Next Article in Journal
Research Progress on Construction of Lutein-Loaded Nano Delivery System and Their Improvements on the Bioactivity
Next Article in Special Issue
Experimental Study on Fatigue Performance of Welded Hollow Spherical Joints Reinforced by CFRP
Previous Article in Journal
Control of Static and Dynamic Parameters by Fuzzy Controller to Optimize Friction Stir Spot Welding Strength
Previous Article in Special Issue
Corrosion Resistance of Mg/Al Vacuum Diffusion Layers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Efficient CuInSe2 Sensitized TiO2 Nanotube Films for Photocathodic Protection of 316 Stainless Steel

1
College of Mechanical and Electrical Engineering, Qingdao University, Qingdao 266071, China
2
State Key Laboratory of Bio-Fibers and Eco-Textiles, Qingdao University, Qingdao 266071, China
*
Author to whom correspondence should be addressed.
Coatings 2022, 12(10), 1448; https://doi.org/10.3390/coatings12101448
Submission received: 1 September 2022 / Revised: 25 September 2022 / Accepted: 27 September 2022 / Published: 30 September 2022

Abstract

:
CuInSe2 nanoparticles were successfully deposited on the surface of TiO2 nanotube arrays (NTAs) by a solvothermal method for the photocathodic protection (PCP) of metals. Compared with TiO2 NTAs, the CuInSe2/TiO2 composites exhibited stronger visible light absorption and higher photoelectric conversion efficiency. After 316 Stainless Steel (SS) was coupled with CuInSe2/TiO2, the potential of 316 SS could drop to −0.90 V. The photocurrent density of CuInSe2/TiO2 connected to 316 SS reached 140 μA cm−2, which was four times that of TiO2 NTAs. The composites exhibited a protective effect in the dark state for more than 8 h after 4 h of visible light illumination. The above could be attributed to increased visible light absorption, the extended lifetime of photogenerated electrons, and generation of oxygen vacancies.

Graphical Abstract

1. Introduction

316 SS is widely used for industrial applications because of its good corrosion resistance and excellent mechanical properties. Nevertheless, stainless steel (SS) is prone to pitting corrosion in Cl-rich solution [1,2,3]. Metal corrosion is extremely harmful, causing huge economic losses and even safety accidents every year [4,5]. Many anti-corrosion methods, including corrosion inhibitors [6], anticorrosion coatings [7] and cathodic protection [8], have been developed to inhibit the corrosion of steel. Photocathodic protection (PCP) is a new type of green anti-corrosion technology. Its principle is that when the semiconductor is irradiated by light which possesses higher energy than the band gap (Eg) of the semiconductor, electrons are excited from the valence band (VB) to the conduction band (CB); then, the photogenerated electrons are transferred from the semiconductor to the connected metal to reduce the potential of the metal surface, thereby inhibiting the metal corrosion. Compared with traditional cathodic protection techniques, PCP technology neither consumes energy nor releases metal ions into the environment [9,10,11]. Therefore, the development of PCP systems to protect metal against substrates is a promising approach for many industrial applications.
TiO2 has attracted great interest as a photoanode material for PCP application due to its stable physical and chemical properties, lack of toxicity, excellent photoelectric properties and low cost [12,13,14]. Unfortunately, some inherent defects of TiO2 limit its application. First of all, TiO2 is unable to utilize most sunlight (less than 5% of solar energy) due to its wide Eg [15]. Secondly, the photogenerated carriers in TiO2 are easy to recombine, which greatly reduces its photoelectric conversion efficiency [16], making it unable to protect metals in dark environments. Therefore, it is necessary to modify the TiO2, for example, through doping, with metal elements (W [17], Fe [18], Ni [19], etc.) or non-metal elements (N [20], B [21], etc.), and co-sensitizing with narrow gap semiconductors (MoS2 [22], Co(OH)2 [23], FeS2 [24], etc.).
Ternary semiconductors have aroused great attention because of their adjustable Eg and electronic energy level, controllable composition and internal structure [25]. Polymetallic sulfides/selenides have good electrochemical properties due to the synergistic effect of two metal atoms [26,27]. At present, AgInS2 and AgInSe2 have been used to modify TiO2 and have achieved good PCP effects [28,29]. In addition, selenides have faster heterogeneous electron transfer rates than sulfides because the electronegativity of Se is lower than S [30]. CuInSe2 is regarded as a promising photovoltaic material due to its adjustable band gap, high optical stability and excellent photoelectric conversion efficiency. Therefore, the sensitization of CuInSe2 may improve the composites’ efficiency of utilization of sunlight [31]. More importantly, the CB potential (ECB) of CuInSe2 was more negative than that of TiO2, so it is possible to construct a CuInSe2/TiO2 heterojunction, which can facilitate the efficient transfer of electrons from CuInSe2 to TiO2 [32]. Therefore, CuInSe2 may be an ideal semiconductor material for modifying TiO2. According to previous reports, some researchers have constructed CuInSe2/TiO2 nanostructures for photocatalytic degradation of organic pollutants [33] and solar cell applications [34]. However, there are no reports on the PCP performance of CuInSe2/TiO2 for metals.
In this paper, CuInSe2/TiO2 composites were synthesized by anodic oxidation and solvothermal methods. The morphologies, crystal structures, composition and light absorption properties of the composite materials were studied. The PCP performances of photoanodes were studied by electrochemical test methods. Through density functional theory (DFT) calculations, the electronic structure changes of composites were calculated. The PCP mechanism of CuInSe2/TiO2 on 316 SS was studied.

2. Materials and Methods

2.1. Chemicals and Reagents

The Ti foils (40 mm × 10 mm × 0.1 mm; 99.9% purity) were purchased from Shanghai Gao Dewei Co., Shanghai, China, and 316 SS was purchased from Shanghai Baosteel Co., Shanghai, China with the following ingredients (wt%): C 0.08%, Si 0.90%, P 0.045%, S 0.029%, Mn 1.80%, Ni 14.00%, Cr 17% and balanced Fe. CuCl2·2H2O (99.0%), InCl3·4H2O (99.9%), Se powder (99.0%), Na2S·9H2O (98%), NaOH (99%) and NaCl (99.5%) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China).

