Next Article in Journal
Scientific Research on a Gold- and Silver-Inlaid Bronze Zun from the Han Dynasty
Previous Article in Journal
Effect of Microcapsules on Mechanical, Optical, Self-Healing and Electromagnetic Wave Absorption in Waterborne Wood Paint Coatings
Previous Article in Special Issue
Effect of Fabric Electrode Surface Coating Medium on ECG Signal Quality under Dynamic and Static Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Efficient Photocathodic Protection Performance of ZIS@CNNs Composites under Visible Light

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 2023, 13(9), 1479; https://doi.org/10.3390/coatings13091479
Submission received: 30 June 2023 / Revised: 4 August 2023 / Accepted: 15 August 2023 / Published: 22 August 2023

Abstract

:
Low isolation efficiency of photogenerated electron-hole pairs and inadequate utilization of visible light limit the application of g-C3N4 nanosheets (CNNs) in photocathodic protection (PCP). Therefore, indium zinc sulfide (ZnIn2S4, ZIS) nanolayers with nano-leaf structures were fabricated on CNNs using a simple hydrothermal method and used as visible light sensitizer and electron donor to improve its PCP performance. Under visible light illumination, the 30% ZIS@CNNs photoelectrode coupled with 316 stainless steel (SS) exhibited the largest photocurrent density of 17.30 μA cm−2 and the highest potential drop of 0.37 V, which was approximately 4 and 7.5 times higher than that of pure CNNs, respectively. The improvement in protection performance may be attributable to the crucial increase in visible light absorption and the terrific enhancement in rapid migration pathways provided using heterogeneous junctions.

1. Introduction

Photocathodic protection (PCP) technology has attracted extensive attention because it has been a promising green and clean technology in the past decades [1,2,3,4]. This technology takes advantage of the photosensitive characteristics of semiconductor photoanode materials to protect metals from corrosion by transferring electrons to the metal surface and decreasing the metal’s potential.
As a metal-free semiconductor, g-C3N4 (CN, ~2.7 eV) has wide applications such as photocatalytic hydrogen production [5,6,7], degradation of organic pollutants [8,9,10], and CO2 reduction, etc. [11,12] due to the advantages of visible light response, obvious chemical stability, easy preparation, rich reserves, and strong environmental affinity. Compared with metal-semiconductor, CN has incomparable advantages in avoiding secondary pollution [13]. However, the protection of metals by CN materials may not be ideal due to many defects, such as limited surface area and inefficient utilization of visible light [14,15,16,17]. In order to address these issues, various approaches, including nanostructure design [18,19], elemental doping [20,21], and heterojunction construction [22,23], have been applied to modify CN. Among these approaches, heterojunction construction is recognized as a more efficient approach to suppress the carrier recombination and improve the absorption efficiency of visible light, such as NaNbO3/CN [24], KBiFe2O5/CN [25], Co3O4/CN heterojunction [26].
As a unique ternary sulfide n-type semiconductor, ZnIn2S4 (ZIS, 2.34~2.55 eV) has been extensively applied in the photocatalytic fields due to its exceptional layered structure, tunable band gap, and excellent visible light absorption properties [27,28,29]. Compared with conventional binary metal sulfides, ZIS has higher photo-corrosion resistance and better chemical stability, which has aroused substantial attention of researchers in the field of PCP [30]. However, ZIS still has some limitations in its applications, including poor photoelectron mobility and low carrier separation efficiency [31,32,33]. Coupling CN with ZIS to construct heterojunction composites is a feasible method to solve the above limitations. Dang et al. reported that CN/ZIS composite showed better photocatalytic hydrogen production activity than pure ZIS and CN [34]. Hou et al. reported that CN/ZIS composite exhibited good photocatalytic degradation of ofloxacin performance in comparison with pure CN and ZIS [35]. However, no research on CN/ZIS composites in the field of PCP has been reported. Therefore, the PCP performances of the CN/ZIS composites were investigated, and the PCP mechanism was studied. In addition, based on the reported literature [34,36,37] and our previous work, we found that less than 50% of ZIS in composites yielded better results. Therefore, we intend to explore the best of the following three proportions (20%, 30%, and 40%), which may provide some guidance for exploring the best proportions in the future.
The morphological structure and size of semiconductors also affect the properties of the material. The ultrathin CN nanosheets (CNNs) are regarded as an ideal structure that can significantly enlarge the specific surface area and expose a considerable number of active sites, thus effectively improving the performance of CN [38,39]. Moreover, it has been reported that the 2D morphology in the heterojunction composites can significantly facilitate electron migration and boost the absorption of visible light [40], which may improve the utilization efficiency of photogenerated carriers and the PCP performance.
In this work, we fabricated the ZIS@CNNs composites using two-step calcination and hydrothermal method and then coated the composites onto the fluorine-doped tin oxide (FTO) substrates. The microstructures, chemical compositions, and crystalline structures of CNNs and different ZIS@CNNs composites were identified using scanning electron microscope (SEM), transmission electron microscopy (TEM), X-ray photoelectron spectroscopy (XPS), and X-ray diffraction (XRD). UV-Vis diffuse reflectance spectroscopy (DRS) was used to analyze optical absorption properties. Photoluminescence (PL) spectroscopy was used to analyze the isolation efficiency of carriers. The PCP performances of the composites were evaluated using photoinduced open circuit potential (OCP) and photocurrent densities. The PCP mechanism was proposed using analyses of electrochemical impedance spectroscopy (EIS) and Mott-Schottky (M-S) spectra.

2. Experimental

2.1. Materials

Chemicals:melamine (C3N3(NH2)3)), Zinc trinitrate hexahydrate (Zn(NO3)2·6H2O), thioacetamide (TAA), indium nitrate tetrahydrate (In(NO3)3∙4H2O), sodium citrate dihydrate (C6H5O7Na3·2H2O) were all acquired from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). FTO glasses (7 Ω cm−2, 10 × 20 × 2.2 mm3) were purchased from Guluo Glass Co., Ltd. (Luoyang, China).

