Next Article in Journal
Increasing Electrical Resistivity of P-Type BiFeO3 Ceramics by Hydrogen Peroxide-Assisted Hydrothermal Synthesis
Next Article in Special Issue
Erosion–Corrosion Behavior of Friction Stud Welded Joints of X65 Pipelines in Simulated Seawater under Different Flow Rates
Previous Article in Journal
Magnetic Biochar Obtained by Chemical Coprecipitation and Pyrolysis of Corn Cob Residues: Characterization and Methylene Blue Adsorption
Previous Article in Special Issue
Study of the Chlorine Influence on the Corrosion of Three Steels to Be Used in Water Treatment Municipal Facilities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption Characteristics between Ti Atoms of TiO2(100) and Corrosive Species of CO2-H2S-Cl System in Oil and Gas Fields

1
School of Materials Science and Engineering, Xi’an Shiyou University, Xi’an 710065, China
2
Shaanxi Key Laboratory of Carbon Dioxide Sequestration and Enhanced Oil Recovery, Shaanxi Yanchang Petroleum (Group) Co., Ltd., Xi’an 710065, China
3
School of Chemical Engineering, Northwest University Shaanxi, Xi’an 710069, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(8), 3129; https://doi.org/10.3390/ma16083129
Submission received: 8 March 2023 / Revised: 10 April 2023 / Accepted: 14 April 2023 / Published: 16 April 2023
(This article belongs to the Special Issue Advances in Surface Corrosion Science)

Abstract

:
The service environment of OCTG (Oil Country Tubular Goods) in oil and gas fields is becoming more and more severe due to the strong affinity between ions or atoms of corrosive species coming from solutions and metal ions or atoms on metals. While it is difficult for traditional technologies to accurately analyze the corrosion characteristics of OCTG in CO2-H2S-Cl systems, it is necessary to study the corrosion-resistant behavior of TC4 (Ti-6Al-4V) alloys based on an atomic or molecular scale. In this paper, the thermodynamic characteristics of the TiO2(100) surface of TC4 alloys in the CO2-H2S-Cl system were simulated and analyzed by first principles, and the corrosion electrochemical technologies were used to verify the simulation results. The results indicated that all of the best adsorption positions of corrosive ions (Cl, HS, S2−, HCO3, and CO32−) on TiO2(100) surfaces were bridge sites. A forceful charge interaction existed between Cl, S, and O atoms in Cl, HS, S2−, HCO3, CO32−, and Ti atoms in TiO2(100) surfaces after adsorption in a stable state. The charge was transferred from near Ti atoms in TiO2 to near Cl, S, and O atoms in Cl, HS, S2−, HCO3, and CO32−. Electronic orbital hybridization occurred between 3p5 of Cl, 3p4 of S, 2p4 of O, and 3d2 of Ti, which was chemical adsorption. The effect strength of five corrosive ions on the stability of TiO2 passivation film was S2− > CO32− > Cl > HS > HCO3. In addition, the corrosion current density of TC4 alloy in different solutions containing saturated CO2 was as follows: NaCl + Na2S + Na2CO3 > NaCl + Na2S > NaCl + Na2CO3 > NaCl. At the same time, the trends of Rs (solution transfer resistance), Rct (charge transfer resistance), and Rc (ion adsorption double layer resistance) were opposite to the corrosion current density. The corrosion resistance of TiO2 passivation film to corrosive species was weakened owing to the synergistic effect of corrosive species. Severe corrosion resulted, especially pitting corrosion, which further proved the simulation results mentioned above. Thus, this outcome provides the theoretical support to reveal the corrosion resistance mechanism of OCTG and to develop novel corrosion inhibitors in CO2-H2S-Cl environments.

