.; et al. TiO2/Ti3C2 as an efficient photocatalyst for selective oxidation of benzyl alcohol to benzaldehyde. Appl. Catal. B Environ. 2021, 286, 119885. [Google Scholar] [CrossRef]
  30. Wang, Z.; Wang, F.; Hermawan, A.; Zhu, J.; Yin, S. Surface engineering of Ti3C2Tx MXene by oxygen plasma irradiation as room temperature ethanol sensor. Funct. Mater. Lett. 2022, 15, 2251007. [Google Scholar] [CrossRef]
  31. Wang, Z.; Wang, F.; Hermawan, A.; Zhu, J.; Yin, S. A facile method for preparation of porous nitro-gen-doped Ti3C2Tx MXene for highly responsive acetone detection at high temperature. Funct. Mater. Lett. 2021, 14, 2151043. [Google Scholar] [CrossRef]
  32. Sun, G.; Kürti, J.; Rajczy, P.; Kertesz, M.; Hafner, J.; Kresse, G. Performance of the Vienna ab initio simulation package (VASP) in chemical applications. J. Mol. Struct. 2015, 624, 37–45. [Google Scholar] [CrossRef]
  33. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1997, 78, 1396. [Google Scholar] [CrossRef] [Green Version]
  34. Shein, I.R.; Ivanovskii, A.L. Graphene-like titanium carbides and nitrides Tin+1Cn, Tin+1Nn (n = 1, 2, and 3) from de-intercalated MAX phases: First-principles probing of their structural, electronic properties and relative stability. Comput. Mater. Sci. 2012, 65, 104–114. [Google Scholar] [CrossRef]
  35. Hu, Q.; Sun, D.; Wu, Q.; Wang, H.; Wang, L.; Liu, B.; Zhou, A.; He, J. MXene: A New Family of Promising Hydrogen Storage Medium. J. Phys. Chem. A Mol. Spectrosc. Kinet. Environ. Gen. Theory 2013, 117, 14253–14260. [Google Scholar] [CrossRef] [PubMed]
  36. Henkelman, G.; Arnaldsson, A.; Jónsson, H. A fast and robust algorithm for Bader decomposition of charge density. Comput. Mater. Sci. 2006, 36, 354–360. [Google Scholar] [CrossRef]
  37. Teng, B.-T.; Zhao, Y.; Wu, F.-M.; Wen, X.-D.; Chen, Q.-P.; Huang, W.-X. A density functional theory study of CF3CH2I adsorption and reaction on Ag(111). Surf. Sci. 2012, 606, 1227–1232. [Google Scholar] [CrossRef]
  38. Savin, A.; Nesper, R.; Wengert, S.; Fässler, T.F. ELF: The electron Localization Function. Angew. Chem. Int. Ed. Engl. 1992, 36, 1808–1832. [Google Scholar] [CrossRef]
Figure 1. The adsorption configuration of HCHO at different sites on the surface of Ti3C2.
Figure 1. The adsorption configuration of HCHO at different sites on the surface of Ti3C2.
Catalysts 12 00387 g001
Figure 2. Side and top views of the differential charge density plots of the stable adsorption configuration of HCHO on the Ti3C2 surface. (Isosurface value is 0.002 e/Å3).
Figure 2. Side and top views of the differential charge density plots of the stable adsorption configuration of HCHO on the Ti3C2 surface. (Isosurface value is 0.002 e/Å3).
Catalysts 12 00387 g002
Figure 3. (a) Intrinsic Ti3C2 relaxation side view, (b) Cu-doped Ti3C2 configuration relaxation side view, (c) Pt-doped Ti3C2 configuration relaxation side view, (d) Co-doped Ti3C2 configuration relaxation side view, (e) Si-doped Ti3C2 configuration relaxation side view, (f) F-doped Ti3C2 configuration relaxation side view, (g) Cl-doped Ti3C2 configuration relaxation side view, (h) side view of the relaxation of the Br-doped Ti3C2 configuration.
Figure 3. (a) Intrinsic Ti3C2 relaxation side view, (b) Cu-doped Ti3C2 configuration relaxation side view, (c) Pt-doped Ti3C2 configuration relaxation side view, (d) Co-doped Ti3C2 configuration relaxation side view, (e) Si-doped Ti3C2 configuration relaxation side view, (f) F-doped Ti3C2 configuration relaxation side view, (g) Cl-doped Ti3C2 configuration relaxation side view, (h) side view of the relaxation of the Br-doped Ti3C2 configuration.