2.2. Fabrication of CuInSe2/TiO2 Photoelectrodes

Figure 1a illustrates the process for preparation of CuInSe2/TiO2 nanotube array (NTA) photoanodes. TiO2 NTAs were synthesized on a Ti foil by electrochemical anodization. The Ti foil was first chemically buffed in a mixed solution of 2.5 mL H2O, 0.45 g NH4F, 6.0 mL H2O2 and 6.0 mL HNO3, and then cleaned with first deionized water (DI water) and then absolute ethanol. The cleaned Ti foil and the Pt plate were used as the anode and cathode, respectively. The Ti foil was immersed in the electrolyte solution (0.36 g NH4F, 4.0 mL H2O, 60 mL ethylene glycol) for anodic oxidation at 30 V for 30 min, and then rinsed with, respectively, DI water and ethanol. The oxidized Ti foil was placed in a muffle furnace at 450 °C for 2 h, and then TiO2 NTAs were obtained.
CuInSe2 nanoparticles were synthesized on the TiO2 by a solvothermal method. CuCl2·2H2O (0.1, 0.2, 0.3 mmol), InCl3·4H2O (0.1, 0.2, 0.3 mmol) and Se powder (0.2, 0.4, 0.6 mmol) were mixed with 60 mL methanol and magnetically stirred for 20 min. The mixed solution and the as-fabricated TiO2 were immersed in a Teflon-lined autoclave and sintered at 200 °C for 12 h. Finally, CuInSe2/TiO2 composites were obtained. The synthesized photoanodes were marked, respectively, as CuInSe2/TiO2-A, CuInSe2/TiO2-B and CuInSe2/TiO2-C.

2.3. Characterization

The morphologies of the photoelectrodes were observed using a scanning electron microscope (SEM, Hitachi SU8220, Tokyo, Japan) with Quantax75 energy-dispersive X-ray spectroscopy (EDS, Hitachi SU8220, Tokyo, Japan). The crystal structures of the photoelectrodes were obtained using X-ray diffraction (XRD, D8-advance, Bruker AXS Co., Karlsruhe, Germany) the Cu Kα radiation. The chemical components and element chemical states were analyzed by X-ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific, Waltham, MA, USA, Al-Kα radiation, 1486.6 eV). The optical properties of the photoelectrode were tested using a UV-Vis diffuse reflectance spectrophotometer (DRS, Hitachi UH4150, Tokyo, Japan). Photoluminescence (PL, Ex: 320 nm) spectra were measured using an FLS980 series fluorescence spectrum scanner. Surface morphologies of 316 SS electrodes were obtained by a metallographic microscope (Axiocam 105 color, Oberkochen, Germany).

2.4. Photoelectrochemical Measurements

All measurements were carried out on an electrochemical workstation (CHI 760E, Chenhua Instrument Co., Ltd., Shanghai, China). The test device was composed of a corrosion cell (3.5 wt% NaCl) and a photoanode cell (0.1 M Na2S + 0.2 M NaOH). The two cells were connected by Nafion film. The 316 SS electrode (1 cm × 1 cm × 1 cm) was encapsulated in epoxy resin with an exposed area of 1 cm2. A 300 W xenon lamp (PLS-SXE 300C, Beijing Perfectlight Technology Co., Ltd., Beijing, China) with a 420 nm cut-off filter was used as a visible light source to illuminate the photoanode material vertically. The open circuit potentials (OCP), Tafel curves and electrochemical impedance spectroscopy (EIS) were recorded by the two-cell system (Figure 1b). The corrosion potential (Ecorr) and corrosion current density (Jcorr) were recorded by Tafel curves which were measured at a scanning rate of 0.5 mV·s–1 from −250 to 250 mV vs. OCP. The EIS data were obtained in the frequency range of 105–10–2 Hz with OCP as the initial potential and the amplitude of AC signal was 5 mV. The Mott–Schottky (M–S) curves were recorded from −1.0 V to 0.5 V with a frequency and an amplitude of 1000 Hz and 10 mV, respectively. The I–V and M–S tests were both carried out in 0.1 M Na2SO4 solution. The 316 SS in the corrosion cell and the photoanodes in the photoanode cell were connected by wires to the working electrode (WE). The saturated calomel electrode (SCE) and Pt plate were set as reference electrode (RE) and counter electrode (CE), respectively. The setup used to measure the photocurrent densities is shown in Figure 1c; CE and RE were connected with wires, 316 SS and photoanodes were connected to ground wire and WE, respectively.

3. Results and Discussion

3.1. Morphology and Chemical Compositions

The morphologies of TiO2 and CuInSe2/TiO2-B were studied by SEM. Figure 2a,b shows that the average inner diameter and tube length of TiO2 nanotubes were about 40 and 1600 nm, respectively. The ordered structure of TiO2 NTAs can promote the separation and transport of e/h+ pairs [35]. As shown in Figure 2c,d, CuInSe2 nanoparticles were successfully loaded on the TiO2 surface. It is evident that the deposition of the CuInSe2 nanoparticles had no effect on the morphology of the nanotubes. The elemental mapping results of CuInSe2/TiO2-B show that the composite material is composed of Ti, O, Cu, In and Se elements. It further indicates that CuInSe2 nanoparticles were formed uniformly on the surface of the nanotubes in the composite by the simple solvothermal method.
XRD spectra were used to determine the crystal structural information of the synthesized photoanodes. Figure 3 shows the XRD patterns of TiO2 and CuInSe2/TiO2-B. For TiO2, the XRD peaks at 25.3°, 49.5° and 53.9° corresponded to (101), (200) and (105) planes of anatase TiO2 (JCPDS No. 21-1272), respectively. The other peaks of TiO2 were derived from the Ti substrate. For CuInSe2/TiO2-B photoanode, the XRD peaks at 27.3°, 28.3°, 43.9° and 44.2° corresponded to (112), (103), (105) and (301) crystal planes of tetragonal chalcopyrite CuInSe2. No diffraction peak of the impurity phase was detected by XRD. Combined with the results of SEM and XRD, the CuInSe2/TiO2 photoanode materials with high purity were successfully prepared by anodic oxidation and solvothermal methods.
The elemental chemical state of CuInSe2/TiO2-B was determined by XPS spectra. Figure 4a shows that the dominant elements of the composite material were Se, In, Ti, O, and Cu, in addition to C element. This result is consistent with the EDS mapping results. The peaks at 458.6 and 464.3 eV were in agreement with Ti 2p3/2 and Ti 2p1/2, respectively, indicating that Ti exists in the form of Ti4+, which is derived from TiO2 [36] (Figure 4b). Figure 4c shows the peaks of 529.8 and 531.3 eV that were attributed to the lattice oxygen (OL) and adsorbed oxygen (OA) [37], respectively. The presence of OA indicates that oxygen vacancies were generated on the surface of the synthesized sample. Oxygen vacancy can facilitate electron transfer to the material surface, thus reducing the recombination of photogenerated carriers [38]. The binding energies of In 3d5/2 and 3d3/2 located at 444.5 and 452.4 eV, respectively (Figure 4d), corresponded to the binding energies of In3+ from CuInSe2 [39]. As shown in Figure 4e, the peak at 54.2 eV was indexed to Se 3d5/2, indicating the presence of Se2− in the composite, and no selenium oxide was formed [40]. Two main peaks with the binding energy peaks at 932.7 and 952.5 eV in Figure 4f were attributed to Cu 2p3/2 and Cu 2p1/2, respectively. The difference between the two binding energy peaks is 19.8 eV confirming Cu2+ was reduced to Cu+ during the process of reaction [41]. The above analysis further demonstrated the successful synthesis of CuInSe2/TiO2 photoanode composites.