1. Introduction

Corrosion has been considered as one of the major social problems in pipelines and industries using such materials since the early industrial revolution. A large number of accidents occur frequently, which leads to a greater threat to the safe production of oil and gas [1]. The annual cost of corrosion in China is about CNY 2.3 trillion, accounting for 3.3% of GDP [2].
In recent years, with the development of deep and ultra-deep wells to meet the social demand for energy, the working environment of tubing and casing is becoming more and more complex. In addition to the stringent service conditions, metal OCTG are inevitably subjected to different degrees of corrosion, and the working properties of OCTG have decreased. For example, the presence of H2S leads to severe localized corrosion, as well as cracks caused by stress and hydrogen [3]. In some special working conditions, CO2 and H2S exist at the same time [4], which greatly deteriorates the service environment of OCTG [5]. The high temperature, pressure, acid gas content, and Cl concentration of oil and gas wells increase the requirements for corrosion-resistant OCTG.
The TC4 titanium alloy (Ti-6Al-4V) is now considered to be the ideal material applied in oil and gas fields, accounting for about half of the market share of titanium alloys currently used in the world [6]. A dense TiO2 oxide film with a thickness of 4~6 nm of TC4 will be spontaneously formed at room temperature [7], which can effectively prevent the matrix in a solution from being corroded by corrosive ions (such as H+, Cl, etc.) [8]. However, the film is not always able to maintain its integrity; it is very likely to be destroyed in some medium containing some corrosive species, resulting in serious corrosion of the titanium alloy matrix [9]. It is difficult for traditional technologies to accurately analyze the corrosion characteristics of OCTG in CO2-H2S-Cl systems.
Therefore, the first-principles calculation software (Materials Studio) on account of DFT (Density Functional Theory) was selected to research the interface characteristics between the corrosive species and TiO2 passivation film on the surface of TC4 alloys in CO2-H2S-Cl systems containing Cl, HS, S2−, HCO3, and CO32− based on an atomic or molecular scale. Additionally, the corrosion characteristics of the TC4 alloy in NaCl, NaCl + Na2CO3, NaCl + Na2S, and NaCl + Na2S + Na2CO3 solutions containing saturated CO2 were carried out by the electrochemical technologies to verify the simulation results above.

2. Research Methods

2.1. First Principles

2.1.1. Modeling

TiO2 passivation film on titanium alloy surfaces has three crystal structures: rutile, anatase, and brookite [10]. Figure 1 shows the Raman spectra of TiO2 film on TC4 alloy, and Table 1 shows the frequency shift positions of Raman spectral characteristics of three crystalline TiO2. The four peaks, 145.53 cm−1, 241.76 cm−1, 612.53 cm−1, and 824.16 cm−1 in Figure 1, are consistent with the corresponding peak value of the rutile TiO2 in Table 1. Some scholars found that the composition of titanium alloy passivation film was rutile TiO2 [11]. Therefore, rutile TiO2 was selected as the research object in this paper.
There are (110), (100), and (001) low index surfaces in Rutile phase TiO2. The characteristics of the various ions on TiO2(110) surfaces have been studied, including our previous research [2], but few reports were focused on the adsorption of TiO2(100) and TiO2(001) surfaces. Furthermore, compared with TiO2(001) surfaces, TiO2(100) surfaces present a higher possibility of stable existence at high temperatures [12]. Therefore, the adsorption properties of various corrosive ions (Cl, HS, S2−, HCO3, and CO32−) on rutile TiO2(100) surfaces were studied.
The CASTEP in Material Studio, the first-principles computing software, was used to conduct geometric optimization for all adsorption configurations [13]. According to the setting requirements of the CASTEP module, a 2 × 3 × 1 three-dimensional supercell structure with periodic boundary conditions was established for the rutile TiO2(100) surface. In addition, a vacuum area with a thickness of 20 Å was added between the two plates to prevent interaction between them [14]. Figure 2 reveals the boundary surface models of various corrosive ions (Cl, HS, S2−, HCO3, and CO32−) at different adsorption sites (top, bridge, and hole) on a TiO2(100) surface.

2.1.2. Computing Method

Using the PBE functional of GGA, the pseudopotentials were constructed using the plane wave ultrasoft pseudopotential SCF [1,12,15], where the truncation energy of the plane wave was set as 400 eV, the convergence accuracy in the iteration process was 2 × 10−6 eV/atom, the self-consistent iteration was 300 times, the force converge was 0.03 eV/atom, the tolerance deviation was not higher than 0.005, the stress deviation was under 0.08 GPa, and the k-points value was 2 × 3 × 1 in the Brillouin zone.

2.2. Electrochemical Test

2.2.1. Preparation of Experimental Materials

The electrochemical test sample was TC4 titanium alloy, which was ø10 mm × 3 mm. A wire was welded to one end of the sample and tested for conductivity with a multimeter to verify whether the wire was welded correctly. The surface at the other end of the sample was the electrochemical test surface. The surface other than the electrochemical test surface was glued and stamped with epoxy resin AB glue and then polished with sandpaper with mesh sizes of 400#, 800#, 1200#, 1500#, and 2000#. For the purpose of reaching the test requirements for sample roughness, the sample surface was polished to 2000#, cleaned with distilled water, degreased with acetone, dehydrated and desiccated with alcohol, and dried with cold air for later use.