2.2. Preparation of CNNs

CNNs were fabricated by twice calcination (Figure 1). 5 g melamine was heated from room temperature to 520 °C at a heating rate of 5 °C min−1 and maintained at this temperature for 4 h in a tube oven. Afterward, the resultant sample was kept calcining (heating rate: 2.5 °C min−1) at 550 °C for 2 h. The light yellow CNNs were collected (Figure S1).
ZIS nano-leaves were fabricated onto CNNs by a simple hydrothermal method. Firstly, precursor solutions were obtained by adding 0.1 mmol Zn(NO3)2∙6H2O, 0.2 mmol In(NO3)3∙4H2O, and 0.6 mmol sodium citrate dihydrate to the beaker filled with 20 mL deionized water and stirring vigorously for 0.5 h at room temperature. Secondly, different weights of CNNs were added to the precursor solution. The above solution was stirred for 3 h to obtain a white suspension. Finally, 0.8 mmol TAA was poured into the above solution, ultrasound for a while, and churned for 30 min. After that, the above suspension was transferred into a Teflon-lined stainless-steel autoclave (KH-30, Shenghua Instrument Technology Co., Suzhou, China) and heated for 1 h under 160 °C. The fabricated materials were washed with deionized water and anhydrous ethanol, respectively, then collected by centrifugation and dried at 60 °C for 24 h. ZIS@CNNs composites (Figure S1) with different proportions of ZIS were obtained and marked as 20% ZIS@CNNs, 30% ZIS@CNNs, and 40% ZIS@CNNs composites, respectively. As a reference, pure ZIS nano-leaves were also synthesized under similar conditions without CNNs.

2.3. Fabrication of ZIS@CNNs Photoelectrodes

The treated FTO was sealed with tape, and only a 10 × 15 mm2 conductive surface was left as the working surface. 20 mg ZIS@CNNs composite was added to a mixed solution containing 10 μL naphthol and 1 mL anhydrous ethanol and sonicated for 15 min. A uniform suspension was obtained. 40 μL of the above solution was applied drop by drop to the conductive surface of the FTO so that the suspension was distributed as uniformly as possible on the surface of the FTO (about 0.528 mg cm−2). Then, the FTO was dried in the air atmosphere and transferred into an oven to heat at 100 °C for 2 h to obtain ZIS@CNNs photoelectrodes.

2.4. Characterization

The crystal structures of CNNs, ZIS, and ZIS@CNNs were detected by XRD diffractometer with Cu Kα radiation (Rigaku Ultima IV, Tokyo, Japan). The elemental components and keying messages for different ZIS@CNNs composites were analyzed using XPS (Thermo ESCALAB 250XI Al Kα as the X-rays, Tampa, FL, USA). The microstructure of the composite was identified by HRTEM (JEM 200CX, Fukuoka, Japan). The absorption ability of materials was tested by a UV-Vis diffuse reflectance spectrometer (Shimadzu UV-3600i Plus, Kyoto, Japan) equipped with BaSO4 plates used as a reference. Photoluminescence (PL, EI FLS-980, Livingston, EH, USA) spectrum was used to assess the removal efficiency of photogenerated carriers for a variety of materials with an excitation wavelength of 320 nm. The Fourier transform infrared spectroscopy (FT-IR, Thermo Scientific Nicolet iS20, Watertown, MA, USA) was tested at the wavelength range from 400 to 4000 cm−1.

2.5. Electrochemical Tests

An electrochemistry workstation (CHI760E, Chenhua Instrument Co., Ltd., Shanghai, China) was chosen to investigate the electrochemistry properties of the fabricated photoelectrodes. The H-type electrochemical cell was selected for electrochemical performance testing, which was filled with a 3.5 wt.% NaCl solution and a mixture solution of 0.1 M Na2S and 0.2 M NaOH in the corrosion chamber and photochemical chamber, respectively. Conductivity between electrolyte solutions in two electrolytic cells was performed using the Nafion membrane (N-117, DuPont Co., Wilmington, DE, USA). The ZIS@CNNs photoanode was installed in the photochemical chamber, whereas the protected metal electrode with an active area of 1 cm2 was placed in the corrosion chamber as the test subject. A 300 W Xe lamp (PLS-SXE300, Perfect-light Co., Ltd. Beijing, China) was selected as the illumination source (Figure S2a). The light illuminated the photoelectrode through a crystal window with a diameter of 30 mm on the side of the photochemical chamber. A 420 nm filter was used to filter out light below 420 nm. During the measurements of OCP, EIS, and Tafel, the photoanode and 316 SS were connected by wires as working electrodes (WE). The Pt sheet and saturated calomel electrode (SCE) were regarded as counter electrode (CE) and reference electrode (RE), respectively. All electrochemical and photocatalytic experiments were carried out without stirring. The EIS tests were conducted under light and performed at a frequency range of 10−2~105 Hz, with an alternating current amplitude of 5 mV. The Tafel test was conducted under light and performed at a range of −0.25~0.25 V (vs. OCP). During the measurement of photocurrent densities, connecting the photoanode and 316 SS with WE and ground electrode, respectively, and RE and CE were connected in a short circuit (Figure S2b). The M-S tests were conducted under the light and measured with a frequency of 1000 Hz. A conventional three-electrode system with 0.1 mol L−1 Na2SO4 solution was chosen as the electrolyte solution. The photoelectrode, Pt sheet, and SCE were regarded as the WE, CE, and RE, respectively.