3.2. Optical Properties Analysis

The light absorption performances of pure TiO2 and CuInSe2/TiO2-B were studied by UV-vis DRS. Figure 5a shows that pure TiO2 can only absorb UV light with a wavelength less than 370 nm. In addition, the light absorption of TiO2 in the range of 400–800 nm may be due to light scattering caused by cracks in TiO2 [42]. The absorption edge was o red-shifted to 520 nm after depositing CuInSe2 on the TiO2. This indicates that the sensitization of CuInSe2 nanoparticles improved the visible light absorption capacity of TiO2 NTAs.
The Eg of the photoanode materials can be calculated by the formula [43]:
( α h v ) 2 = A ( h v E g )
where α, h, v, and A stand for absorption coefficient, Planck’s constant, optical frequency, and characteristic constant, respectively. Figure 5b shows that the Eg of TiO2 and CuInSe2/TiO2 NTA composites were 3.2 and 2.5 eV, respectively.
The recombination rates of photogenerated carriers of CuInSe2/TiO2 and TiO2 NTAs were analyzed by PL spectra. Figure 5c shows that the peak intensity of TiO2 was larger than that of CuInSe2/TiO2, suggesting the photogenerated carrier recombination rate of CuInSe2/TiO2 decreases. The lifetime of the photogenerated carriers of was evaluated by time-resolved PL spectra (Figure 5d), which were used to characterize the lifetime of photogenerated carriers. The attenuation curve was fitted by the formula:
R ( t ) = B 1 e ( t / τ 1 ) + B 2 e ( t / τ 2 )
The emission-decay time-constant values associated with τ1, τ2, B1 and B2 can be calculated by the formula:
R ( t ) = ( B 1 τ 1 2 + B 2 τ 2 2 ) / ( B 1 τ 1 + B 2 τ 2 )
Table 1 displays the fitting and calculation results from Figure 5d. The mean electron lifetimes of TiO2 and CuInSe2/TiO2 were 1.26 and 1.32 ns, respectively. This indicated that the formation of the CuInSe2/TiO2 heterojunction inhibited the recombination of e/h+ pairs, thus enhancing the performance of PCP.