2.2.2. Experimental Methods and Equipment

The electrochemical test was carried out by Princeton P4000 electrochemical workstation, in which the working electrode was TC4 alloy, the reference electrode was polytetrafluoro silver chloride, and the auxiliary electrode was a platinum electrode. Before the electrochemical test, high-purity nitrogen was used to deoxygenate the required corrosive medium for 1 h., and the temperature was heated up to the preset temperature (80 °C). The electrochemical test was performed when the entire test system reached stability, and each experiment was performed three times.
The working electrode was pre-polarized at a voltage set value of −1200 mV for 3 min in advance of the electrochemical test. After the oxide film spontaneously took shape on the surface of the sample in the air and was eliminated, the working electrode was put in the prepared medium and stood for 30 min to form new film. The test frequency was set to 10−2 HZ~105 HZ, the measured signal amplitude was 10 mV sine wave, and the number of points was 50. The scanning rat was set as 0.3333 mV/s, and the potential was −1000 mV~+1600 mV.
The corrosive medium was 35 g/L NaCl, 35 g/L NaCl + 1 g/L Na2CO3, 35 g/L NaCl + 1 g/L Na2S, and 35 g/L NaCl + 1 g/L Na2CO3 + 1 g/L Na2S, respectively, which were all chemically pure agents.

3. Results and Discussion

3.1. Thermodynamic Stability of Passivation Film

3.1.1. Stable Adsorption Model

To simulate the species in the CO2-H2S-Cl environment (CO2+H2O→H2CO3, H2CO3→H++HCO3, HCO3→H++CO32−, H2S→H++HS, HS→H++S2−), the final energy of five corrosive ions (Cl, HS, S2−, HCO3, and CO32−) at different adsorption sites on TiO2(100) surface after geometric optimization is shown in Table 2. By comparison, it was found that the energy of each corrosive ion was the lowest at the bridge site of the TiO2(100) surface. If the energy of the adsorption system were more negative, its structure would be more stable [16]. Therefore, it can be determined that all of the best adsorption sites of Cl, HS, S2−, HCO3, and CO32− on the TiO2(100) surface were bridge sites. The final energy of each corrosive ion at the bridge site of the TiO2(100) surface was in the following order: S2− > HS > Cl > CO32− > HCO3.

3.1.2. Charge Density

Figure 3 reveals the charge density distribution of each corrosive ion (Cl, HS, S2−, HCO3, and CO32−) at the bridge site of the TiO2(100) surface. It could be seen that a forceful charge interaction exists between the Cl, S, O, and Ti atoms which was in the Cl, HS and S2−, HCO3, CO32−, and TiO2(100) surface, respectively.
Table 3 reveals the charge numbers of Cl, S, and O atoms in various corrosive ions. It could be seen that the absolute values are as follows: S2−(S) > CO32−(O) > Cl(Cl) > HS(S) > HCO3(O), which are in accordance with our previous research [2]. The metal surface with a higher charge density value is more likely to be corroded by the corrosive ions, leading to the TiO2 passivation film suffering from stronger corrosion. It could be seen that the stability of the TiO2 in the environment containing corrosive species was the following: S2− < CO32− < Cl < HS < HCO3. That is, TiO2 film on the surface of TC4 alloy is more easily damaged in the mediums containing S2− than in CO32−, Cl, HS, and HCO3.

3.1.3. Charge Density Difference

Figure 4 shows the charge density of the corrosive ions (Cl, HS, S2−, HCO3, and CO32−) at the bridge site of the TiO2(100) under the stable state adsorption. It could be seen that a very distinct charge transfer appearance was presented between Cl, S, O, and Ti atoms which was in the Cl, HS and S2−, HCO3 and CO32−, and TiO2(100) surfaces, respectively. Charge segregation and electronegativity decreased near Cl, S, and O atoms, while charge dissipation and electronegativity increased near the Ti atom in TiO2 [17]. Therefore, the interface binding energy between Cl, S, O, and Ti atoms was in the Cl, HS and S2−, HCO3 and CO32−, and TiO2(100) surfaces, respectively. Finally, the specific charge transfer process moved from the Ti atom on the TiO2(100) surface to Cl, S, and O atoms.

3.1.4. Density of States

Figure 5 shows PDOS (Projected Density of States) diagrams of five corrosive ions at the bridge site of the TiO2(100) surface, which can be calculated to investigate the characteristics of various ions on the TiO2(100) surface deeply [18]. It could be seen that a certain extent of the charge interaction existed between Cl, S, O, and Ti atoms, indicating that the adsorption process was chemical adsorption [19]. The charge interaction and interfacial bonding were primarily made of hybrid orbitals between 3d2 of the Ti atoms and 3p5 of Cl, 3p4 of S, and 2p4 of O.