Catalysts 12 00387 g003
Figure 4. Comparison of the formaldehyde adsorption energy of various doping systems.
Figure 4. Comparison of the formaldehyde adsorption energy of various doping systems.
Catalysts 12 00387 g004
Figure 5. Differential charge density diagram of HCHO at different adsorption sites on Cu-Ti3C2, Si-Ti3C2 and F-Ti3C2 surfaces. (Iso Ti3C2surface value is 0.002 e/Å3).
Figure 5. Differential charge density diagram of HCHO at different adsorption sites on Cu-Ti3C2, Si-Ti3C2 and F-Ti3C2 surfaces. (Iso Ti3C2surface value is 0.002 e/Å3).
Catalysts 12 00387 g005
Figure 6. The state density of each adsorption system.
Figure 6. The state density of each adsorption system.
Catalysts 12 00387 g006
Figure 7. Electron localization function.
Figure 7. Electron localization function.
Catalysts 12 00387 g007
Table 1. System energy, adsorption energy and related parameters of the three forms of formaldehyde adsorption on the intrinsic Ti3C2 surface.
Table 1. System energy, adsorption energy and related parameters of the three forms of formaldehyde adsorption on the intrinsic Ti3C2 surface.
Adsorption MethodrC–O r C H 1 r C H 2 Etotal/eVEads/eV
Top1.091.411.41−421.32−7.58
Hollow1.101.421.41−421.29−7.55
Bridge-1.181.18−419.16−5.42
Table 2. Energy and adsorption energy of different adsorption systems.
Table 2. Energy and adsorption energy of different adsorption systems.
Adsorption SiteAdsorption of BasalEtotal/eVEsubstrate/eVrC–O r C H 1 r C H 2 Eads/eV
TopTi3C2−421.32−391.611.091.411.41−7.58
Cu-Ti3C2−416.26−386.091.101.421.43−8.04
Pt-Ti3C2−416.23−389.03-1.131.11−5.07
Co-Ti3C2−412.73−389.96-1.131.11−0.64
Si-Ti3C2−418.21−388.341.091.381.43−7.74
F-Ti3C2−415.40−385.52-1.101.13−7.75
Cl-Ti3C2−413.81−385.11-1.101.12−6.57
Br-Ti3C2−413.23−384.92-1.101.12−6.18
HollowTi3C2−421.28−391.611.101.421.41−7.54
Cu-Ti3C2−414.88−386.091.101.301.28−6.66
Pt-Ti3C2−414.03−389.04-1.111.13−2.86
Co-Ti3C2−411.83−389.95-1.111.110.25
Si-Ti3C2−418.56−388.341.101.431.43−8.09
F-Ti3C2−412.09−385.531.551.101.10−4.43
Cl-Ti3C2−414.62−385.12-1.111.12−7.37
Br-Ti3C2−413.23−384.92-1.131.10−6.18
BridgeTi3C2−419.16−391.61-1.181.18−5.42
Cu-Ti3C2−413.67−386.09-1.181.18−5.45
Pt-Ti3C2−413.84−389.031.901.101.09−2.68
Co-Ti3C2−412.33−389.95-1.131.11−0.25
Si-Ti3C2−415.92−388.34-1.181.18−5.45
F-Ti3C2−415.57−385.52-1.111.13−7.92
Cl-Ti3C2−414.16−385.11-1.101.13−6.92
Br-Ti3C2−408.43−384.92-1.091.09−1.38
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guo, Q.; Zhu, B.; Zhu, Z.; Chen, M.; Guo, J. Density Functional Theory Study on the Influence of Cation and Anion Elements Doping on the Surface of Ti3C2 on the Adsorption Performance of Formaldehyde. Catalysts 2022, 12, 387. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12040387

AMA Style

Guo Q, Zhu B, Zhu Z, Chen M, Guo J. Density Functional Theory Study on the Influence of Cation and Anion Elements Doping on the Surface of Ti3C2 on the Adsorption Performance of Formaldehyde. Catalysts. 2022; 12(4):387. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12040387

Chicago/Turabian Style

Guo, Qianyu, Baikang Zhu, Zhouhao Zhu, Mengshan Chen, and Jian Guo. 2022. "Density Functional Theory Study on the Influence of Cation and Anion Elements Doping on the Surface of Ti3C2 on the Adsorption Performance of Formaldehyde" Catalysts 12, no. 4: 387. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12040387

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