3.3. PCP Performance and Stability Evaluation

The PCP performances were evaluated by comparing the values of the photocurrent densities of 316 SS connected to different photoelectrode materials. As shown in Figure 6a, the photocurrent densities of all the photoanodes were almost zero before illumination, indicating that the photoanodes offered no protective effect for 316 SS in the dark state. When the light source was turned on, the photocurrent response of the composite material was considerably improved compared to TiO2. The composite materials prepared with different precursor concentrations also showed different photocurrent densities, and the CuInSe2/TiO2-B connected to 316 SS exhibited the largest photocurrent density (140 μA cm−2).
The potential change of the metal coupled with the photoelectrode is also a critical parameter for evaluating the properties of PCP. The more the potential drops, the better the performance of PCP [44]. In order to further study the PCP properties of photoelectrode materials, the OCP change curves of 316 SS connected with different photoelectrode materials were tested. As demonstrated in Figure 6b, the potential of 316 SS in 3.5 wt% NaCl was −0.19 V (vs. SCE). When the light source was turned on, the potentials of 316 SS connected with TiO2 and CuInSe2/TiO2 NTAs both shifted negatively and then tended to be stable. The potential drops of TiO2 under visible light irradiation was 0.29 V. For CuInSe2/TiO2 composites, the CuInSe2/TiO2-B showed the largest drop (0.71 V), which was consistent with the results of the photocurrent densities. With the increase in precursor concentration, the protective effect of the composite on 316 SS first increased and then decreased, which may be due to the excess of CuInSe2 blocking the pores of the TiO2 NTA and hindering the absorption of visible light. The potentials of the composite were still lower than those of 316 SS after the light source was closed, suggesting that the composite can also protect 316 SS for a period of time in the dark state.
Figure 7a shows the Tafel curves of 316 SS, 316 SS coupled with TiO2 and CuInSe2/TiO2-B NTAs in light and dark states. The fitting data of the Tafel curves are shown in Table 2. In the absence of light, the Ecorr of 316 SS connected with TiO2 and CuInSe2/TiO2-B negatively shifted to –0.44 and –0.47 V (vs. SCE), respectively, which may be because of the galvanic effect [20]. Under visible light irradiation, the negative shift of the Ecorr of 316 SS coupled with CuInSe2/TiO2-B was larger than that of TiO2, indicating that CuInSe2/TiO2-B has a better protection effect. In addition, the Jcorr of 316 SS connected with CuInSe2/TiO2-B was significantly increased compared to TiO2, which may be due to the increased electrochemical reaction rate at the interface caused by the polarization of photogenerated electrons [8].
The interfacial properties of CuInSe2/TiO2-B were investigated by EIS. Figure 7b,c displays the Nyquist plots and Bode-phase curves of pure 316 SS and 316 SS connected with different photoanodes under intermittent visible light. The impedance arc radius of 316 SS connected with CuInSe2/TiO2-B was smaller than that of TiO2 under light, which may be because more electrons were transferred from CuInSe2/TiO2-B to 316 SS, thus facilitating the electrochemical reaction rate of the interface [45]. In addition, the resistance arc radius of CuInSe2/TiO2-B NTAs coupled with 316 SS in the dark state was still smaller than that of 316 SS, indicating that CuInSe2/TiO2-B can also protect 316 SS in the dark state. Figure 7d shows the fitted equivalent circuit models from EIS data, where Rs represent solution resistance, Rp and Qdl represent polarization resistance and double-electric-layer capacitance, respectively, and Rf and Qf represent surface-film resistance and capacitance, respectively. The equivalent circuit model of bare 316 SS can be described as Rs(RpQdl). The equivalent circuit of the 316 SS coupled with different photoanode materials can be fitted as Rs(RpQdl)(RfQf). Table 3 shows the electrochemical parameters fitted from the equivalent circuits. The values of Rp can reflect the difficulty of corrosion [46,47,48]. The smaller value of Rp means more electrons were transferred to 316 SS. The Rp value of 316 SS was significantly reduced after 316 SS was connected to CuInSe2/TiO2-B under visible light, indicating that CuInSe2/TiO2-B had a higher separation efficiency of photoinduced carriers than TiO2.
Figure 8a shows the photoinduced I–V curves of CuInSe2/TiO2-B and TiO2 with visible light turned on and off. The photocurrent densities of the CuInSe2/TiO2-B NTA photoelectrode were higher than those of TiO2. This indicates that the heterojunction structure formed between CuInSe2 and TiO2 can increase the transfer rate of photoelectrons and promote the separation of photogenerated carriers.
Figure 8b–d displays the M–S plots of different photoanodes. The three curves all show a positive slope, suggesting that prepared photoanodes have the characteristics of an n-type semiconductor. The flat band potential (Efb) of a semiconductor can be estimated using C−2 = 0 in the M–S curve [49]. Figure 8b–d shows that the Efb of TiO2, CuInSe2 and CuInSe2/TiO2-B were −0.20, −0.68 and −0.41 V (vs. SCE), respectively. Obviously, the Efb of CuInSe2/TiO2-B was more negative than that of TiO2, indicating that the modification of CuInSe2 can promote charge transfer in TiO2 NTAs [50]. The slope of M–S plot is negatively correlated with the charge density [51]. The slope of the CuInSe2/TiO2-B curve was more negative that of TiO2, demonstrating that CuInSe2/TiO2-B had a higher free carrier density and superior photoelectrochemical performance than TiO2.
The stability of the photoanodes is important for their PCP applications. Therefore, a long-term OCP test was performed. As shown in Figure 9a, after the CuInSe2/TiO2-B was illuminated for 4 h, the potential of 316 SS coupled with CuInSe2/TiO2-B can be stabilized at −0.82 V. The potentials were still lower than the self-corrosion potential of 316 SS for more than 8 h in the dark state. This may be due to the extended lifetime of photogenerated electrons and the generation of oxygen vacancies. Figure 9b shows the XRD spectra of CuInSe2/TiO2-B before and after long-term OCP measurements. The XRD results showed that the crystal structure of the photoanode did not change after long-term testing. It indicated that CuInSe2/TiO2 composite had superior stability. The surface of the 316 SS before and after the experiment was characterized by metallographic microscopy. According to Figure 9c,e, the surface of 316 SS protected by CuInSe2/TiO2-B was consistent with that before the experiment, while several pitting holes appeared on the surface of the unprotected 316 SS. These demonstrated the protective performance of CuInSe2/TiO2 for 316 SS.

3.4. DFT Analysis and PCP Mechanism

The total electronic density of states (TDOS) and partial electronic density of states (PDOS) of the prepared CuInSe2/TiO2 NTAs and pure TiO2 NTAs were calculated using first-principles density functional theory (calculation methods were provided in the Supplementary Materials) In pure TiO2, the maximum value of valence band (VBM) is provided by the electrons of O 2p state electrons, and the minimum value of conduction band (CBM) is provided by the electrons of Ti 3d state electrons (Figure 10a). The Eg can be estimated by the difference between VBM and CBM. The calculated Eg (2.0 eV) of TiO2 is smaller than the DRS result (3.2 eV), which is due to the generalized gradient approximation (GGA) theory underestimating the Hubbard interaction [52]. In CuInSe2, the primary contribution of VBM is of Cu 3d electrons (Figure 10b), and CBM is mainly provided by Se 4p electrons. When CuInSe2 was deposited on the surface of TiO2 NTAs, the 3d electrons of Cu were hybridized with the 3d electrons of Ti and the 2p electrons of O at VB, which further improved the mobility of the photogenerated carriers. Compared with TiO2, after the formation of the CuInSe2/TiO2 heterostructure, the CBM of TiO2 shifted negatively, and the Eg decreased significantly, indicating the enhanced visible light absorption of the CuInSe2/TiO2 heterostructures. The DFT calculation results are consistent with the DRS results.
The possible mechanism for the improved protection of CuInSe2/TiO2 photoanodes was analyzed (Figure 11). The Efb of TiO2 and CuInSe2 obtained from the M–S curves were −0.20 and −0.68 V (vs. SCE), respectively, which are equal to 0.04 and −0.44 V vs. NHE, pH = 7, respectively. The ECB of an n-type semiconductor is 0.2 eV more negative than Efb [30], so the ECB of TiO2 and CuInSe2 were −0.16 and −0.64 eV, respectively. The EVB of TiO2 and CuInSe2 were 3.04 and 1.13 eV, respectively, obtained from the empirical formula EVB = Eg + ECB. Under visible light irradiation, photoelectrons migrated from the CB of CuInSe2 to the CB of TiO2, because the ECB of CuInSe2 is more negative than that of TiO2, and then to the surface of 316 SS. As a result, the potentials of 316 SS were lower than the self-corrosion potential, thereby inhibiting the anodic oxidation reaction of 316 SS. In addition, the photogenerated holes in the VB of TiO2 were transferred to the VB of CuInSe2 and consumed by reaction with the hole scavenger, thereby reducing the recombination of e/h+ pairs.