3.1.5. Binding Energy

The corrosiveness of each corrosive ion to the matrix can be ensured through the interface binding energy, which was calculated as follows [20]:
E = E t - ( E 1 + E 2 )
Et is the total energy of whole model after geometry optimization; E1 is the energy after geometric optimization of TiO2(100); E2 is the energy after geometric optimization of each corrosive ion.
Table 4 displays the final energy between Cl, HS, S2−, HCO3, CO32−, and TiO2(100) after geometric optimization. According to Table 2 and Table 4, combined with Formula (1), the interface binding energies of various corrosive ions (Cl, HS, S2−, HCO3, and CO32−) at the bridge site of the TiO2(100) surface were obtained, as shown in Table 5. It could be seen that when HCO3 was adsorbed on the TiO2(100) surface, the entire adsorption system had low energy. Compared with Cl, HS, HCO3, and CO32−, the interface between S2− and TiO2(100) was easier to bond and react, indicating that S2− had a stronger adsorption capacity on TiO2. Therefore, TiO2 has poor stability in the environment containing S2−. The steadier the interface model is, the smaller interface binding energy is [20], so the film stability of TiO2 in the solutions containing Cl, HS, S2−, HCO3, and CO32− was S2− < CO32− < Cl < HS < HCO3, which is consistent with the charge density results mentioned above.

3.2. Corrosion Behavior

3.2.1. Alternating-Current Impedance

The alternating-current impedances of TC4 alloy in NaCl, NaCl + Na2CO3, NaCl + Na2S and NaCl + Na2S + Na2CO3 solutions containing saturated CO2 are shown in Figure 6. It could be seen that the radius of the capacitive arc of TC4 alloy in four corrosive solution was NaCl > NaCl + Na2CO3 > NaCl + Na2S > NaCl + Na2S + Na2CO3. The radius of electrochemical Nyquist impedance spectroscopy can determine the corrosion resistance of materials; the larger the radius of the electrochemical Nyquist impedance spectrum is, the stronger the corrosion resistance of materials to local corrosion is [21]. Therefore, the corrosiveness of four corrosive solutions to TC4 alloy was NaCl + Na2S + Na2CO3 > NaCl + Na2S > NaCl + Na2CO3 > NaCl.
The equivalent circuit was shown in Figure 7. It can be seen that Cdl (double layer capacitance) and Cc (ion adsorption double layer capacitance on the electrode surface) increased, and that both Rct (charge transfer resistance) and Rc (ion adsorption double layer resistance) decreased, concluding that the TC4 alloy has poor corrosion resistance [22].
As seen in Table 6, when there was only NaCl in the electrolyte, the Cc value of ion adsorption double layer capacitance on the electrode surface was 3.617 × 10−7, the Cdl value of double layer capacitance was 4.925 × 10−6, the Rc value was 1565 Ω·cm2, and the Rct value was 3.135 × 104 Ω·cm2. With the addition of CO32− and S2−, the values of Cc and Cdl increased to varying degrees, while the values of Rc and Rct decreased. When CO32− and S2− exited together, the corresponding electrochemical parameters increased. The corrosion resistance of the TC4 alloy to four solutions is NaCl > NaCl + Na2CO3 > NaCl + Na2S > NaCl + Na2S + Na2CO3, which is consistent with the above numerical simulation results.

3.2.2. Polarization Curve

Figure 8 displays the polarization curves of the TC4 titanium alloy in four corrosive media (NaCl, NaCl + Na2CO3, NaCl + Na2S, NaCl + Na2S + Na2CO3). Table 7 shows the fitting results. The icorr (self-corrosion current density) was 1.689 × 10−4 mA/cm2, and the Ecorr (self-corrosion potential) of TC4 alloy in NaCl solution containing saturated CO2 was −578 mV. With the addition of CO32− or/and S2−, the Ecorr of the electrode decreased, and icorr increased. The Ecorr can reflect the tendency of corrosion [23], and the icorr represents the speed of corrosion rate. The value of the icorr is larger, indicating that the corrosion rate is more rapid [24]. It could be seen that the TC4 titanium alloy showed excellent corrosion resistance in a corrosive solution containing only NaCl. In a NaCl + Na2CO3 solution, the resistance of the TC4 alloy decreased. While in the NaCl + Na2S + Na2CO3 solution, the TC4 alloy suffered from the most severe corrosion. This finding is consistent with the above alternating-current impedance results and numerical simulation results.
The results of the electrochemical experiments mentioned above also are in good accordance with the previous research in a 35% NaCl + 0.4% Na2S solution at 80 °C [25], as shown in Table 8.

