Next Article in Journal
Characterization of the Isothermal and Thermomechanical Fatigue Behavior of a Duplex Steel Considering the Alloy Microstructure
Next Article in Special Issue
Influence of Rare Earth Samarium/Ytterbium Salt on Electrochemical Corrosion Behavior of Aluminum-Based Anode for Batteries
Previous Article in Journal
A New Extruded Mg-6Bi-3Al-1Zn Alloy with Excellent Tensile Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles Study on the Adsorption Characteristics of Corrosive Species on Passive Film TiO2 in a NaCl Solution Containing H2S and CO2

1
School of Materials Science and Engineering, Xi’an Shiyou University, Xi’an 710065, China
2
Middle East E&P, Research Institute of Petroleum Exploration & Development, Beijing 100083, China
3
School of Chemical Engineering, Northwest University, Xi’an 710069, China
4
State Key Laboratory for Performance and Structure Safety of Petroleum Tubular Goods and Equipment Materials, CNPC Tubular Goods Research Institute, Xi’an 710077, China
5
College of Chemistry & Chemical Engineering, Xi’an Shiyou University, Xi’an 710065, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(7), 1160; https://doi.org/10.3390/met12071160
Submission received: 15 June 2022 / Revised: 1 July 2022 / Accepted: 5 July 2022 / Published: 7 July 2022
(This article belongs to the Special Issue Corrosion and Protection Technology of Metal Matrix Composites)

Abstract

:
The adsorption characteristics of corrosive anions (Cl, HS, S2−, HCO3 and CO32−) on TiO2 of TC4 titanium alloy in a NaCl solution containing H2S and CO2 were studied by density functional theory (DFT). The stable adsorption configuration of each corrosive species on the TiO2 (110) surface was obtained by geometric optimization, and the electronic structure and interface binding energy were calculated and analyzed. The results showed that the optimal adsorption positions of Cl, HS, S2−, HCO3 and CO32− on TiO2 (110) were all bridge positions. There was a strong charge interaction between the negatively charged Cl, S and O atoms in Cl, HS, S2−, HCO3 and CO32− and the positively charged Ti atoms of TiO2. The interface bonding was mainly caused by charge movement from around Ti atoms to around Cl, O, S atoms. The energy levels were mainly caused by the electron orbital hybridization of Cl-3p5, S-3p4, O-2p4 and Ti-3d2. All adsorption configurations were chemical adsorption. The order of influence of the five ions on the stability of TiO2 was S2− > CO32− > Cl > HS > HCO3. Finally, a novel corrosion mechanism was proposed to illustrate the dynamic evolution processes of pits.

1. Introduction

In the process of oil and gas exploitation and transportation, corrosive media such as H2S, CO2 and Cl are often encountered, oil country tubular goods (OCTGs) will suffer from corrosion, which is sometimes serious, and the cost of corrosion is increasing daily [1]. According to statistics, the consumption of the petroleum industry is approximately CNY 10 billion for OCTG every year in China, and most of these expenses are caused by corrosion [2,3]. With the rapid development of the petroleum industry, drilling conditions are becoming increasingly severe. The common materials used in oil and gas fields, such as carbon steel and conventional stainless steel, have been unable to meet the needs for the technical development of modern drilling and the extraction and transportation of oil and gas [4,5,6].
Titanium alloy has been widely used in many fields because of its high strength-to-weight ratio, its excellent mechanical properties and its good corrosion resistance [7,8,9,10,11,12,13,14,15]; among them, the excellent corrosion resistance is attributed to the spontaneous oxide passive film (TiO2) on the surface of titanium alloy [16,17]. The stability of the passive film on the surface of titanium alloy exposed to the working environment depends on its electronic properties, which are closely related to the electrochemical reactivity and the redox reaction at the metal/oxide interface [9].
The passive film on titanium alloy is only a few nanometers thick, and the adsorption characteristics of corrosive species such as Cl, HS and HCO3 on the surface of corrosion-resistant alloys have only been studied on a limited basis. Therefore, it is difficult to study the thermodynamic stability of the interface between the passive film and the solution and to further study the electronic structure, bonding type and bonding strength after adsorption on the surface of passive film only by experimental means. First-principles calculations, especially the density functional theory (DFT) method, can link the microscopic properties of metal materials with their structural and thermodynamic properties, providing a powerful tool for the study of adsorption systems [18]. Fu [19] found that the adsorption capacity of H2S at the top and bridge positions on the Mo2C (001) surface was also weak, with adsorption energies of 23.23 kcal/mol and 26.12 kcal/mol, respectively. Lin [20] believed that when CO molecules were adsorbed on the surface of SiC (111), the adsorption of O atoms in CO molecules in the vertical direction of Si atoms was the most stable, and the adsorption energy was −1.24 eV. Lin [21] argued that there was almost no charge transferred from the SiC (111) surface to the CH4 molecule during the adsorption of CH4 on the SiC (111) surface, and it could be seen that the adsorption mechanism was physical adsorption based on the diagram of the electron density difference and the density of states.
Therefore, the adsorption characteristics (electronic structure and interfacial binding energy) of corrosive ions (Cl, HS, S2−, HCO3 and CO32−) on the passive film (TiO2) of TC4 titanium alloy were investigated by using the CASTEP module in the first-principles Materials Studio simulation software based on density functional theory, and the thermodynamic stability characteristics of the passive film interface on TC4 titanium alloy were obtained. Finally, the corrosion resistance mechanism of TC4 titanium alloy in a harsh corrosion environment (CO2-H2S-Cl medium system) was revealed from the micro perspective to provide a theoretical basis for the development of titanium alloy and its applicability in oil and gas fields.