4. Conclusions

In this study, novel CuInSe2/TiO2 NTA photoanodes were successfully fabricated by electrochemical anodic oxidation and a solvothermal method. A highly efficient heterojunction was formed between tetragonal chalcopyrite CuInSe2 and anatase TiO2. The sensitization of CuInSe2 improved the absorption capacity of the composites to visible light, inhibited the recombination of electron-hole pairs and improved the electron transfer ability. The CuInSe2/TiO2-B NTA photoanode exhibited the best PCP performance. The photocurrent density of the composite connected to 316 SS could reach 140 μA cm−2 under visible light, and the potential drops to −0.90 V, which is much lower than the self-corrosion potential of 316 SS (−0.19 V). In addition, the protection effect can still be maintained for more than 8 h after 4 h of visible light irradiation. The optical images of the protected 316 SS fully demonstrated the excellent protection capability of CuInSe2/TiO2 NTAs. Therefore, CuInSe2/TiO2 NTAs show great potential application in the field of PCP.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/coatings12101448/s1. Supplementary Materials: Computational methods.

Author Contributions

Experiment, writing—original draft, review and editing, Z.Y.; writing—review and editing, supervision, funding acquisition, H.L.; experiment, picture drawing, X.C.; experiment, J.Z.; resources, Y.L.; resources, funding acquisition, P.Z.; resources, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by National Natural Science Foundation of China [grant number: 51801109] and Science and Technology Support Plan for Youth Innovation of Colleges in Shandong Province [grant number: DC2000000891].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Akhtar, S.; Matin, A.; Kumar, A.; Brahim, A.; Laoui, I.T. Enhancement of anticorrosion property of 304 stainless steel using silane coatings. Appl. Surf. Sci. 2018, 440, 1286–1297. [Google Scholar] [CrossRef]
  2. Zhang, X.; Chen, G.; Li, W.; Wu, D. Graphitic carbon nitride homojunction films for photocathodic protection of 316 stainless steel and Q235 carbon steel. J. Electroanal. Chem. 2020, 857, 113703. [Google Scholar] [CrossRef]
  3. Zhang, B.; Wang, J.; Wu, B.; Guo, X.; Wang, Y.; Chen, D.; Zhang, Y.; Du, K.; Oguzie, E.; Ma, X. Unmasking chloride attack on the passive film of metals. Nat. Commun. 2018, 9, 2559. [Google Scholar] [CrossRef] [PubMed]
  4. El Ibrahimi, B.; Jmiai, A.; Bazzi, L.; El Issami, S. Amino acids and their derivatives as corrosion inhibitors for metals and alloys. Arab. J. Chem. 2020, 13, 740–771. [Google Scholar] [CrossRef]
  5. Hou, B.; Li, X.; Ma, X.; Du, C.; Zhang, D.; Zheng, M.; Xu, W.; Lu, D.; Ma, F. The cost of corrosion in China. NPJ Mater. Degrad. 2017, 1, 4. [Google Scholar] [CrossRef]
  6. Tan, B.; He, J.; Zhang, S.; Xu, C.; Chen, S.; Liu, H.; Li, W. Insight into anti-corrosion nature of Betel leaves water extracts as the novel and eco-friendly inhibitors. J. Colloid Interface Sci. 2021, 585, 287–301. [Google Scholar] [CrossRef]
  7. Mirhashemihaghighi, S.; Swiatowska, J.; Maurice, V.; Seyeux, A.; Klein, L.; Salmi, E.; Ritala, M.; Marcus, P. Interfacial native oxide effects on the corrosion protection of copper coated with ALD alumina. Electrochim. Acta 2016, 193, 7–15. [Google Scholar] [CrossRef]
  8. Kear, G.; Barker, B.; Stokes, K.; Walsh, F. Corrosion and impressed current cathodic protection of copper-based materials using a bimetallic rotating cylinder electrode (BRCE). Corros Sci. 2005, 47, 1694–1705. [Google Scholar] [CrossRef]
  9. Delfani, F.; Rahbar, N.; Aghanajafi, C.; Heydari, A.; KhalesiDoost, A. Utilization of thermoelectric technology in converting waste heat into electrical power required by an impressed current cathodic protection system. Appl. Energy 2021, 302, 117561. [Google Scholar] [CrossRef]
  10. Prasad, N.; Pathak, A.; Kundu, S.; Mondal, K. Highly active and efficient hybrid sacrificial anodes based on high p pig iron, Zn and Mg. J. Electrochem. Soc. 2021, 168, 111504. [Google Scholar] [CrossRef]
  11. Yuan, J.; Tsujikawa, S. Characterization of Sol-Gel-Derived TiO2 Coatings and Their Photoeffects on Copper Substrates. J. Electrochem. Soc. 1995, 142, 3444–3450. [Google Scholar] [CrossRef]
  12. Shao, J.; Wang, X.; Xu, H.; Zhao, X.; Niu, J.; Zhang, Z.; Huang, Y.; Duan, J. Photoelectrochemical Performance of SnS2 Sensitized TiO2 Nanotube for Protection of 304 Stainless Steel. J. Electrochem. Soc. 2021, 168, 016511. [Google Scholar] [CrossRef]
  13. Chen, R.; Xu, Y.; Xie, X.; Li, C.; Zhu, W.; Xiang, Q.; Li, G.; Wu, D.; Li, X.; Wang, L. Synthesis of TiO2 nanotubes/nickel-gallium layered double hydroxide heterostructure for highly-efficient photocathodic anticorrosion of 304 stainless steel. Surf. Coat. Technol. 2021, 424, 127641. [Google Scholar] [CrossRef]
  14. Jiang, X.; Sun, M.; Chen, Z.; Jing, J.; Lu, G.; Feng, C. Boosted photoinduced cathodic protection performance of ZnIn2S4/TiO2 nanoflowerbush with efficient photoelectric conversion in NaCl solution. J. Alloy Compd. 2021, 876, 160144. [Google Scholar] [CrossRef]