4. Conclusions

(1)
All of the most suitable adsorption sites of corrosive ions (Cl, HS, S2−, HCO3, and CO32−) on the TiO2(100) surface were bridge sites, then hole sites and top sites.
(2)
A forceful charge interaction occurred between Cl, S, O, and Ti atoms. The charge was transferred from near the Ti atoms in the TiO2(100) surface to near Cl, Ss, and O atoms in Cl, HS, S2−, HCO3, and CO32−, respectively. Interface binding energy was primarily formed by electronic orbital hybridization between 3p5 of Cl, 3p4 of S, 2p4 of O, and 3d2 of Ti, and they were chemical adsorption.
(3)
Interface binding energy between five corrosive species and the TiO2(100) was as follows: S2− > CO32− > Cl > HS > HCO3.
(4)
With the addition of CO32− and S2−, local corrosion of the TC4 alloy in an NaCl solution containing saturated CO2 increased, especially the synergistic effect between Cl, CO32−, and/or S2−, which made the corrosion electrochemical parameters of TC4 alloy change by two orders of magnitude.

Author Contributions

Designing the experiments, S.Z.; performing the experiments, K.W. and P.D.; contributing the reagents, materials, and analysis tools, K.W. and H.M.; analyzing the data, P.D. and H.M.; writing—original draft preparation, P.D.; writing—review and editing, S.Z. All authors have discussed the results, reviewed the manuscript, and approved the decision to publish the results. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (51974245), Open Foundation of Shaanxi Key Laboratory of Carbon Dioxide Sequestration and Enhanced Oil Recovery (YJSYZX22SKF0003), and Key Research and Development Program of Shaanxi Province (2022GY-128, 2022SF-045).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

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.