3. Results and Discussion

3.1. Characterization of ZIS@CNNs Composites

Figure 2a–c show the morphological images of CNNs, ZIS, and ZIS@CNNs composites, respectively. It could be observed from Figure 2a that pure CNNs prepared via the two-step calcination method had distinct flaky images and exhibited overlapping and stacking morphologies. As can be seen from Figure 2b, the morphology of pure ZIS synthesized by the hydrothermal method resembles nano-leaves. As shown in Figure 2c, the bottom layer of the ZIS@CNNs composite material is CNNs, and ZIS nano-leaves are evenly attached to its surface. These results indicate that ZIS nano-leaves are successfully grown on CNNs. Figure 2d shows the chemical composition of the ZIS@CNNs composites obtained by EDS testing, which shows the presence of Zn, In, S, C, and N elements.
When CNNs were added to the ZIS precursor solution with dissolved sodium citrate dihydrate, Zn2+ and In3+ in the solution moved rapidly to CNNs, and then the added TAA released S2−, which reacted with Zn2+ and In3+ attached to the surface to form ZIS nanoparticles. Under hydrothermal conditions, sodium citrate dihydrate exerted an inhibitory effect on the growth of nanoparticles, and then the nanoparticles gradually became nano-leaves. Eventually, a large number of ZIS nano-leaves accumulated on the surface of CNNs, forming ZIS@CNNs heterojunctions.
The microscopic morphologies of CNNs, ZIS, and 30% ZIS@CNNs were further investigated using TEM. From Figure 3a,b, it was clear that the morphologies of CNNs and ZIS were nanosheets and nano-leaves, respectively. 30% of the ZIS@CNNs composite’s surface is rougher than that of pure CNNs due to the deposition of the ZIS nano-leaves, which is in accordance with the SEM results. Figure 3d displayed the HRTEM image of 30% ZIS@CNNs. The crystal spacing of 0.321 nm is consistent with the (002) crystallographic plane of CNNs, and the crystal spacing of 0.315 nm is consistent with the (104) crystallographic plane of ZIS (JCPDS No. 49-1562).
Figure 4a,b shows the N2 adsorption-desorption curves and pore size distributions isotherms of CNNs and 30% ZIS@CNNs composite, respectively. According to the adsorption isotherm, the SBET of CNNs was 656.965 m2/g, while the SBET of 30% ZIS@CNNs composite decreased sharply to 47.239 m2/g. The pore size distribution of 30% ZIS@CNNs composite and CNNs were analyzed according to the Barrett-Joyner-Halenda (BJH) approach [41], The average pore diameters of 30% ZIS@CNNs composite and CNNs are about 13.049 and 4.380 nm, respectively, indicating that both materials are mesoporous structures.
Figure 5a display the XRD patterns of pure CNNs, pure ZIS, and three different ZIS@CNNs composites. It could be obviously seen from Figure 5a that there were two evident characteristic peaks in the XRD patterns of CNNs. The weak peak at 12.9° can be assigned to an in-plane repeating motif of the continuous heptazine network [42,43]. Due to the interlayer-stacking (002) crystal plane of melon networks, there is a prominent and sharp diffraction peak at 27.7° [44]. From the XRD patterns of ZIS, the peaks at 28.3°, 47.0°, and 55.8° can be assigned to the diffraction peaks of (104), (110), and (024) hexagonal crystal planes of ZIS (JCPDS No. 49-1562), respectively. For the three different ZIS@CNNs composites, the peak of the (002) crystal plane of CNNs overlaps with the peak of the (104) hexagonal crystal plane of ZIS, and the remaining two peaks are the characteristic diffraction peaks of ZIS. In addition, due to the introduction of ZIS, the diffraction peak density of CNNs is weakened. However, the peak position of CNNs does not move, showing that the addition of ZIS does not alter the crystalline phase structure of CNNs.
FT-IR spectra were chosen to characterize the variation of CNNs in the reaction process. As displayed in Figure 5b, several strong peaks appeared in the range of 900–1700 cm−1 are caused by the skeleton vibration of C–N and C=N heterocycles [45]. The peak located at 808 cm−1 may be attributed to the breathing vibration of the s-triazine units [46]. The peaks between 3000 and 3500 cm−1 are caused by H2O molecules bound with surfaces of CNNs and ZIS@CNNs and the amino groups formed by the hydrothermal reaction [47]. Compared with the FT-IR spectra of CNNs and three different ZIS@CNNs composites, the peaks of the four curves were similar, which indicates that the addition of ZIS does not damage the skeleton structure of CNNs. ZIS and CNNs can co-exist well in the hydrothermal reaction. It also shows that CNNs have excellent structural stability.
XPS was considered an effective means to further study the surface chemical state of the ZIS@CNNs composites. As could be seen in Figure 6a, five elements, Zn, In, N, S, and C, were clearly present in the tested material. Figure 6b show the high-resolution XPS spectrum of element C. The peaks of CNNs and ZIS@CNNs composites were very similar, which indicates the presence of CNNs in the composites. The strong peak located at 287.6 eV is assigned to N–C=N, and the weaker peak at 284.3 eV is assigned to graphite carbon nitride [48]. The peak of ZIS@CNNs composites negatively shifts to 287.5 and 284.2 eV, respectively. The binding energy of ZIS@CNNs composites decreases by 0.1 eV compared to CNNs. This may be caused by the intense electronic interactions between CNNs and ZIS, where CNNs gain photoelectrons [49]. Figure 6c show the high-resolution XPS spectrum of element N. Four peaks can be seen in Figure 6c, and the intensity of these four peaks gradually decreased from 398.0 to 403.7 eV. The peak with the highest binding energy at 398.0 eV is caused by the C–N=C bond. The peak at 399.6 eV is caused by N–(C)3 groups. The peak at 400.5 eV indicates the presence of N–H bonds [50]. The weak peak at 403.7 eV is caused by charge buildup due to electron transfer. In comparison with CNNs, the binding energy of elements C and N of the ZIS@CNNs composites negatively shift, indicating the state of CNNs gaining photoelectrons. To further analyze the electronic state of ZIS, high-resolution XPS spectra of Zn, In, and S elements were analyzed. The peaks of ZIS and ZIS@CNNs composites are very similar, which indicates the existence of ZIS in the composites. Figure 6d shows the high-resolution XPS spectrum of element S. It could be seen two peaks with different peak widths. The narrow peak at 160.6 eV is the characteristic peak of S 2p3/2, and the broad peak at 161.8 eV is the characteristic peak of S 2p1/2, indicating the S2− chemical state in the ZIS@CNNs composite. Figure 6e shows the high-resolution XPS spectrum of element In, the peaks at 451.6 and 444.2 eV are the characteristic peaks of In 3d3/2 and In 3d5/2, respectively, which indicated the In3+ chemical state in ZIS@CNNs composite [51]. Figure 6f shows two peaks at 1044.8 and 1021.3 eV, which are the characteristic peaks of Zn 2p1/2 and Zn 2p3/2, respectively, indicating the Zn2+ chemical state in the ZIS@CNNs composite [52]. In comparison with ZIS, the binding energy of element S, In, Zn of the ZIS@CNNs composites positively shifts, indicating the state of ZIS losing photoelectrons [53]. Combining the above results of the gain and loss of photoelectrons in CNNs and ZIS, it can be judged that photoelectrons flow from ZIS to CNNs.