  15. Sood, S.; Mehta, S.; Sinha, A.; Kansal, S. Bi2O3/TiO2 heterostructures: Synthesis, characterization and their application in solar light mediated photocatalyzed degradation of an antibiotic, ofloxacin. Chem. Eng. J. 2016, 290, 45–52. [Google Scholar] [CrossRef]
  16. Wang, M.; Sun, L.; Lin, Z.; Cai, J.; Xie, K.; Lin, C. p-n Heterojunction photoelectrodes composed of Cu2O-loaded TiO2 nanotube arrays with enhanced photoelectrochemical and photoelectrocatalytic activities. Energy Environ. Sci. 2013, 6, 1211–1220. [Google Scholar] [CrossRef]
  17. Momeni, M.; Khansari-Zadeh, S.; Farrokhpour, H. Fabrication of tungsten-iron-doped TiO2 nanotubes via anodization: New photoelectrodes for photoelectrochemical cathodic protection under visible light. SN Appl. Sci. 2019, 1, 1160. [Google Scholar] [CrossRef]
  18. Liu, Y.; Xu, C.; Feng, Z. Characteristics and anticorrosion performance of Fe-doped TiO2 films by liquid phase deposition method. Appl. Surf. Sci. 2014, 314, 392–399. [Google Scholar] [CrossRef]
  19. Sun, M.; Chen, Z.; Yu, J. Highly efficient visible light induced photoelectrochemical anticorrosion for 304 SS by Ni-doped TiO2. Electrochim. Acta. 2013, 109, 13–19. [Google Scholar] [CrossRef]
  20. Qiu, P.; Sun, X.; Lai, Y.; Gao, P.; Chen, C.; Ge, L. N-doped TiO2@TiO2 visible light active film with stable and efficient photocathodic protection performance. J. Electroanal. Chem. 2019, 844, 91–98. [Google Scholar] [CrossRef]
  21. Momeni, M.; Taghinejad, M.; Ghayeb, Y.; Bagheri, R.; Song, Z. Preparation of various boron-doped TiO2 nanostructures by in situ anodizing method and investigation of their photoelectrochemical and photocathodic protection properties. J. Iran. Chem. Soc. 2019, 16, 1839–1851. [Google Scholar] [CrossRef]
  22. Ma, X.; Ma, Z.; Lu, D.; Jiang, Q.; Li, L.; Liao, T.; Hou, B. Enhanced photoelectrochemical cathodic protection performance of MoS2/TiO2 nanocomposites for 304 stainless steel under visible light. J. Mater. Sci. Technol. 2021, 64, 21–28. [Google Scholar] [CrossRef]
  23. Lu, X.; Liu, L.; Ge, J.; Cui, Y.; Wang, F. Morphology controlled synthesis of Co(OH)2/TiO2 p-n heterojunction photoelectrodes for efficient photocathodic protection of 304 stainless steel. Appl. Surf. Sci. 2021, 537, 148002. [Google Scholar] [CrossRef]
  24. Wang, N.; Wang, J.; Liu, M.; Ge, C.; Hou, B.; Liu, N.; Ning, Y.; Hu, Y. Preparation of FeS2/TiO2 nanocomposite films and study on the performance of photoelectrochemistry cathodic protection. Sci. Rep. 2021, 11, 7509. [Google Scholar] [CrossRef] [PubMed]
  25. Aldakov, D.; Lefrançois, A.; Reiss, P. Ternary and quaternary metal chalcogenide nanocrystals: Synthesis, properties and applications. J. Mater. Chem. C. 2013, 1, 3756–3776. [Google Scholar] [CrossRef]
  26. Bhosale, R.; Agarkar, S.; Agrawal, I.; Naphade, R.; Ogale, S. Nanophase CulnS2 nanosheets/CuS composite grown by the SILAR method leads to high performance as a counter electrode in dye sensitized solar cells. Rsc. Adv. 2014, 4, 21989–21996. [Google Scholar] [CrossRef]
  27. Loo, A.; Bonanni, A.; Sofer, Z.; Pumera, M. Exfoliated transition metal dichalcogenides (MoS2, MoSe2, WS2, WSe2): An electrochemical impedance spectroscopic investigation. Electrochem. Commun. 2015, 50, 39–42. [Google Scholar] [CrossRef]
  28. Li, H.; Song, W.; Cui, X.; Li, Y.; Hou, B.; Zhang, X.; Wang, Y.; Cheng, L.; Zhang, P.; Li, J. AgInS2 and graphene co-sensitized TiO2 photoanodes for photocathodic protection of Q235 carbon steel under visible light. Nanotechnology 2020, 31, 305704. [Google Scholar] [CrossRef]
  29. Jiang, X.; Sun, M.; Chen, Z.; Jing, J.; Feng, C. High-efficiency photoelectrochemical cathodic protection performance of the TiO2/AgInSe2/In2Se3 multijunction nanosheet array. Corros. Sci. 2020, 176, 108901. [Google Scholar] [CrossRef]
  30. Zhao, X.; Yang, Y.; Li, Y.; Cui, X.; Zhang, Y.; Xiao, P. NiCo-selenide as a novel catalyst for water oxidation. J. Mater. Sci. 2016, 51, 3724–3734. [Google Scholar] [CrossRef]
  31. Jiang, Z.; Feng, L.; Zhu, J.; Liu, B.; Li, X.; Chen, Y.; Khan, S. Construction of a hierarchical NiFe2O4/CuInSe2 (p-n) heterojunction: Highly efficient visible-light-driven photocatalyst in the degradation of endocrine disruptors in an aqueous medium. Ceram. Int. 2021, 47, 8996–9007. [Google Scholar] [CrossRef]
  32. Wu, Z.; Tong, X.; Sheng, P.; Li, W.; Yin, X.; Zou, J.; Cai, Q. Fabrication of high-performance CuInSe2 nanocrystals-modified TiO2 NTs for photocatalytic degradation applications. Appl. Surf. Sci. 2015, 351, 309–315. [Google Scholar] [CrossRef]
  33. Liao, Y.; Zhang, H.; Zhong, Z.; Jia, L.; Bai, F.; Li, J.; Zhong, P.; Chen, H.; Zhang, J. Enhanced visible-photocatalytic activity of anodic TiO2 nanotubes film via decoration with CuInSe2 nanocrystals. ACS Appl. Mater. Interfaces 2013, 5, 11022–11028. [Google Scholar] [CrossRef]