References

  1. Peng, Z.G.; Lv, F.L.; Feng, Q.; Zheng, Y. Enhancing the CO2-H2S corrosion resistance of oil well cement with a modified epoxy resin. Constr. Build. Mater. 2022, 326, 126854. [Google Scholar] [CrossRef]
  2. Dong, P.; Zhang, Y.G.; Zhu, S.D.; Nie, Z.; Ma, H.X.; Liu, Q.; Li, J.L. 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. [Google Scholar] [CrossRef]
  3. Zhang, Z.; Zheng, Y.S.; Li, J.; Liu, W.Y.; Liu, M.Q.; Gao, W.X.; Shi, T.H. Stress corrosion crack evaluation of super 13Cr tubing in high-temperature and high-pressure gas wells. Eng. Fail. Anal. 2019, 95, 263–272. [Google Scholar] [CrossRef]
  4. Zheng, S.; Zhu, J.Y. Overview on plant extracts as green corrosion inhibitors in the oil and gas fields. J. Mater. Res. Technol. 2021, 15, 5078–5094. [Google Scholar]
  5. Muhammad, W.B.; Djukic, M. External corrosion of oil and gas pipelines: A review of failure mechanisms and predictive preventions. J. Nat. Gas. Sci. Eng. 2022, 100, 104467. [Google Scholar]
  6. Chen, L.Y.; Zhang, L.C. Corrosion of titanium alloys and composites in aqueous solutions. Encycl. Mater. Met. Alloy. 2022, 1, 200–211. [Google Scholar]
  7. dos Santos Monteiro, E.; de Souza Soares, F.M.; Fernandes Nunes, L.; Carvalho Santana, A.I.; de Biasi, R.S.; Nelson Elias, C. Comparison of the wettability and corrosion resistance of two biomedical Ti alloys free of toxic elements with those of the commercial ASTM F136 (Ti–6Al–4V) alloy. J. Mater. Res. Technol. 2020, 9, 16329–16338. [Google Scholar] [CrossRef]
  8. Qin, P.; Chen, L.Y.; Liu, Y.J.; Jia, Z.; Liang, S.X.; Zhao, C.H.; Sun, H.; Zhang, L.C. Corrosion and passivation behavior of laser powder bed fusion produced Ti-6Al-4V in static/dynamic NaCl solutions with different concentrations. Corros. Sci. 2021, 191, 109728. [Google Scholar] [CrossRef]
  9. 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]
  10. Patel, V.R.; Sonvane, Y.; Thakor, P.B. Structural and electrical properties of TiO2 monolayers using first-principle calculations. Mater. Today Process. 2021, 47, 563–566. [Google Scholar] [CrossRef]
  11. Alves, V.A.; Reis, R.Q.; Santos, I.C.; Souza, D.G.; Gonçalves, T.D.; Pereira-da-Silva, M.A.; Rossi, A.; Da Silva, L.A. In situ impedance spectroscopy study of the electrochemical corrosion of Ti and Ti–6Al–4V in simulated body fluid at 25 °C and 37 °C. Corros. Sci. 2009, 51, 2473–2482. [Google Scholar] [CrossRef]
  12. Yang, F.; Wen, L.Y.; Zhao, Y.; Xu, J.; Zhang, S.F.; Yang, Z.Q. Adsorption and reaction of both C and Cl2 on TiO2(100) surface based on the first principles calculation. J. Chin. Process Eng. 2020, 20, 569–575. [Google Scholar]
  13. Ding, J.J.; Yang, M.Y.; Chen, H.X.; Fu, H.W.; Peng, J.H. First-principles investigation on electrical and adsorption properties of the ZnO/Silicene heterostructures: The role of Ag and N co-doping and external electric field. FlatChem 2022, 33, 100369. [Google Scholar] [CrossRef]
  14. Hu, X.Y.; Gui, Y.G.; Zhu, S.P.; Chen, X.P. First-principles study of the adsorption behavior and sensing properties of C2H4 and C2H6 molecules on (CuO/TiO2) n (n= 1–3) cluster modified MoTe2 monolayer. Surf. Interfaces 2022, 31, 102003. [Google Scholar] [CrossRef]
  15. Li, S.H.; Cui, X.H.; Li, X.H.; Yan, H.T.; Zhang, R.Z.; Cui, H.L. Effect of coexistence of vacancy and strain on the electronic properties of NH3 adsorption on the Hf2CO2 MXene from first-principles calculations. Vacuum 2022, 196, 110774. [Google Scholar] [CrossRef]
  16. Wen, P.; Li, C.F.; Zhao, Y.; Zhang, F.C.; Tong, L.H. First principles calculation of occupancy; bonding characteristics and alloying effect of Cr, Mo, Ni in bulk α-Fe (C). Acta Phys. Sin. 2014, 63, 280. [Google Scholar]
  17. Jia, Y. First principle calculations on adsorption and diffusion behavior of Li on graphene surface. J. Chin. Mater. Rep. 2019, 33, 43–47. [Google Scholar]
  18. 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]
  19. Luo, H.; Wang, D.Y.; Liu, L.X.; Li, L.C. Theoretical study on the adsorption characteristics of trichlorophenol on TiO2(101) surface. J. Chin.At. Mol. Phy. 2020, 37, 349–353. [Google Scholar]