3.2. Light Absorption Characteristics

To study the absorption properties of CNNs and three different composites for visible light, their UV-Vis DRS was measured [54]. Figure 7a shows that the absorption edges of the three different ZIS@CNNs composites are significantly redshift to the visible region in comparison with pure CNNs, and the 30% ZIS@CNNs composite shows the largest absorption threshold, which indicates that the ZIS@CNNs composites have stronger absorption of visible light than pure CNNs, and the 30% ZIS@CNNs composite has the best absorption performance. From the results of TEM and SEM as well as BET, it is obvious that the introduction of ZIS makes the CNNs porous structures as well as the increased area of visible light contact, which allows more visible light to be captured through the pores into the gaps between the CNNs; thus, the composites exhibit a strong light capture capability. In addition, the Tauc curves (Figure 7b) are obtained using Kubelka–Munk means [55,56]:
( α h v ) = A ( h v - E g ) n
where the absorption index, the Planck constant, the characteristic constant, and the frequency are defined as α, h, A, and v, respectively. The value of n is determined using the material’s transition characteristics: n = 1/2 for the direct band gap and n = 2 for the indirect band gap. The Eg of pure CNNs, ZIS, 20% ZIS@CNNs, 30% ZIS@CNNs, and 40% ZIS@CNNs composites are 2.7, 2.39, 2.63, 2.52, and 2.59 eV, respectively.
The isolation efficiency of photogenerated carriers has an essential influence on the PCP performance of the material. To study the carrier separation efficiency of different materials [57], PL tests were performed, which can be seen in Figure 8. The greater the decay of PL peak intensity, the higher the carrier separation efficiency [58]. Compared with pure CNNs, the peak intensities of the three different ratios of composites decayed significantly, with the 30% ZIS@CNNs composite showing the greatest decay. This may be due to the introduction of ZIS, which makes the ZIS@CNNs composites exhibit significant fluorescence quenching. The above results indicate that the reunited efficiency of the photogenerated electron-hole pairs of the composites is significantly reduced, and the 30% ZIS@CNNs composite has the lowest carrier recombination efficiency.