2. Modeling and Computing

TiO2 is a metal oxide semiconductor with polycrystalline properties [22], and it has three types of crystalline structures: rutile, anatase and brookite. They are all composed of octahedral structures, but their arrangement, linkage and lattice structure are different [23]. Rutile TiO2 belongs to the tetragonal crystal system structure, and its space group is P42/MNM [24], as shown in Figure 1a, around which each octahedron is connected to ten octahedrons, with two common edges and eight common apex angles. Each protocell consists of two TiO2 molecules, and thus, the molecular formula is Ti2O4. Anatase TiO2 belongs to the tetragonal crystal system structure, as shown in Figure 1b. Each octahedron is connected to eight octahedrons, with four common sides and four common apical angles. Each primitive cell consists of four TiO2 molecules, and thus, the molecular formula is Ti4O8. Brookite TiO2 belongs to the orthorhombic crystal system, as shown in Figure 1c. Each primitive cell consists of six TiO2 molecules, and thus, the molecular formula is Ti6O12.
The rutile phase is stable under atmospheric conditions and is also the main component phase of passive films on titanium alloys. Burnside and Labat [25,26] simulated the relative energy of rutile TiO2 (110), (100) and (001) crystal faces and found that the relative energy of the TiO2 (110) crystal face was the lowest and most stable. Studies have shown that the (110) crystal face of TiO2 was the close-packed surface and the thermodynamically stable surface with the lowest energy [27,28]. Zhao [29] also obtained that the rutile phase characteristic peak was the (110) face by XRD analysis and found that the 0.320 nm lattice fringes shown in the sample correspond to the (110) face of the rutile phase through HR-TEM. Therefore, the (110) crystal face of rutile TiO2 was selected for the first-principles calculation in this paper.

2.1. Modeling

The interface characteristics of all the adsorption models were calculated by the CASTEP module based on density functional theory (DFT) in Material Studio software [30,31]. Taking full account of the degree of conformity with the actual situation, the actual calculation strength of the server and the calculation requirements of CASTEP, the supercell structure of the TiO2 (110) face in three-dimensional space with 2 × 1 × 1 periodic boundary conditions was established. To avoid the interaction between the plates, a vacuum region of 15 Å was added between the plates. Figure 2 shows the interface models of different corrosive species (Cl, HS, S2−, HCO3 and CO32−) at three adsorption positions (top, bridge and hole) on the TiO2 (110) crystal face.

2.2. Computational Methods

The PBE functional form under the generalized gradient approximation (GGA) was used for all calculations, and the ultrasoft pseudopotential self-consistent field (SCF) was used to construct the pseudopotential [32,33,34]. The plane wave cutoff energy was 425 eV, and the self-consistent iteration was 500 times. The convergence accuracy was 2 × 10−6 eV/atom, the force converged to 0.03 eV/atom, the stress deviation was less than 0.08 GPa, the tolerance deviation was less than 0.005 and the value of the k point in the Brillouin zone was 4 × 4 × 1.

3. Results and Discussion

3.1. Optimum Structure and Stable Adsorption Configuration

(1)
Optimum structure
Table 1 is a comparison of the lattice constants of the rutile TiO2 in this paper to those in the literature [35]. The errors of “a” and “c” were only ±0.02% and ±0.07%, respectively, indicating that the constructed model conformed to the actual requirements.
(2)
Stable adsorption configuration
Since each atom in the nonrelaxation state of TiO2 is fixed, it is equivalent to the mechanical accumulation in a specific position. However, the atoms in the relaxation state of TiO2 can move within the crystal cell. After the geometric optimization of the adsorption structures with corrosive species adsorbed on the surface of TiO2, an optimal position was needed, in which the energy of the system was the lowest—that is, the system was the most stable at this time. Therefore, the geometries of all adsorption positions were optimized to select the most stable adsorption configuration. Table 2 shows the final energies of Cl, HS, S2−, HCO3 and CO32− at each adsorption position (top, bridge and hole) on the TiO2 (110) crystal face after geometric optimization.
The final energies of Cl, HS, S2−, HCO3 and CO32− at the bridge position on the TiO2 (110) crystal face were the lowest; the lower the total cell energy is, the more stable the cell structure is [36]. Therefore, the optimal adsorption positions of Cl, HS, S2−, HCO3 and CO32− on the TiO2 (110) surface were all bridge positions, and the final energy order of the corrosive species at the bridge position on the TiO2 (110) crystal face was S2− > HS > Cl > CO32− > HCO3.

3.2. Electron Density

Figure 3 shows the charge density distributions of Cl, HS, S2−, HCO3 and CO32− at the bridge position on the TiO2 (110) crystal face, which can directly reflect the bonding characteristics between atoms [37].
As shown in Figure 3, when the adsorption of corrosive species on the TiO2 (110) surface reached a stable state, there was a strong charge interaction between the Cl atoms in Cl, the S atoms in HS-, the S atoms in S2−, the O atoms in HCO3, the O atoms in CO32− and the Ti atoms on TiO2 [38]. The red region represents a large density of electrons, which is an active region available for chemical reactions. The deeper the color of the red, the larger the charge density [39]. Interface bonding [40] mainly occurred between the negatively charged atoms in the anion and the positively charged Ti atoms in the TiO2 (110) crystal face.
Table 3 shows the charge number of the negatively charged atoms in each corrosive species. The absolute value order of the charge number of Cl atoms in the Cl, S atoms in HS, the S atoms in S2−, the O atoms in HCO3 and the O atoms in CO32− was S2− > CO32− > Cl > HS > HCO3. The charge density represents the number of valence electrons in a volume of pure physical space, and the larger the charge density is, the easier the ions are adsorbed on the metal surface, resulting in the stronger corrosive effect of the ions on the passive film. Therefore, the stability of TiO2 in the medium containing S2− was the worst, followed by CO32−, Cl, HS and HCO3.