  20. Li, Y.Y. Study on stability of titanium alloy passivation film under severe corrosion environment. J. Chin. Xi’an Shiyou Univ. Nat. Sci. Ed. 2018, 33, 120–126. [Google Scholar]
  21. Li, J.C.; Chen, P.; Wang, Y.; Wang, G.Y. Corrosion resistance of surface texturing epoxy resin coatings reinforced with fly ash cenospheres and multiwalled carbon nanotubes. Prog. Org. Coat. 2021, 158, 106388. [Google Scholar] [CrossRef]
  22. Qiao, C.; Wang, M.N.; Hao, L.; Jiang, X.L.; Liu, X.H.; Thee, C.; An, X.Z. In-situ EIS study on the initial corrosion evolution behavior of SAC305 solder alloy covered with NaCl solution. J. Alloy. Compd. 2021, 852, 156953. [Google Scholar] [CrossRef]
  23. Li, X.; Dong, Y.C.; Dan, Z.H.; Chang, H.; Fang, Z.G.; Guo, Y.H. Corrosion behavior of ultrafine grained pure Ti processed by equal channel angular pressing. Acta Metall. Sin. China 2019, 55, 967–975. [Google Scholar]
  24. Xie, F.X.; He, X.B.; Cao, S.L.; Mei, M.; Qu, X.H. Influence of pore characteristics on microstructure, mechanical properties and corrosion resistance of selective laser sintered porous Ti–Mo alloys for biomedical applications. Electrochim. Acta 2013, 105, 121–129. [Google Scholar] [CrossRef]
  25. Chen, C.; Shuang, Y.H.; Chen, J.X.; Zhang, Y.Z.; Liu, B.S.; Zhang, S.H. Research on electrochemical characteristics and material selection of corrosion resistant titanium alloy in petroleum field. Light Met. China 2023, 1, 54–59. [Google Scholar]
Figure 1. Raman spectra of TiO2 passivation film on TC4 titanium alloy surface.
Figure 1. Raman spectra of TiO2 passivation film on TC4 titanium alloy surface.
Materials 16 03129 g001
Figure 2. Interface model of TiO2100) including corrosive ions. (a) TiO2(100)-Cl(top); (b) TiO2(100)-Cl(bridge); (c) TiO2(100)-Cl(hole); (d) TiO2(100)-HS(top); (e) TiO2(100)-HS(bridge); (f) TiO2(100)-HS(hole); (g) TiO2(100)-S2−(top); (h) TiO2(100)-S2−(bridge); (i) TiO2(100)-S2−(hole); (j) TiO2(100)-HCO3(top); (k) TiO2(100)-HCO3(bridge); (l) TiO2(100)-HCO3(hole); (m) TiO2(100)-CO32−(top); (n) TiO2(100)-CO32−(bridge); (o) TiO2(100)-CO32−(hole).
Figure 2. Interface model of TiO2100) including corrosive ions. (a) TiO2(100)-Cl(top); (b) TiO2(100)-Cl(bridge); (c) TiO2(100)-Cl(hole); (d) TiO2(100)-HS(top); (e) TiO2(100)-HS(bridge); (f) TiO2(100)-HS(hole); (g) TiO2(100)-S2−(top); (h) TiO2(100)-S2−(bridge); (i) TiO2(100)-S2−(hole); (j) TiO2(100)-HCO3(top); (k) TiO2(100)-HCO3(bridge); (l) TiO2(100)-HCO3(hole); (m) TiO2(100)-CO32−(top); (n) TiO2(100)-CO32−(bridge); (o) TiO2(100)-CO32−(hole).
Materials 16 03129 g002aMaterials 16 03129 g002b
Figure 3. Charge density distribution of five ions on the TiO2(100) surface. (a) Cl; (b) HS; (c) S2−; (d) HCO3; (e) CO32−.
Figure 3. Charge density distribution of five ions on the TiO2(100) surface. (a) Cl; (b) HS; (c) S2−; (d) HCO3; (e) CO32−.
Materials 16 03129 g003
Figure 4. Differential charge density distribution of five ions on the TiO2(100) surface at the bridge site. (a) Cl; (b) HS; (c) S2−; (d) HCO3; (e) CO32−.
Figure 4. Differential charge density distribution of five ions on the TiO2(100) surface at the bridge site. (a) Cl; (b) HS; (c) S2−; (d) HCO3; (e) CO32−.
Materials 16 03129 g004
Figure 5. PDOS curves of (a) Cl, (b) HS, (c) S2−, (d) HCO3, and (e) CO32− adsorbing on the TiO2(100) surface at the bridge site.
Figure 5. PDOS curves of (a) Cl, (b) HS, (c) S2−, (d) HCO3, and (e) CO32− adsorbing on the TiO2(100) surface at the bridge site.
Materials 16 03129 g005
Figure 6. Alternating-current impedance diagram of TC4 titanium alloy under different corrosion environments.
Figure 6. Alternating-current impedance diagram of TC4 titanium alloy under different corrosion environments.
Materials 16 03129 g006
Figure 7. Equivalent circuit diagram.
Figure 7. Equivalent circuit diagram.
Materials 16 03129 g007
Figure 8. Polarization curves of TC4 titanium alloy under four corrosion environments.
Figure 8. Polarization curves of TC4 titanium alloy under four corrosion environments.
Materials 16 03129 g008
Table 1. Raman spectra characteristic frequency shift position of three crystalline TiO2.