3.3. PCP Effects

In order to test the protection effect of CNNs and three different ZIS@CNNs composites on 316 SS, we conducted OCP tests on the above four materials [59]. It could be seen from Figure 9a that the OCP of all materials dropped steeply as the light source was turned on, and the potential drops of three different ZIS@CNNs composites were significantly higher than that of pure CNNs. In addition, 30% of ZIS@CNNs exhibited the largest potential drop (0.37 V), which was 7.4 times greater than that of pure CNNs (0.05 V). After the power was turned off, the OCP first rosed sharply and then slowly and finally stabilized around a certain value. As could be seen from Figure 9a, the stable values of the rebound potentials of both ZIS@CNNs composites and CNNs were lower than the self-corrosion potential of 316 SS (−0.18 V vs. SCE), which may be due to the fact that the electrons in ZIS@CNNs composites and CNNs did not disappear completely after the light was turned off. There are still some electrons transferred from three different ZIS@CNNs composites and CNN photoanodes to the 316 SS surface, indicating that they still played a continuous protective role for 316 SS in the dark state. In addition, three different ZIS@CNNs composites rebounded to more negative potential values compared with pure CNNs, indicating that the composite had a better continuous protection effect than pure CNNs in the dark state. The potentials of three different ZIS@CNNs composites remained stable after four on-off photocycles, indicating that the composites can provide stable protection to 316 SS for a long time. In addition, to test the protection effect of the composite material on other metals, we conducted OCP tests on 304 SS under the same conditions as 316 SS. As shown in Figure S3, the potentials of 304 SS connected to CNNs and 30% ZIS@CNNs composites both dropped. It can be clearly seen that ZIS@CNNs composites have obvious protection for 304 SS, and 30% ZIS@CNNs composite has a much better protection effect for 304 SS than CNNs. In addition, in comparison with the protection of CNNs and 30% ZIS@CNNs composite for 316 SS (Figure 9e), the protection for 304 SS is weaker than 316 SS.
CV curves were used to analyze the electrochemical activity of pure CNNs and three different ZIS@CNNs composites. The CV curves of four materials were tested at different sweep speeds, and the results are shown in Figure 10a–d. The area surrounded by the CV curves of the three ZIS@CNNs composites was larger than that of the CNNs at the same sweep speed, and the area surrounded by the CV curves of the 30% ZIS@CNNs composite was the largest. The scatter plot of Figure 10e was obtained with the sweep speed as the vertical dimension and the current variation of the CV curve as the vertical coordinate. The data were fitted to obtain four straight lines with different slopes, and the slope allowed us to obtain the capacitance values of different materials. The larger the slope, the larger the capacitance value [60]. Further, the electrochemical active area correlated with the size of the capacitance value; the larger the capacitance value, the larger the active area. Therefore, the slope size of the four lines in Figure 10e could be used to qualitatively analyze the effect of the size of the electrochemically active area of the four materials on the protection abilities of the materials. Figure 10e shows the slopes of the three different ZIS@CNNs composites were significantly larger than that of CNNs, and the slope of the 30% ZIS@CNNs composite was the largest, which indicates that the capacitance value of the three different ZIS@CNNs composites was larger than that of CNNs and the 30% ZIS@CNNs composite had the largest capacitance. Therefore, the electrochemically active areas of three different composites were greater than that of CNNs. 30% ZIS@CNNs composite had the largest electrochemically active area. This is in agreement with the 30% ZIS@CNNs composites exhibiting the optimal protective properties. In addition, the slope of the 40% ZIS@CNNs composite was slightly smaller than that of the 20% ZIS@CNNs composite, which may be attributed to the increase in the proportion of ZIS, which caused a large amount of ZIS to accumulate, resulting in smaller gaps between the nano-leaves, resulting in a reduced active area. However, more heterojunction channels were formed between the ZIS and CNNs in the 40% ZIS@ CNNs composites, which allowed the 40% ZIS@CNNs composites to exhibit better protective properties than the 20% ZIS@CNNs composites.
The photoanodes generate current under light; therefore, photocurrent density is another effective approach to characterize the protection effect of photoanodes on 316 SS [61]. The i–t curves of 316 SS coupled with CNNs and three different ZIS@CNNs composites were measured. Figure 9b shows that when the light source was switched on, the photocurrent densities of all materials increased rapidly to a certain value. It was obvious from Figure 9b that the photocurrent densities of three different ZIS@CNNs composites were larger than that of pure CNNs and ZIS, and 30% of ZIS@CNNs composites exhibited the largest photocurrent densities (17.30 μA cm−2), which was more than four times that of pure CNNs (4.22 μA cm−2)and pure ZIS (3.02 μA cm−2). The current drops in the 100~150 s and 200~250 s closed-light phases For CNNs and the three different composites were not as rapid as in the open-light phase but more slowly, and the final steady current was not zero, which may be due to the fact that the carriers generated in the open-light phase did not disappear completely. Combined with the above and OCP results, it can be indicated that three different ZIS@CNNs composites could provide better protection for 316 SS than pure CNNs under light conditions, and the 30% ZIS@CNNs composite had the best protection effect.
Figure 9c shows the Tafel graph of 316 SS combined with CNNs and three different ZIS@CNNs composites. Table 1 shows the fitted values of the above curves. The order of Ecorr for different materials was that 30% ZIS@CNNs composite < 40% ZIS@CNNs composite < 20% ZIS@CNNs composite < CNNs. The 30% ZIS@CNNs composite has the most negative photogenerated potential, which is in agreement with the OCP data. In addition, the order of icorr for different materials is that 30% ZIS@CNNs composite > 40% ZIS@CNNs composite > 20% ZIS@CNNs composite > CNNs. The icorr for 30% ZIS@CNNs composite is more than three times that of CNNs. A larger icorr means faster electrode reaction rates [62]. icorr for 30% ZIS@CNNs is the largest, indicating this composite has the highest protection efficiency.
In order to study the electron transfer efficiency inside and between materials, we measured the EIS plots of 316 SS coupled with CNNs and three different ZIS@CNNs composites, as shown in Figure 9d. The impedance arcs of three different ZIS@CNNs composites were smaller than that of CNNs under the light condition, and 30% of the ZIS@CNNs composite has the smallest impedance arc. This indicates that the ZIS@CNNs composites generate a greater number of photogenerated electrons, and the formation of heterojunction channels accelerates the electron transfer, which enhances the protection of the material for 316 SS. According to the measured impedance data, the fitted circuit used is the Rs (Qf Rf) (Cdl Rct) circuit. The fitting circuit diagram is shown in the internal diagram of Figure 9d. Rf, Rs, and Rct were the internal resistance of the photoanode, the resistance of 3.5 wt% NaCl solution, and the charge transfer resistance, respectively. Qf and Cdl were defined as internal film capacitance of the material and double-layer capacitance, respectively. Table 2 shows the values of the fitting. The smaller the Rct, the easier the charge transfer [63]. The order of Rct of different materials was that 30% ZIS@CNNs (426.5 Ω cm−2) < 40% ZIS@CNNs (556.6 Ω cm−2) < 20% ZIS@CNNs (874.2 Ω cm−2) < CNNs (934.1 Ω cm−2). It is obvious that the Rct of CNNs was about two times that of 30% ZIS@CNNs composite, which means that the transfer efficiency of electrons in 30% ZIS@CNNs composite is higher than that in pure CNNs. Furthermore, 30% ZIS@CNNs composite has the highest electron transfer efficiency.
The stability of the material is an essential element in the practical implementation of the material. Figure 9f shows the XRD patterns of the material before and after the electrochemical test; the XRD images before and after the reaction are not significantly changed. This indicates that the photoanode material has good stability.