3.3. Electron Density Difference

The electron density difference can verify the relevant conclusions of the charge density diagram and more intuitively observe the charge transfer before and after adsorption [41] and the bonding situation [42]. The analysis of the electron density difference was conducted on the steady-state adsorption configuration of Cl, HS, S2−, HCO3 and CO32− adsorbed at the crystal plane bridge position of TiO2 (110), as shown in Figure 4.
When the adsorption of corrosive species on the TiO2 (110) surface reached a stable state, there was an obvious charge transfer phenomenon between the Cl atoms in Cl, the S atoms in HS, the S atoms in S2−, the O atoms in HCO3-, the O atoms in CO32− and the Ti atoms on the surface of TiO2. In addition, the charge accumulation and electronegativity around the negatively charged atoms in the species decreased, while the charge dissipation and electronegativity around the Ti atoms increased. The interface bonding mainly existed between the Cl atoms in Cl, the S atoms in HS-, the S atoms in S2−, the O atoms in HCO3, the O atoms in CO32− and the Ti atoms. The bonding was mainly caused by the charge moving from around the Ti atom to around the Cl atom in Cl, the S atom in HS, the S atom in S2−, the O atoms in HCO3 and the O atoms in CO32−.

3.4. Density of States

The electron transfer of the interaction between atoms mainly occurs with the valence electrons of each atom. To further investigate the nature of electronic interactions in interface bonding after adsorption and understand the contribution of the electrons of each element to the total density of states of the alloy [43], partial density of states (PDOS) analysis of the steady-state adsorption configurations of Cl, HS, S2−, HCO3 and CO32− at the bridge position of the TiO2 (110) crystal face was carried out, as shown in Figure 5.
Figure 5 shows that there was a certain degree of charge interaction between the Cl atoms in Cl, the S atoms in HS, the S atoms in S2−, the O atoms in HCO3 and the O atoms in CO32− with the Ti atoms on the surface of TiO2, indicating that the adsorption processes were all chemisorption [44]. The charge interactions took place near −6 eV and 0.5 eV~1.5 eV, 0.5 eV~1.5 eV, −6 eV and 0.5 eV, 1 eV, −10.5 eV and −6 eV and 0.5 eV, respectively, which were mainly formed by the hybrid orbitals between the 3d2 electrons of Ti and the 3p5 electrons of Cl, the 3p4 electrons of S and the 2p4 electrons of the O atoms. Therefore, after the adsorption of Cl, HS, S2−, HCO3 and CO32− on the TiO2 (110) crystal surface, the interface bonding was mainly caused by the electronic orbital hybridization of Cl-3p5, S-3p4, O-2p4 and Ti-3d2, and the bonding energy levels of the Cl, HS, S2−, HCO3 and CO32− ions were mainly located near the corresponding charge position mentioned above.

3.5. Interface Binding Energy

The stability of the interface can be quantitatively determined by the interface binding energy to determine the corrosion attack of corrosive species in the solution on metallic materials. The calculation equation of the interface binding energy is as follows [35]:
E interface = E t E m + E i
where Et is the total energy of the entire adsorption system after geometric optimization, Em is the energy of the TiO2 (110) system and Ei is the energy of the corrosive medium system (Cl, HS, S2−, HCO3 and CO32−).
Table 4 shows the final energies of the single Cl, HS, S2−, HCO3, CO32− and TiO2 (110) models after geometric optimization. The binding energies of the Cl, HS, S2−HCO3, CO32− and TiO2 (110) surfaces at bridge positions were obtained by Table 2 and Table 4, as shown in Table 5. Table 5 shows that, compared with Cl, HS, HCO3 and CO32−, it was easier for S2− to bind and react with TiO2, indicating that S2− had strong adsorption on the TiO2 (110) surface. The interface binding energy is an important parameter to characterize the interface thermodynamic stability; the smaller the value is, the more stable the interface structure is [35]. Therefore, the stability of TiO2 in this environment containing S2- was the poorest. The binding energy order of Cl, HS, S2−, HCO3 and CO32− with the TiO2 (110) surface was S2− > CO32− > Cl > HS > HCO3. The stability of TiO2 in the environments containing Cl, HS, S2−, HCO3 and CO32− was S2− < CO32− < Cl < HS < HCO3, which is in good accordance with the electron density results.