Table 1. Raman spectra characteristic frequency shift position of three crystalline TiO2.
Crystal Structure of TiO2Raman Frequency Shift (cm−1)
Brookite127, 150, 193, 212, 247, 286, 318, 366, 412, 462, 502, 544, 582, 645
Anatase143, 196, 326, 395, 512, 645
Rutile143, 244, 440, 610, 825
Table 2. Final energy of five corrosive ions on the TiO2(100) surface at different sites.
Table 2. Final energy of five corrosive ions on the TiO2(100) surface at different sites.
ModelsFinal Energy/eV
TiO2(top)-Cl−69,578.2607299930
TiO2(bridge)-Cl−69,578.6903762455
TiO2(hole)-Cl−69,578.6028239061
TiO2(top)-HS−69,462.8589004437
TiO2(bridge)-HS−69,463.9047289728
TiO2(hole)-HS−69,463.8965710229
TiO2(top)-S2−−69,443.5255285524
TiO2(bridge)-S2−−69,445.7491433063
TiO2(hole)-S2−−69,445.3972256848
TiO2(top)-HCO3−70,647.8573602164
TiO2(bridge)-HCO3−70,651.6819378591
TiO2(hole)-HCO3−70,649.4503215808
TiO2(top)-CO32−−70,633.5143850651
TiO2(bridge)-CO32−−70,633.5232150766
TiO2(hole)-CO32−−70,633.5217354935
Table 3. Charge numbers of Cl, O, and S atoms in corrosive ions.
Table 3. Charge numbers of Cl, O, and S atoms in corrosive ions.
AtomCl(Cl)HCO3(O)CO32−(O)HS(S)S2−(S)
Charge/e−0.15−0.12−0.22−0.14−0.34
Table 4. The final energy of five ions, TiO2, and TiO2(100).
Table 4. The final energy of five ions, TiO2, and TiO2(100).
ModelFinal Energy/eV
Cl−411.7437852366
HCO3−1483.4813928615
CO32−−1467.9443448936
HS−296.3156703881
S2−−280.4987150773
TiO2−4961.8922752454
TiO2(100)−69171.5619281723
Table 5. Interface binding energies of five ions on TiO2(100) surface at bridge sites.
Table 5. Interface binding energies of five ions on TiO2(100) surface at bridge sites.
ModelInterface Binding Energy/eV
TiO2(bridge)-Cl4.6153371666
TiO2(bridge)-HS3.9728695881
TiO2(bridge)-S2−6.3114999473
TiO2(bridge)-HCO33.3613831813
TiO2(bridge)-CO32−5.9830579936
Table 6. Alternating-current fitting results of TC4 titanium alloy in various corrosion environments.
Table 6. Alternating-current fitting results of TC4 titanium alloy in various corrosion environments.
Corrosion EnvironmentsRs
/Ω·cm2
Cc × 10−7
/F·cm−2
Rc
/Ω·cm2
Cdl × 10−6
/F·cm−2
Rct × 104
/Ω·cm2
NaCl68.12 ± 0.723.617 ± 0.0351565 ± 334.925 ± 0.0623.135 ± 0.039
NaCl + Na2CO354.38 ± 0.553.826 ± 0.0291324 ± 155.236 ± 0.0512.754 ± 0.058
NaCl + Na2S33.05 ± 0.614.182 ± 0.0421069 ± 2456.480 ± 0.0382.376 ± 0.066
NaCl + Na2S + Na2CO39.923 ± 0.588.793 ± 0.053948 ± 1656.641 ± 0.0420.633 ± 0.045
Table 7. Electrochemical parameters of TC4 titanium alloy in various corrosion environments.
Table 7. Electrochemical parameters of TC4 titanium alloy in various corrosion environments.
Corrosion MediaEcorr
/mV
bc
/mV
ba
/mV
icorr × 10−4
/mA·cm−2
NaCl−578 ± 3.5635.704 ± 0.08629.744 ± 0.0661.689 ± 0.034
NaCl + Na2CO3−595 ± 2.3533.951 ± 0.09241.011 ± 0.0859.399 ± 0.042
NaCl + Na2S−667 ± 4.6227.320 ± 0.07527.218 ± 0.07319.173 ± 0.071
NaCl + Na2S + Na2CO3−1014 ± 4.8522.659 ± 0.08832.778 ± 0.069252.519 ± 0.099
Table 8. Electrochemical parameters of titanium alloy in 35% NaCl + 0.4% Na2S [25].
Table 8. Electrochemical parameters of titanium alloy in 35% NaCl + 0.4% Na2S [25].
Temperature/°CMaterialsRs/Ω·cm2Rct/Ω·cm2Ecorr/Vicorr× 10−4
/mA·cm−2
25TC462.49947,740−0.5961.751
TC4ELI9.15290,050−0.5532.468
80TC415.5262,373−0.5465.651
TC4ELI6.1593,617−0.4024.105
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

Zhu, S.; Wang, K.; Ma, H.; Dong, P. Adsorption Characteristics between Ti Atoms of TiO2(100) and Corrosive Species of CO2-H2S-Cl System in Oil and Gas Fields. Materials 2023, 16, 3129. https://doi.org/10.3390/ma16083129

AMA Style

Zhu S, Wang K, Ma H, Dong P. Adsorption Characteristics between Ti Atoms of TiO2(100) and Corrosive Species of CO2-H2S-Cl System in Oil and Gas Fields. Materials. 2023; 16(8):3129. https://doi.org/10.3390/ma16083129

Chicago/Turabian Style

Zhu, Shidong, Ke Wang, Haixia Ma, and Pan Dong. 2023. "Adsorption Characteristics between Ti Atoms of TiO2(100) and Corrosive Species of CO2-H2S-Cl System in Oil and Gas Fields" Materials 16, no. 8: 3129. https://doi.org/10.3390/ma16083129

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