  34. Yu, Y.; Chien, W.; Ko, Y.; Chen, S. Preparation and characterization of P3HT: CuInSe2: TiO2 thin film for hybrid solar cell applications. Thin Solid Film. 2011, 520, 1503–1510. [Google Scholar] [CrossRef]
  35. Zhang, T.; Rahman, Z.; Wei, N.; Liu, Y.; Liang, J.; Wang, D. In situ growth of single-crystal TiO2 nanorod arrays on Ti substrate: Controllable synthesis and photoelectro-chemical water splitting. Nano Res. 2017, 10, 1021–1032. [Google Scholar] [CrossRef]
  36. Deng, X.; Zhang, H.; Guo, R.; Cui, Y.; Ma, Q.; Zhang, X.; Cheng, X.; Li, B.; Xie, M.; Cheng, Q. Effect of fabricating parameters on photoelectrocatalytic performance of CeO2/TiO2 nanotube arrays photoelectrode. Sep. Purif. Technol. 2018, 193, 264–273. [Google Scholar] [CrossRef]
  37. Jiang, X.; Sun, M.; Chen, Z.; Jing, J.; Feng, C. An ultrafine hyperbranched CdS/TiO2 nanolawn photoanode with highly efficient photoelectrochemical performance. J. Alloy. Compd. 2020, 816, 152533. [Google Scholar] [CrossRef]
  38. Antony, R.; Mathews, T.; Dash, S.; Tyagi, A.; Raj, B. X-ray photoelectron spectroscopic studies of anodically synthesized self aligned TiO2 nanotube arrays and the effect of electrochemical parameters on tube morphology. Mater. Chem. Phys. 2012, 132, 957–966. [Google Scholar] [CrossRef]
  39. Anuroop, R.; Pradeep, B. Structural, optical, ac conductivity and dielectric relaxation studies of reactively evaporated In6Se7 thin films. J. Alloy. Compd. 2017, 702, 432–441. [Google Scholar] [CrossRef]
  40. Zhang, H.; Fang, W.; Wang, W.; Qian, N.; Ji, X. Highly Efficient Zn-Cu-In-Se Quantum Dot-Sensitized Solar Cells through Surface Capping with Ascorbic Acid. ACS Appl. Mater. Interfaces 2019, 11, 6927–6936. [Google Scholar] [CrossRef]
  41. Jia, H.; Cheng, S.; Zhang, H.; Yu, J.; Lai, Y. Band alignment at the Cu2SnS3/In2S3 interface measured by X-ray photoemission spectroscopy. Appl. Surf. Sci. 2015, 353, 414–418. [Google Scholar] [CrossRef]
  42. Dai, G.; Yu, J.; Liu, G. Synthesis and Enhanced Visible-Light Photoelectrocatalytic Activity of p-n Junction BiOI/TiO2 Nanotube Arrays. J. Phys. Chem. C 2011, 115, 7339–7346. [Google Scholar] [CrossRef]
  43. Tauc, J. Optical Properties and Electronic Structure of Amorphous Ge and Si. Mater. Res. Bull. 1968, 3, 37–46. [Google Scholar] [CrossRef]
  44. Lu, X.; Liu, L.; Xie, X.; Cui, Y.; Oguzie, E.; Wang, F. Synergetic effect of graphene and Co(OH)2 as cocatalysts of TiO2 nanotubes for enhanced photogenerated cathodic protection. J. Mater. Sci. Technol. 2020, 37, 55–63. [Google Scholar] [CrossRef]
  45. Sun, Z.; Li, F.; Zhao, M.; Xu, L.; Fang, S. A comparative study on photoelectrochemical performance of TiO2 photoanodes enhanced by different polyoxometalates. Electrochem. Commun. 2013, 30, 38–41. [Google Scholar] [CrossRef]
  46. Saha, S.; Dutta, A.; Ghosh, P.; Sukul, D.; Banerjee, P. Adsorption and corrosion inhibition effect of Schiff base molecules on the mild steel surface in 1 M HCl medium: A combined experimental and theoretical approach. Phys. Chem. Chem. Phys. 2015, 17, 5679–5690. [Google Scholar] [CrossRef]
  47. Solmaz, R.; Kardaş, G.; Çulha, M.; Yazıcı, B.; Erbil, M. Investigation of adsorption and inhibitive effect of 2-mercaptothiazoline on corrosion of mild steel in hydrochloric acid media. Electrochim. Acta 2008, 53, 5941–5952. [Google Scholar] [CrossRef]
  48. Özcan, M.; Dehri, İ.; Erbil, M. Organic sulphur-containing compounds as corrosion inhibitors for mild steel in acidic media: Correlation between inhibition efficiency and chemical structure. Appl. Surf. Sci. 2004, 236, 155–164. [Google Scholar] [CrossRef]
  49. Tong, X.; Shen, W.; Chen, X. Enhanced H2S sensing performance of cobalt doped free-standing TiO2 nanotube array film and theoretical simulation based on density functional theory. Appl. Surf. Sci. 2019, 469, 414–422. [Google Scholar] [CrossRef]
  50. Zhou, M.; Hou, Z.; Chen, X. Graphitic-C3N4 nanosheets: Synergistic effects of hydrogenation and n/n junctions for enhanced photocatalytic activities. Dalton Trans. 2017, 46, 10641–10649. [Google Scholar] [CrossRef]
  51. Wang, C.; Long, X.; Wei, S.; Wang, T.; Li, F.; Gao, L.; Hu, Y.; Li, S.; Jin, J. Conformally Coupling CoAl-LDH on Fluorine-Doped Hematite: Surface and Bulk Co-modification for Enhanced Photoelectrochemical Water Oxidation. ACS Appl. Mater. Interfaces 2019, 11, 29799–29806. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, X.; Lei, J.; Shao, Q.; Li, X.; Ning, X.; Shao, J.; Duan, J.; Hou, B. Preparation of ZnWO4/TiO2 composite film and its photocathodic protection for 304 stainless steel under visible light. Nanotechnology 2019, 30, 045710. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of (a) the synthesis process for the CuInSe2/TiO2 NTAs. Test devices for measuring (b) OCP, Tafel and EIS, and (c) photocurrent densities.
Figure 1. Schematic illustration of (a) the synthesis process for the CuInSe2/TiO2 NTAs. Test devices for measuring (b) OCP, Tafel and EIS, and (c) photocurrent densities.
Coatings 12 01448 g001