3.4. Mechanism Analysis

To determine the band edge positions of different materials and the valence band (VB) and conduction band (CB) positions of CNNs and ZIS and to elucidate the mechanism of electron transfer. The M-S curves of the five materials were measured. It can be seen from Figure 11 that the slopes of all the curves are positive, which indicates that the experimentally prepared ZIS, CNNs, and the three different ZIS@CNNs composites are n-type semiconductors [64]. Compared with CNNs, the flat-band potentials (Efb) of composites have a negative shift, and 30% ZIS@CNNs composite has the most negative shift. It is an indication that a greater number of photogenerated electrons can be transferred from the 30% ZIS@CNNs composite to the 316 SS surface for their protection than the other materials. It could be seen from Figure 11 that the Efb of CNNs was −0.6 V vs. SCE, and ECB could be calculated using the following formulae:
E fb ( vs .   NHE ) = E fb ( vs .   SCE ) 0.24
E CB = E fb ( vs .   NHE ) + 0.20
Thus, the ECB of CNNs is −0.56 eV. Similarly, the ECB of ZIS is −0.83 eV [65,66]. EVB of the semiconductor photoanodes can be calculated using the formula:
E VB = E fb E CB
Thus, the EVB of CNNs and ZIS were estimated to be 2.14 and 1.56 eV, respectively, which were consistent with the XPS VB spectra (Figure S4). Obviously, The ECB of the ZIS was slightly more critical than that of the CNNs, so the photogenerated electrons can be transferred from the ZIS to the CNNs surface and moved to the 316 SS surface (Figure 12). Therefore, the ZIS@CNNs composites generate more electrons than pure CNNs.

4. Conclusions

In summary, three different ZIS@CNNs composites were fabricated using calcination and hydrothermal methods. According to the results of SEM, TEM, and BET, the compounding of CNNs and ZIS leads to many mesopores being formed, resulting in a significant increase in visible light absorption efficiency and generating lots of high-speed nanochannels for charge transfer. In addition, UV-Vis DRS shows that the absorption band edges of composites have a red shift that results in an increase in the absorption of visible light by the material. Photocatalysis is a complex process of carrier generation, separation, compounding, and transfer. The result of PL shows that due to the introduction of ZIS, the isolation efficiency of photogenerated carriers of ZIS@CNNs composites is greatly improved. EIS shows that photogenerated carriers could rapidly transfer from the ZIS@CNNs photoanodes to the surface of 316 SS. Therefore, the introduction of ZIS brings a significant increase in both carrier isolation and migration efficiency, which leads to the improved PCP efficiency of ZIS@CNNs nanocomposite under visible light. The photocurrent density of the 30% ZIS@CNNs composite can achieve 17 μA cm−2, and the photogenerated potential drops to about −0.58 V, showing the effective protection of the composite. In conclusion, the ZIS@CNNs composite is a very friendly photoanode material to protect the metal from corrosion in the ocean environment.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/coatings13091479/s1, Figure S1: Picture of CNNs and three different ZIS@CNNs composite material; Figure S2: Schematic diagrams of experimental setup for (a) OCP, EIS, Tafel (b) photocurrent densities tests; Figure S3: OCP-t curves of 304 SS coupled with CNNs and 30% ZIS@CNNs composite; Figure S4: XPS valence band (VB) spectra of CNNs and ZIS; Table S1: Comparison of ZIS@CNNs photoanode with other photoanodes.