4. Corrosion Mechanism

According to the above study, when corrosive ions (Cl, HS, S2−, HCO3 and CO32−) adsorbed on the TiO2 (110) surface, their order of corrosion attack on the TiO2 (110) surface was S2− > CO32− > Cl > HS > HCO3. The corrosion resistance of TiO2 in the environment containing H2S was the worst, followed by CO2, and the stability of TiO2 in the environment containing Cl was better than the former two. In addition, the TC4 titanium alloy would suffer more serious corrosion in the CO2-H2S-Cl system due to the interaction between corrosive ions. The corrosion mechanism of the TC4 titanium alloy in the NaCl solution containing H2S and CO2 is shown in Figure 6.
The following is the specific reaction process of each corrosive ion (Cl, HS, S2−, HCO3 and CO32−) on the passive film of the TC4 titanium alloy surface. The specific oxidation process of titanium is as follows [45,46]:
Ti 2 e Ti 2 +
Ti 2 + e Ti 3 +
Ti 3 + e Ti 4 +
The total reaction equation is:
Ti 2 + + O 2 TiO
2 Ti 3 + + 3 O 2 Ti 2 O 3
Ti 4 + + 2 O 2 TiO 2
The oxide film formed is mainly TiO2. In the environment where titanium alloy is in service, H2S will undergo an ionization reaction to ionize HS, S2− and H+, and the specific ionization process is as follows [47]:
2 H 2 S ( aq ) 2 HS + 2 H +
HS ( aq ) S 2 + H +
Obviously, the local concentration of HS increases with increasing sulfide concentrations, and the concentration of S2− increases with increasing HS- concentrations. The specific formation process of TiS2 is mainly manifested in two situations [46]. The first one is shown in Equation (10), and the second one may be caused by chemical reaction after the electron transfer reaction. The specific reaction process is shown in Equation (11).
TiO 2 + HS + 2 H + TiS 2 + 2 H 2 O
Ti 4 + + 2 S 2 TiS 2
The corrosion of titanium alloy in a saturated CO2 water environment is also electrochemical corrosion. CO2 is dissolved in water to form H2CO3. After ionization, H2CO3 generates HCO3 and CO32−. The specific reaction steps are as follows [47]:
CO 2 + H 2 O H 2 CO 3
H 2 CO 3 HCO 3 + H +
HCO 3 CO 3 2 + H +
According to Equations (2)–(14) and the binding energy between the corrosive ions and the corrosive ions with the rutile TiO2 (110) surface, when the initial corrosion system only contains the NaCl medium, Cl has a strong penetration, which can destroy the passive film TiO2 on the titanium alloy surface. However, when H2S and CO2 are introduced into the system, although Cl has a superadsorption effect, the interfacial binding energy of Cl between TiO2 is lower than those of S2− and CO32−, and the interfacial binding energy of S2− between TiO2 is higher than that of CO32−. The literature showed that, in a corrosive environment where CO2 and H2S coexisted, the chemical adsorption capacity of HS was stronger than that of HCO3 [48].
Therefore, S2− will first exclude Cl and show a high concentration on the surface of the passive film. At this time, the damage of S2− on the titanium alloy surface passive film TiO2 is stronger than that of Cl. Therefore, TiO2 firstly reacts with H2S (or its hydrolysate) in the interface between the bulk solution and the metal matrix to form TiS2. At the same time, HS- and S2− ionized by H2S can promote the formation of TiS2 compounds [45]. TiS2 is mainly contained in the outer layer of the passive film. However, the source of some TiS2 can be attributed to the competition of sulfur and oxygen for titanium bond orbitals, which is actually a central issue [49]. Since S and O are elements in the same main group, S more easily replaces O in passive film TiO2. That is, the nucleation and growth of partial TiS2 is a process of replacing Ti-O bonds with Ti-S bonds [49]. Some researchers believe that the formation of sulfides on metals is much faster than the formation of corresponding oxides [50].
When the formed corrosion scale TiS2 is deposited on the substrate surface and reaches a certain saturation state, the formation of the corrosion scale TiS2 gradually replaces the passive film TiO2. At the moment, CO32− ionized from H2CO3 formed by CO2 dissolved in water will act on TiO2 to generate titanium carbonate. As a typical weak acid and alkali salt, titanium carbonate is easily hydrolyzed, but it will cause the dissolution of passive film or corrosion product film. Therefore, the existence of CO2 will promote the dissolution of the TiS2 film, destroy the structure of the TiS2 film as well as the integrity of the passive film and corrosion scale and further erode the metal matrix.
H2CO3 formed by CO2 dissolved in water provides a further acidified environment for Cl corrosion attack, resulting in a significant decrease in the stability of the passive film (or corrosion product film) in an acidic medium. At the same time, Cl itself has a corresponding influence on the formation of the corrosion product layer [51]. With the influence of H2CO3, Cl further destroys the integrity of TiS2 on the metal matrix and accelerates the dissolution of the matrix [52], resulting in an increase in the number and depth of corrosion pits. The above discussion is consistent with the experimental results in the relevant literature: compared with the CO2 corrosion environment in formation water, the passive film on the surface of titanium alloy is more vulnerable to damage in the water solution containing the H2S corrosion medium [35]. Therefore, TC4 titanium alloy will suffer serious corrosion in the environment containing H2S, followed by the environment containing CO2, and it has relatively good corrosion resistance to Cl. When the corrosions H2S, CO2 and Cl coexist in the corrosion medium, TC4 titanium alloy will suffer more serious corrosion.

5. Conclusions

(1)
The optimal adsorption positions of Cl, HS, S2−, HCO3 and CO32− on the surface of TiO2 (110) were all bridge positions, followed by hole and top positions.
(2)
When the corrosive ion adsorption on TiO2 (110) reached a stable state, there was a strong charge interaction between the negatively charged Cl, S, O atoms in Cl, HS, S2−, HCO3 and CO32− and the positively charged Ti atoms in TiO2. The bonding was caused by the transfer of the charge from around the Ti atom to around the Cl, O and S atoms, forming the electron orbital hybridization of Cl-3p5, S-3p4, O-2p4 and Ti-3d2, and the adsorption mechanism was chemical adsorption.
(3)
The binding energies of Cl, HS, S2−, HCO3 and CO32− with TiO2 (110) were in the order of S2− > CO32− > Cl > HS > HCO3. Titanium alloy would be corroded in the system containing S2−, followed by CO32−, Cl, HS- and HCO3, and the combined action of H2S, CO2 and Cl- further accelerated the corrosion of titanium alloy.

Author Contributions

Conceptualization, P.D. and S.Z.; methodology, P.D. and Y.Z.; software, H.M.; validation, P.D.; formal analysis, Q.L.; investigation, Z.N.; data curation, J.L.; writing—original draft preparation, P.D.; writing—review and editing, S.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (51974245) and the Key Research and Development Program of Shaanxi Province (2022GY-128, 2022SF-045).

Data Availability Statement

Not applicable.

Acknowledgments

We thank the associate editor and the reviewers for their useful feedback that improved this paper, along with the Youth Innovation Team of Shaanxi University for the helpful discussions on topics related to this work.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses or interpretation of data; in the writing of the manuscript; or in the decision to publish the results. All authors agreed to submit the manuscript to the journal.