Figure 2. (a) SEM Top-view and (b) cross-section of the TiO2 NTAs, (c) SEM top-view, (d) cross-section view of the CuInSe2/TiO2-B NTAs, and (e) EDS elemental mapping of the CuInSe2/TiO2-B NTAs.
Figure 2. (a) SEM Top-view and (b) cross-section of the TiO2 NTAs, (c) SEM top-view, (d) cross-section view of the CuInSe2/TiO2-B NTAs, and (e) EDS elemental mapping of the CuInSe2/TiO2-B NTAs.
Coatings 12 01448 g002
Figure 3. XRD patterns of TiO2 and CuInSe2/TiO2-B.
Figure 3. XRD patterns of TiO2 and CuInSe2/TiO2-B.
Coatings 12 01448 g003
Figure 4. (a) The total XPS survey spectrum; high-resolution spectra of (b) Ti 2p, (c) O 1s, (d) In 3d, (e) Se 3d and (f) Cu 2p of CuInSe2/TiO2-B NTAs.
Figure 4. (a) The total XPS survey spectrum; high-resolution spectra of (b) Ti 2p, (c) O 1s, (d) In 3d, (e) Se 3d and (f) Cu 2p of CuInSe2/TiO2-B NTAs.
Coatings 12 01448 g004
Figure 5. (a) UV-vis DRS, (b) Tauc plots, (c) PL spectra and (d) time-resolved PL spectra of the prepared TiO2 and CuInSe2/TiO2-B.
Figure 5. (a) UV-vis DRS, (b) Tauc plots, (c) PL spectra and (d) time-resolved PL spectra of the prepared TiO2 and CuInSe2/TiO2-B.
Coatings 12 01448 g005
Figure 6. (a) I–t curves and (b) OCP–t curves between the different materials and 316 SS.
Figure 6. (a) I–t curves and (b) OCP–t curves between the different materials and 316 SS.
Coatings 12 01448 g006
Figure 7. (a) Tafel curves, (b) Nyquist plots, and (c) Bode-phase curves of pure 316 SS and 316 SS coupled with different photoanodes under intermittent visible light; (d) the equivalent circuit for fitting the impedance data.
Figure 7. (a) Tafel curves, (b) Nyquist plots, and (c) Bode-phase curves of pure 316 SS and 316 SS coupled with different photoanodes under intermittent visible light; (d) the equivalent circuit for fitting the impedance data.
Coatings 12 01448 g007
Figure 8. (a) I–V curves of the prepared TiO2 and CuInSe2/TiO2-B; (bd) M–S plots of TiO2, CuInSe2 and CuInSe2/TiO2-B.
Figure 8. (a) I–V curves of the prepared TiO2 and CuInSe2/TiO2-B; (bd) M–S plots of TiO2, CuInSe2 and CuInSe2/TiO2-B.
Coatings 12 01448 g008
Figure 9. (a) Long-term OCP change of 316 SS connected to the CuInSe2/TiO2-B photoanode under on and off visible light illumination; (b) XRD images of CuInSe2/TiO2-B before and after long-term OCP measurements; the optical images of the 316 SS (c) before the experiment, (d) unprotected and (e) protected by CuInSe2/TiO2-B for 14 h.
Figure 9. (a) Long-term OCP change of 316 SS connected to the CuInSe2/TiO2-B photoanode under on and off visible light illumination; (b) XRD images of CuInSe2/TiO2-B before and after long-term OCP measurements; the optical images of the 316 SS (c) before the experiment, (d) unprotected and (e) protected by CuInSe2/TiO2-B for 14 h.
Coatings 12 01448 g009
Figure 10. The calculated TDOS and PDOS for (a) TiO2, (b) CuInSe2 and (c) CuInSe2/TiO2.
Figure 10. The calculated TDOS and PDOS for (a) TiO2, (b) CuInSe2 and (c) CuInSe2/TiO2.
Coatings 12 01448 g010
Figure 11. Possible electron-transfer mechanism of the CuInSe2/TiO2.
Figure 11. Possible electron-transfer mechanism of the CuInSe2/TiO2.
Coatings 12 01448 g011
Table 1. Fitting results from Figure 5d.
Table 1. Fitting results from Figure 5d.
Samplesτ1B1τ2B2t
TiO21.26202.541.26245.591.26
CuInSe2/TiO21.32308.751.32362.611.32
Table 2. Electrochemical parameters obtained by Figure 7a.
Table 2. Electrochemical parameters obtained by Figure 7a.
SamplesEcorr (V vs. SCE)Jcorr (μA cm−2)
316 SS−0.191.58
TiO2 dark−0.445.22
CuInSe2/TiO2-B dark−0.475.31
TiO2 illumination−0.5932.31
CuInSe2/TiO2-B illumination−0.7676.26
Table 3. Electrochemical impedance parameters of the as-prepared photoanodes obtained from Figure 7b.
Table 3. Electrochemical impedance parameters of the as-prepared photoanodes obtained from Figure 7b.
SamplesRs (Ω·cm2)QfRf (Ω·cm2)QdlRp (Ω·cm2)
Y01 (Sn·Ω−1 cm−2)n1Y02 (Sn·Ω−1 cm−2)n2
3165.559---2.186 × 10−50.92321.001 × 105
TiO2 a7.5584.984 × 10−41.0022.1903.232 × 10−40.77964.357 × 104
CuInSe2/TiO2-B a4.8691.912 × 10−30.80271.948 × 1036.390 × 10−40.83692.840 × 104
TiO2 b3.5672.879 × 10−70.993.6366.784 × 10−40.72433.783 × 103
CuInSe2/TiO2-B b6.6773.167 × 10−30.89161.616 × 1039.438 × 10−40.795820.990
a Dark. b Visible light illumination.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, Z.; Li, H.; Cui, X.; Zhu, J.; Li, Y.; Zhang, P.; Li, J. Highly Efficient CuInSe2 Sensitized TiO2 Nanotube Films for Photocathodic Protection of 316 Stainless Steel. Coatings 2022, 12, 1448. https://doi.org/10.3390/coatings12101448

AMA Style

Yang Z, Li H, Cui X, Zhu J, Li Y, Zhang P, Li J. Highly Efficient CuInSe2 Sensitized TiO2 Nanotube Films for Photocathodic Protection of 316 Stainless Steel. Coatings. 2022; 12(10):1448. https://doi.org/10.3390/coatings12101448

Chicago/Turabian Style

Yang, Zhanyuan, Hong Li, Xingqiang Cui, Jinke Zhu, Yanhui Li, Pengfei Zhang, and Junru Li. 2022. "Highly Efficient CuInSe2 Sensitized TiO2 Nanotube Films for Photocathodic Protection of 316 Stainless Steel" Coatings 12, no. 10: 1448. https://doi.org/10.3390/coatings12101448

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