Author Contributions

Experiment, Writing—Original draft, Picture drawing, review and editing, W.L.; Experiment, Z.Y.; Resources, Y.L.; Resources, P.Z.; Writing—Review and editing, Supervision, Funding acquisition, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Natural Science Foundation of Shandong Province of China (No. ZR2023ME205), the National Natural Science Foundation of China (No. 51801109), and the Science and Technology Support Plan for Youth Innovation of Colleges in Shandong Province (No. 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. Guo, H.; Zhang, Y.; Wang, S.; Li, L.; Wang, W.; Sun, Q. In-situ generation of Bi2S3 to construct WO3/BiVO4/Bi2S3 heterojunction for photocathodic protection of 304 SS. J. Electroanal. Chem. 2022, 907, 116033. [Google Scholar] [CrossRef]
  2. Pu, J.; Wang, X.; Liu, J.; Ren, M.; Nan, Y.; Liu, M.; Xu, H.; Yang, L.; Huang, Y.; Hou, B. PDA decorated spaced TiO2 nanotube array photoanode material for photocathodic protection of 304 stainless steels. J. Electroanal. Chem. 2022, 914, 116319. [Google Scholar] [CrossRef]
  3. Wang, X.; Xu, H.; Nan, Y.; Sun, X.; Duan, J.; Huang, Y.; Hou, B. Research progress of TiO2 photocathodic protection to metals in marine environment. J. Oceanol. Limnol. 2020, 38, 1018–1044. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, J.; Yang, H.; Wang, Y.; Cui, X.; Wen, Z.; Liu, Y.; Fan, L.; Feng, J. Facile fabrication of SnO2 modified TiO2 nanorods film for efficient photocathodic protection of 304 stainless steel under simulated solar light. Corros. Sci. 2020, 176, 108927. [Google Scholar] [CrossRef]
  5. Qiu, J.; Wei, S.; Pan, J.; Liang, Z.; Ma, J.; Liang, Q.; Li, C. g-C3N4-loaded carbon nanofiber for efficient photocatalytic hydrogen production. Integr. Ferroelectr. 2023, 231, 70–77. [Google Scholar] [CrossRef]
  6. Zhang, X.; Liang, H.; Li, C.; Bai, J. 1D CeO2/g-C3N4 type II heterojunction for visible-light-driven photocatalytic hydrogen evolution. Inorg. Chem. Commun. 2022, 144, 109838. [Google Scholar] [CrossRef]
  7. Ren, T.; Dang, Y.; Xiao, Y.; Hu, Q.; Deng, D.; Chen, J.; He, P. Depositing Ag nanoparticles on g-C3N4 by facile silver mirror reaction for enhanced photocatalytic hydrogen production. Inorg. Chem. Commun. 2021, 123, 108367. [Google Scholar] [CrossRef]
  8. Li, X.; Liu, T.; Tian, F.; Tao, X.; Wu, Z.S. Enhanced carrier transport and visible light response in CA-β-CD/g-C3N4/Ag2O 2D/0D heterostructures functionalized with cyclodextrin for effective organic degradation. Korean J. Chem. Eng. 2022, 39, 2972–2982. [Google Scholar] [CrossRef]
  9. Yu, X.; Zhang, X.; Zhao, J.; Xu, L.; Yan, J. Flower-like shaped Bi12TiO20/g-C3N4 heterojunction for effective elimination of organic pollutants: Preparation, characterization, and mechanism study. Appl. Organomet. Chem. 2020, 34, e5702. [Google Scholar] [CrossRef]
  10. Mao, N.; Jiao, Y.; Duan, X. g-C3N4/ZnO heterojunction as Fenton-like catalyst for degradation of organic pollution. Mater. Res. Bull. 2022, 151, 111818. [Google Scholar] [CrossRef]
  11. Zhu, X.; Deng, H.; Cheng, G. Facile construction of g-C3N4-W18O49 heterojunction with improved charge transfer for solar-driven CO2 photoreduction. Inorg. Chem. Commun. 2021, 132, 108814. [Google Scholar] [CrossRef]
  12. Huang, M.; Chen, C.; Wang, T.; Sui, Q.; Zhang, K.; Li, B. Cadmium-sulfide/gold/graphitic-carbon-nitride sandwich heterojunction photocatalyst with regulated electron transfer for boosting carbon-dioxide reduction to hydrocarbon. J. Colloid Interface Sci. 2022, 613, 575–586. [Google Scholar] [CrossRef]
  13. Zhang, R.; Cao, Y.; Doronkin, D.; Ma, M.; Dong, F.; Zhou, Y. Single-atom dispersed Zn-N3 active sites bridging the interlayer of g-C3N4 to tune NO oxidation pathway for the inhibition of toxic by-product generation. Chem. Eng. J. 2023, 454, 140084. [Google Scholar] [CrossRef]
  14. Liang, Y.; Zeng, Z.; Yang, J.; Yang, G.; Han, Y. Designing heterointerface in BiOBr/ g-C3N4 photocatalyst to enhance visible-light-driven photocatalytic performance in water purification. Colloids Surf. A Physicochem. Eng. Asp. 2021, 624, 126796. [Google Scholar] [CrossRef]
  15. Ni, Y.; Wang, M.; Liu, L.; Li, M.; Hu, S.; Lin, J.; Sun, J.; Yue, T.; Zhu, M.; Wang, J. Efficient and reusable photocatalytic river water disinfection by addictive graphitic carbon nitride/magnesium oxide nano-onions with particular “nano-magnifying glass effect”. J. Hazard. Mater. 2022, 439, 129533. [Google Scholar] [CrossRef] [PubMed]
  16. Rao, F.; Zhong, J.; Li, J. Improved visible light responsive photocatalytic hydrogen production over g-C3N4 with rich carbon vacancies. Ceram. Int. 2022, 48, 1439–1445. [Google Scholar] [CrossRef]
  17. Shi, Z.; Rao, L.; Wang, P.; Zhang, L. The photocatalytic activity and purification performance of g-C3N4/carbon nanotubes composite photocatalyst in underwater environment. Environ. Sci. Pollut. Res. 2022, 29, 83981–83992. [Google Scholar] [CrossRef]
  18. Cao, J.; Cai, J.; Li, R.; Han, J.; Liu, J.; Huang, M. A novel 3D yolk-double-shell Au@CdS/ g-C3N4 nanostructure with enhanced photoelectrochemical and photocatalytic properties. J. Phys. Chem. C 2022, 126, 4939–4947. [Google Scholar] [CrossRef]
  19. Saeidpour, S.; Khoshnevisan, B.; Boroumand, Z. Synthesis and characterization of a g-C3N4/TiO2-ZnO nanostructure for photocatalytic degradation of methylene blue. Nano Futures 2022, 6, 035001. [Google Scholar] [CrossRef]