References

  1. Al-Moubaraki, A.H.; Obot, I.B. Top of the line corrosion: Causes, mechanisms, and mitigation using corrosion inhibitors. Arab. J. Chem. 2021, 14, 103116. [Google Scholar] [CrossRef]
  2. Hou, B.R.; Li, X.G.; Ma, X.M.; Du, C.W.; Zhang, D.W.; Zheng, M.; Xu, W.C.; Lu, D.Z.; Ma, F.B. The cost of corrosion in China. NPJ Mater. Degrad. 2017, 1, 4. [Google Scholar] [CrossRef]
  3. Wang, Z.Q.; Zhou, Z.Y.; Xu, W.C.; Yang, L.H.; Zhang, B.B.; Li, Y.T. Study on inner corrosion behavior of high strength product oil pipelines. Eng. Fail. Anal. 2020, 115, 104659. [Google Scholar] [CrossRef]
  4. Stipaničev, M.; Turcu, F.; Esnault, L.; Schweitzer, E.W.; Kilian, R.; Basseguy, R. Corrosion behavior of carbon steel in presence of sulfate-reducing bacteria in seawater environment. Electrochim. Acta 2013, 113, 390–406. [Google Scholar] [CrossRef] [Green Version]
  5. Xin, S.S.; Li, M.C. Electrochemical corrosion characteristics of type 316L stainless steel in hot concentrated seawater. Corros. Sci. 2014, 81, 390–406. [Google Scholar] [CrossRef]
  6. Hoseinieh, S.M.; Homborg, A.M.; Shahrabi, T.; Mol, J.M.C.; Ramezanzadeh, B. A novel approach for the evaluation of under deposit corrosion in marine environments using combined analysis by electrochemical impedance spectroscopy and electrochemical noise. Electrochim. Acta 2016, 217, 226–241. [Google Scholar] [CrossRef]
  7. Su, B.X.; Wang, B.B.; Luo, L.S.; Wang, L.; Su, Y.Q.; Wang, F.X.; Xu, Y.J.; Han, B.S.; Huang, H.G.; Guo, J.J.; et al. The corrosion behavior of Ti-6Al-3Nb-2Zr-1Mo alloy: Effects of HCl concentration and temperature. J. Mater. Sci. Technol. 2021, 74, 143–154. [Google Scholar] [CrossRef]
  8. Tekin, K.C.; Malayoglu, U. Production of plasma electrolytic oxide coatings on Ti-6Al-4V alloy in aluminate-based electrolytes. Surf. Eng. 2017, 33, 787–795. [Google Scholar] [CrossRef]
  9. Gai, X.; Bai, Y.; Li, J.; Li, S.J.; Hou, W.T.; Hao, Y.L.; Zhang, X.; Yang, R.; Misra, R.D.K. Electrochemical behaviour of passive film formed on the surface of Ti-6Al-4V alloys fabricated by electron beam melting. Corros. Sci. 2018, 145, 80–89. [Google Scholar] [CrossRef]
  10. Seo, D.I.; Lee, J.B. Effects of competitive anion adsorption (Br or Cl) and semiconducting properties of the passive films on the corrosion behavior of the additively manufactured Ti-6Al-4V alloys. Corros. Sci. 2020, 173, 108789. [Google Scholar] [CrossRef]
  11. Kaur, S.; Sharma, S.; Bala, N. A comparative study of corrosion resistance of biocompatible coating on titanium alloy and stainless steel. Mater. Chem. Phys. 2019, 238, 121923. [Google Scholar] [CrossRef]
  12. Yang, Z.; Gu, H.; Sha, G.; Lu, W.J.; Yu, W.Q.; Zhang, W.J.; Fu, Y.F.; Wang, K.S.; Wang, L.Q. TC4/Ag metal matrix nanocomposites modified by friction stir processing: Surface characterization, antibacterial property, and cytotoxicity in vitro. ACS Appl. Mater. Interfaces 2018, 10, 41155–41166. [Google Scholar] [CrossRef] [PubMed]
  13. Fazel, M.; Salimijazi, H.R.; Shamanian, M. Improvement of corrosion and tribocorrosion behavior of pure titanium by subzero anodic spark oxidation. ACS Appl. Mater. Interfaces 2018, 10, 15281–15287. [Google Scholar] [CrossRef]
  14. Ren, S.; Du, C.W.; Liu, Z.Y.; Li, X.G.; Xiong, J.H.; Li, S.K. Effect of fluoride ions on corrosion behaviour of commercial pure titanium in artificial seawater environment. Appl. Surf. Sci. 2020, 506, 144759. [Google Scholar] [CrossRef]
  15. Haldar, B.; Karmakar, S.; Saha, P.; Chattopadhyay, A.B. In situ multicomponent MMC coating developed on Ti–6Al–4V substrate. Surf. Eng. 2014, 30, 256–262. [Google Scholar] [CrossRef]
  16. Tang, J.; Luo, H.Y.; Qi, Y.M.; Xu, P.W.; Ma, S.; Zhang, Z.; Ma, Y. The effect of cryogenic burnishing on the formation mechanism of corrosion product film of Ti-6Al-4V titanium alloy in 0.9% NaCl solution. Surf. Coat. Technol. 2018, 345, 123–131. [Google Scholar] [CrossRef]
  17. Rahimipour, S.; Rafiei, B.; Salahinejad, E. Organosilane-functionalized hydrothermal-derived coatings on titanium alloys for hydrophobization and corrosion protection. Prog. Org. Coat. 2020, 142, 105594. [Google Scholar] [CrossRef]
  18. Xing, H.R.; Hu, P.; Li, S.L.; Zuo, Y.G.; Han, J.Y.; Hua, X.J.; Wang, K.S.; Yang, F.; Feng, P.F.; Chang, T. Adsorption and diffusion of oxygen on metal surfaces studied by first-principle study: A review. J. Mater. Sci. Technol. 2021, 62, 180–194. [Google Scholar] [CrossRef]
  19. Fu, D.L.; Guo, W.Y.; Liu, Y.J.; Chi, Y.H. Adsorption and dissociation of H2S on Mo2C(001) surface-A first-principle study. Appl. Surf. Sci. 2015, 351, 125–134. [Google Scholar] [CrossRef]
  20. Lin, L.; Yao, L.W.; Li, S.F.; Shi, Z.G.; Xie, K.; Zhu, L.H.; Tao, H.L.; Zhang, Z.Y. Effect of SiC(111) with different layers on adsorption properties of CO molecules. Mater. Today Commun. 2020, 25, 101596. [Google Scholar] [CrossRef]
  21. Lin, L.; Yao, L.W.; Li, S.F.; Zhu, L.H.; Huang, J.T.; Wang, P.T.; Yu, W.Y.; He, C.Z.; Zhang, Z. The influence of SiC(111) surface with different layers on CH4 adsorption. Surf. Sci. 2020, 702, 121699. [Google Scholar] [CrossRef]
  22. Pantaroto, H.N.; Cordeiro, J.M.; Pereira, L.T.; de Almeida, A.B.; Junior, F.H.N.; Rangel, E.C.; Azevedo Neto, N.F.; da Silva, J.H.D.; Barão, V.A.R. Sputtered crystalline TiO2 film drives improved surface properties of titanium-based biomedical implants. Mater. Sci. Eng. C 2021, 119, 111638. [Google Scholar] [CrossRef]
  23. Mattsson, A.; Österlund, L. Adsorption and photoinduced decomposition of acetone and acetic acid on anatase, brookite, and rutile TiO2 nanoparticles. J. Phys. Chem. C 2010, 114, 14121–14132. [Google Scholar] [CrossRef]
  24. Cui, Y.Y.; Wang, Q.F.; Ren, J.S.; Liu, B.; Yang, G.; Gao, Y.F. Geometric and electronic properties of rutile TiO2 with vanadium implantation: A first-principles calculation. Nucl. Inst. Methods Phys. Res. B 2019, 455, 35–38. [Google Scholar] [CrossRef]
  25. Morgan, B.J.; Watson, G.W. A DFT+ U description of oxygen vacancies at the TiO2 rutile (110) surface. Surf. Sci. 2007, 601, 5034–5041. (In Chinese) [Google Scholar] [CrossRef]
  26. Burnside, S.D.; Shklover, V.; Barbé, C.; Comte, P.; Arendse, F.; Brooks, K.; Grätzel, M. Self-organization of TiO2 nanoparticles in thin films. Chem. Mater. 1998, 10, 2419–2425. [Google Scholar] [CrossRef]
  27. Wang, Y.; Huang, G.S. Adsorption of Au atoms on stoichiometric and reduced TiO2(110) rutile surfaces: A first principles study. Surf. Sci. 2003, 542, 72–80. [Google Scholar] [CrossRef]
  28. Käckell, P.; Terakura, K. First-principle analysis of the dissociative adsorption of formic acid on rutile TiO2(110). Appl. Surf. Sci. 2000, 166, 370–375. [Google Scholar] [CrossRef]
  29. Zhao, Y.F.; Ren, S.Y.; Wu, J.; Ji, Z.H.; Ma, Q.; Yang, M. Experimental study on the fabrication of TiO2 phase-junction nanorods and deep removal of mercury from flue gas by photocatalyst. J. Chin. Soc. Power Eng. 2021, 41, 979–983+1018. [Google Scholar] [CrossRef]
  30. Lin, L.; Shi, Z.G.; Yan, L.B.; Tao, H.L.; Yao, L.W.; Li, S.F.; Xie, K.; Huang, J.T.; Zhang, Z.Y. First-principles study of CO adsorption on Os atom doped anatase TiO2(101) surface. Polyhedron 2020, 191, 114814. [Google Scholar] [CrossRef]
  31. Guo, Y.X.; Hu, R.M.; Zhou, X.L.; Yu, J.; Wang, L.H. A first principle study on the adsorption of H2O2 on CuO(111) and Ag/CuO(111) surface. Appl. Surf. Sci. 2019, 479, 989–996. [Google Scholar] [CrossRef]
  32. Su, K.; Liu, D.M.; Pang, H.; Shao, T.M. Improvement on thermal stability of TiAlSiN coatings deposited by IBAD. Surf. Eng. 2018, 34, 504–510. [Google Scholar] [CrossRef]
  33. Guo, W.B.; She, Z.Y.; Xue, H.T.; Zhang, X.M. Effect of active Ti element on the bonding characteristic of the Ag(111)/α-Al2O3(0001) interface by using first principle calculation. Ceram. Int. 2020, 46, 5430–5435. [Google Scholar] [CrossRef]
  34. Shi, Z.G.; Lin, L.; Chen, R.X.; Yan, L.B. Adsorption of CO molecules on anatase TiO2(001) loaded with noble metals M (M=Ir/Pd/Pt): A study from DFT calculations. Mater. Today Commun. 2021, 28, 102699. [Google Scholar] [CrossRef]
  35. Li, Y.Y. Study on stability of titanium alloy passivation film under severe corrosion environment. J. Xi’an Shiyou Univ. Nat. Sci. Ed. 2018, 33, 120–126. [Google Scholar]
  36. Wen, P.; Li, F.C.; Zhao, Y.; Zhang, F.C.; Tong, L.H. First principles calculation of occupancy, bonding characteristics and alloying effect of Cr, Mo, Ni in bulk alpha-Fe(C). Acta Phys. Sin. 2014, 63, 809. [Google Scholar] [CrossRef]
  37. Mao, Y.L.; Hao, W.P.; Wei, X.L.; Yuan, J.M.; Zhong, J.X. Edge-adsorption of potassium adatoms on graphene nanoribbon: A first principle study. Appl. Surf. Sci. 2013, 280, 698–704. [Google Scholar] [CrossRef]
  38. Yang, S.Q.; Liang, G.X.; Lv, M. First-principle study on adsorption energy and electronic characteristics of Ni plating interface. Hot Work. Technol. 2021, 50, 27–30. [Google Scholar] [CrossRef]
  39. Li, W.; Lu, X.M.; Li, G.Q.; Ma, J.J.; Zeng, P.Y.; Chen, J.F.; Pan, Z.L.; He, Q.Y. First-principle study of SO2 molecule adsorption on Ni-doped vacancy-defected single-walled (8,0) carbon nanotubes. Appl. Surf. Sci. 2016, 364, 560–566. [Google Scholar] [CrossRef]
  40. Gao, X.; Zhou, Q.; Wang, J.X.; Xu, L.N.; Zeng, W. Adsorption of SO2 molecule on Ni-doped and Pd-doped graphene based on first-principle study. Appl. Surf. Sci. 2020, 517, 146180. [Google Scholar] [CrossRef]
  41. Lin, L.; Shi, Z.G.; Huang, J.T.; Wang, P.T.; Yu, W.Y.; He, C.Z.; Zhang, Z.Y. Molecular adsorption properties of CH4 with noble metals doped onto oxygen vacancy defect of anatase TiO2(101) surface: First-principles calculations. Appl. Surf. Sci. 2020, 514, 145900. [Google Scholar] [CrossRef]
  42. Ren, Y.J.; Du, H.Y.; Du, C.Y.; Chen, J.; Li, W.; Qiu, W.; Hsu, J.P.; Jiang, J.Z. Influence of oxygen adsorption on the chemical stability and conductivity of transition metal ceramic coatings: First-principle calculations. Appl. Surf. Sci. 2019, 495, 143530. [Google Scholar] [CrossRef]
  43. Hu, Y.L.; Bai, L.H.; Tong, Y.G.; Deng, D.Y.; Liang, X.B.; Zhang, J.; Li, Y.J.; Chen, Y.X. First-principle calculation investigation of NbMoTaW based refractory high entropy alloys. J. Alloys Compd. 2020, 827, 153963. [Google Scholar] [CrossRef]
  44. Luo, H.; Wang, D.Y.; Liu, L.X.; Li, L.C. Theoretical study on the adsorption characteristics of trichlorophenol on TiO2(101) surface. J. At. Mol. Phys. 2020, 37, 349–353. [Google Scholar]
  45. Yang, X.J.; Du, C.W.; Wan, H.X.; Liu, Z.Y.; Li, X.G. Influence of sulfides on the passivation behavior of titanium alloy TA2 in simulated seawater environments. Appl. Surf. Sci. 2018, 458, 198–209. [Google Scholar] [CrossRef]
  46. Wei, B.X.; Pang, J.Y.; Xu, J.; Sun, C.; Zhang, H.W.; Wang, Z.Y.; Yu, C.K.; Ke, W. Microbiologically influenced corrosion of TiZrNb medium-entropy alloys by Desulfovibrio desulfuricans. J. Alloys Compd. 2021, 875, 160020. [Google Scholar] [CrossRef]
  47. Gong, Q.J.; Xiang, Y.; Zhang, J.Q.; Wang, R.T.; Qin, D.H. Influence of elemental sulfur on the corrosion mechanism of X80 steel in supercritical CO2-saturated aqueous phase environment. J. Supercrit. Fluids 2021, 176, 105320. [Google Scholar] [CrossRef]
  48. Wang, Y.; Wang, B.; He, S.; Zhang, L.; Xing, X.J.; Li, H.X.; Lu, M.X. Unraveling the effect of H2S on the corrosion behavior of high strength sulfur-resistant steel in CO2/H2S/Cl- environments at ultra high temperature and high pressure. J. Nat. Gas Sci. Eng. 2022, 100, 104477. [Google Scholar] [CrossRef]
  49. Li, Y.F.; Singh, A.; Reidy, K.; Jo, S.S.; Ross, F.M.; Jaramillo, R. Making large-area titanium disulfide films at reduced temperature by balancing the kinetics of sulfurization and roughening. Adv. Funct. Mater. 2020, 30, 2003617. [Google Scholar] [CrossRef]
  50. Mantha, D.; Wen, X.; Reddy, R.G. High temperature oxidation of a Ti3Al alloy in argon-5% SO2 environment. High Temp. Mater. Processes 2004, 23, 93–102. [Google Scholar] [CrossRef]
  51. Wei, B.X.; Chen, C.J.; Xu, J.; Yang, L.L.; Jia, Y.X.; Du, Y.; Guo, M.X.; Sun, C.; Wang, Z.Y.; Wang, F.H. Comparing the hot corrosion of (100), (210) and (110) Ni-based superalloys exposed to the mixed salt of Na2SO4-NaCl at 750 °C: Experimental study and first-principles calculation. Corros. Sci. 2022, 195, 109996. [Google Scholar] [CrossRef]
  52. Qin, M.; Liao, K.X.; He, G.X.; Huang, Y.J.; Wang, M.N.; Zhang, S.J. Main control factors and prediction model of flow-accelerated CO2/H2S synergistic corrosion for X65 steel. Process Saf. Environ. Prot. 2022, 160, 749–762. [Google Scholar] [CrossRef]
Figure 1. The crystal structure of TiO2 (light gray atoms represent Ti; red atoms represent O): (a) rutile TiO2; (b) anatase TiO2; (c) brookite TiO2.
Figure 1. The crystal structure of TiO2 (light gray atoms represent Ti; red atoms represent O): (a) rutile TiO2; (b) anatase TiO2; (c) brookite TiO2.
Metals 12 01160 g001
Figure 2. Interface model of TiO2 (110) containing different corrosive species: (a) TiO2 (top)—Cl; (b) TiO2 (bridge)—Cl; (c) TiO2 (hole)—Cl; (d) TiO2 (top)—HS; (e) TiO2 (bridge)—HS; (f) TiO2 (hole)—HS; (g) TiO2 (top)—S2−; (h) TiO2 (bridge)—S2−; (i) TiO2 (hole)—S2−; (j) TiO2 (top)—HCO3; (k) TiO2 (bridge)—HCO3; (l) TiO2 (hole)—HCO3; (m) TiO2 (top)—CO32−; (n) TiO2 (bridge)—CO32−; (o) TiO2 (hole)—CO32−.
Figure 2. Interface model of TiO2 (110) containing different corrosive species: (a) TiO2 (top)—Cl; (b) TiO2 (bridge)—Cl; (c) TiO2 (hole)—Cl; (d) TiO2 (top)—HS; (e) TiO2 (bridge)—HS; (f) TiO2 (hole)—HS; (g) TiO2 (top)—S2−; (h) TiO2 (bridge)—S2−; (i) TiO2 (hole)—S2−; (j) TiO2 (top)—HCO3; (k) TiO2 (bridge)—HCO3; (l) TiO2 (hole)—HCO3; (m) TiO2 (top)—CO32−; (n) TiO2 (bridge)—CO32−; (o) TiO2 (hole)—CO32−.
Metals 12 01160 g002aMetals 12 01160 g002bMetals 12 01160 g002c
Figure 3. Charge density distribution of corrosive species at the bridge position on the TiO2 (110) surface: (a) Cl; (b) HS; (c) S2−; (d) HCO3; (e) CO32−.
Figure 3. Charge density distribution of corrosive species at the bridge position on the TiO2 (110) surface: (a) Cl; (b) HS; (c) S2−; (d) HCO3; (e) CO32−.
Metals 12 01160 g003
Figure 4. Electron density difference distribution of corrosive species at the bridge position on the TiO2 (110) crystal face: (a) Cl; (b) HS; (c) S2−; (d) HCO3; (e) CO32−.
Figure 4. Electron density difference distribution of corrosive species at the bridge position on the TiO2 (110) crystal face: (a) Cl; (b) HS; (c) S2−; (d) HCO3; (e) CO32−.
Metals 12 01160 g004
Figure 5. PDOS curves of corrosive species adsorbed on the bridge position of the TiO2 (110) surface (note: the valence electrons of Cl, S, O and Ti are 3S23P5, 3S23P4, 2S22P4 and 3p63d24s2, respectively): (a) Cl; (b) HS; (c) S2−; (d) HCO3; (e) CO32−.
Figure 5. PDOS curves of corrosive species adsorbed on the bridge position of the TiO2 (110) surface (note: the valence electrons of Cl, S, O and Ti are 3S23P5, 3S23P4, 2S22P4 and 3p63d24s2, respectively): (a) Cl; (b) HS; (c) S2−; (d) HCO3; (e) CO32−.
Metals 12 01160 g005
Figure 6. Corrosion mechanism of the TC4 titanium alloy in the CO2-H2S-Cl system: (a) initial stage and (b) latter stage.
Figure 6. Corrosion mechanism of the TC4 titanium alloy in the CO2-H2S-Cl system: (a) initial stage and (b) latter stage.
Metals 12 01160 g006
Table 1. Comparison of the lattice constants of rutile TiO2 between the calculated values and the literature values.
Table 1. Comparison of the lattice constants of rutile TiO2 between the calculated values and the literature values.
Lattice ConstantCalculated Value/ÅLiterature Value/ÅError/%
a4.59404.5930±0.02
c2.95902.9610±0.07
c/a0.64410.6447±0.09
Table 2. The final energy of each corrosive species at each adsorption position on the TiO2 (110) crystal face after geometric optimization.
Table 2. The final energy of each corrosive species at each adsorption position on the TiO2 (110) crystal face after geometric optimization.
ModelFinal Energy/eV
TiO2 (top)—Cl−30,176.98682254
TiO2 (bridge)—Cl−30,177.36115795
TiO2 (hole)—Cl−30,177.35786327
TiO2 (top)—HS−30,061.99607022
TiO2 (bridge)—HS−30,062.32529929
TiO2 (hole)—HS−30,062.32232474
TiO2 (top)—S2−−30,045.12435310
TiO2 (bridge)—S2−−30,045.13031885
TiO2 (hole)—S2−−30,045.12848408
TiO2 (top)—HCO3−31,247.50645571
TiO2 (bridge)—HCO3−31,250.09178742
TiO2 (hole)—HCO3−31,249.39763267
TiO2 (top)—CO32−−31,233.00592263
TiO2 (bridge)—CO32−−31,233.01017140
TiO2 (hole)—CO32−−31,233.00957590
Table 3. Charge number of atoms with a negative charge in each corrosive species.
Table 3. Charge number of atoms with a negative charge in each corrosive species.
AtomCl (Cl)HS (S)S2 (S)HCO3 (O)CO32 (O)
Charge/e−0.20−0.17−0.41−0.14−0.22
Table 4. Final energy of each corrosive species and TiO2 (110) model after geometric optimization.
Table 4. Final energy of each corrosive species and TiO2 (110) model after geometric optimization.
ModelFinal Energy/eV
Cl−411.7437852366
HS−296.3156703881
S2−−280.4987150773
HCO3−1483.481392861
CO32−−1467.944344893
TiO2 (110)−29,766.99300605
Table 5. Interface binding energy between corrosive species and the bridge position of the TiO2 (110) crystal face.
Table 5. Interface binding energy between corrosive species and the bridge position of the TiO2 (110) crystal face.
ModelInterfacial Binding Energy/eV
TiO2 (bridge)—Cl1.3756333366
TiO2 (bridge)—HS0.9833771481
TiO2 (bridge)—S2−2.3614022773
TiO2 (bridge)—HCO30.382611491
TiO2 (bridge)—CO32−1.927179543
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dong, P.; Zhang, Y.; Zhu, S.; Nie, Z.; Ma, H.; Liu, Q.; Li, J. First-Principles Study on the Adsorption Characteristics of Corrosive Species on Passive Film TiO2 in a NaCl Solution Containing H2S and CO2. Metals 2022, 12, 1160. https://doi.org/10.3390/met12071160

AMA Style

Dong P, Zhang Y, Zhu S, Nie Z, Ma H, Liu Q, Li J. First-Principles Study on the Adsorption Characteristics of Corrosive Species on Passive Film TiO2 in a NaCl Solution Containing H2S and CO2. Metals. 2022; 12(7):1160. https://doi.org/10.3390/met12071160

Chicago/Turabian Style

Dong, Pan, Yanna Zhang, Shidong Zhu, Zhen Nie, Haixia Ma, Qiang Liu, and Jinling Li. 2022. "First-Principles Study on the Adsorption Characteristics of Corrosive Species on Passive Film TiO2 in a NaCl Solution Containing H2S and CO2" Metals 12, no. 7: 1160. https://doi.org/10.3390/met12071160

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