  20. Liu, F.; Bi, S.; Wang, W.; Duan, Q.; Feng, Y.; Chen, J.; Luo, R.; Huang, Y.; Lee, J. Preparation of a modified g-C3N4 catalyst library and realization of a two-dimensional screening reaction. New J. Chem. 2021, 45, 2582–2588. [Google Scholar] [CrossRef]
  21. Deng, L.; Sun, J.; Sun, J.; Wang, X.; Shen, T.; Zhao, R.; Zhang, Y.; Wang, B. Improved performance of photosynthetic H2O2 and photodegradation by K-, P-, O-, and S-co-doped g-C3N4 with enhanced charge transfer ability under visible light. Appl. Surf. Sci. 2022, 597, 153586. [Google Scholar] [CrossRef]
  22. Ge, Y.; Guo, X.; Zhou, D.; Liu, J. Construction and excellent photoelectric synergistic anticorrosion performance of Z-scheme carbon nitride/tungsten oxide heterojunctions. Nanoscale 2022, 14, 12358–12376. [Google Scholar] [CrossRef] [PubMed]
  23. Zhou, W.; Lu, S.; Chen, X. Anionic donor-acceptor conjugated polymer dots/g-C3N4 nanosheets heterojunction: High efficiency and excellent stability for co-catalyst-free photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2022, 608, 912–921. [Google Scholar] [CrossRef]
  24. Yan, X.; Xie, M.; Pan, L.; Ai, T.; Li, Z.; Niu, Y. Piezo-photocatalytic activity of NaNbO3/g-C3N4 heterojunction for dye wastewater degradation. J. Mater. Sci. Mater. Electron. 2023, 34, 695. [Google Scholar] [CrossRef]
  25. Vavilapalli, D.S.; Peri, R.G.; Muthuraaman, B.; Sridharan, K.; Rao, M.R.; Singh, S. Enhanced photocatalytic and photoelectrochemical performance of KBiFe2O5/g-C3N4 heterojunction photocatalyst under visible light. Phys. B Condens. Matter 2023, 648, 414411. [Google Scholar] [CrossRef]
  26. Huang, Y.; Li, B.; Wu, F.; Yang, B. Fabrication of novel flower-like Co3O4/g-C3N4 heterojunction for tetracycline degradation under visible light irradiation. Mater. Lett. 2022, 311, 131538. [Google Scholar] [CrossRef]
  27. Wang, Z.; Zhang, J.; Ji, X.; Wu, H.; Xu, X.; Zhan, J.; Shi, H.; Liu, W.; Tang, T. Construction of CdSe/ZnIn2S4 Z-Scheme heterojunction for enhanced photocatalytic degradation of tetracycline. Mater. Sci. Eng. B 2022, 286, 116065. [Google Scholar] [CrossRef]
  28. Yang, N.; Li, J. Construction of a 0D/2D heterojunction based on ZnO nanoparticles and ZnIn2S4 nanosheets to improve photocatalytic degradation efficiency. Opt. Mater. 2021, 115, 111040. [Google Scholar] [CrossRef]
  29. Wang, Z.; Su, B.; Xu, J.; Hou, Y.; Ding, Z. Direct Z-scheme ZnIn2S4/LaNiO3 nanohybrid with enhanced photocatalytic performance for H2 evolution. Int. J. Hydrogen. Energ. 2020, 45, 4113–4121. [Google Scholar] [CrossRef]
  30. Zhu, J.; Li, H.; Cui, X.; Yang, Z.; Chen, B.; Li, Y.; Zhang, P.; Li, J. Efficient photocathodic protection performance of ZnIn2S4 nanosheets/SnO2 quantum dots/TiO2 nanotubes composite for 316 SS under visible light. J. Alloys Compd. 2022, 926, 166901. [Google Scholar] [CrossRef]
  31. Cao, S.; Yu, J.; Wageh, S.; Al-Ghamdi, A.; Mousavi, M.; Ghasemi, J.; Xu, F. H2-production and electron-transfer mechanism of a noble-metal-free WO3@ZnIn2S4 S-scheme heterojunction photocatalyst. J. Mater. Chem. A 2022, 10, 17174–17184. [Google Scholar] [CrossRef]
  32. Wang, K.; Shao, X.; Cheng, Q.; Li, K.; Le, X.; Wang, G.; Wang, H. In situ-Illuminated X-Ray photoelectron spectroscopy investigation of S-scheme Ta2O5/ZnIn2S4 core-shell hybrid nanofibers for highly efficient solar-driven CO2 overall splitting. Solar RRL 2022, 6, 2200736. [Google Scholar] [CrossRef]
  33. Zhang, L.; Qiu, J.; Dai, D.; Zhou, Y.; Liu, X.; Yao, J. Cr-metal-organic framework coordination with ZnIn2S4 nanosheets for photocatalytic reduction of Cr(VI). J. Clean. Prod. 2022, 341, 130891. [Google Scholar] [CrossRef]
  34. Dang, X.; Xie, M.; Dai, F.; Guo, J.; Liu, J.; Lu, X. Ultrathin 2D/2D ZnIn2S4/g-C3N4 nanosheet heterojunction with atomic-level intimate interface for photocatalytic hydrogen evolution under visible light. Adv. Mater. Interfaces 2021, 8, 2196–7350. [Google Scholar] [CrossRef]
  35. Hou, L.; Li, W.; Wu, Z.; Wei, Q.; Yang, H.; Jiang, Y.; Wang, T.; Wang, Y.; He, Q. Embedding ZnCdS@ZnIn2S4 into thiazole-modified g-C3N4 by electrostatic self-assembly to build dual Z-scheme heterojunction with spatially separated active centers for photocatalytic H2 evolution and ofloxacin degradation. Sep. Purif. Technol. 2022, 290, 120858. [Google Scholar] [CrossRef]
  36. Zhang, G.P.; Zhu, X.W.; Chen, D.Y.; Li, N.J.; Xu, Q.F.; Li, H.; He, J.H.; Xu, H.; Lu, J.M. Hierarchical Z-scheme g-C3N4/Au/ZnIn2S4 photocatalyst for highly enhanced visible-light photocatalytic nitric oxide removal and carbon dioxide conversion. Environ. Sci. Nano 2020, 7, 676–687. [Google Scholar] [CrossRef]
  37. Liu, X.M.; Wang, S.Q.; Yang, F.; Zhang, Y.C.; Yan, L.S.; Li, K.X.; Guo, H.Q.; Yan, J.J.; Lin, J. Construction of Au/g-C3N4/ZnIn2S4 plasma photocatalyst heterojunction composite with 3D hierarchical microarchitecture for visible-light-driven hydrogen production. Int. J. Hydrogen Energy 2022, 47, 2900–2913. [Google Scholar] [CrossRef]
  38. Li, Y.; Lu, Y.; Wang, Y.; Dong, L.; Chao, M.; Sun, J.; Zhao, Z.; Zhang, J. One-step synthesis of high photocatalytic graphitic carbon nitride porous nanosheets. Nanotechnology 2020, 31, 464001. [Google Scholar] [CrossRef]
  39. Ai, L.; Fan, H. CTAB-melamine molecular crystals as precursor for synthesis of layered carbon nitride porous nanostructures with enhanced photocatalytic activity for hydrogen production. Mater. Today Commun. 2021, 29, 102780. [Google Scholar] [CrossRef]
  40. Zhu, K.; Luan, X.; Matras-Postolek, K.; Yang, P. 2D/2D MoS2/g-C3N4 layered heterojunctions with enhanced interfacial electron coupling effect. J. Electroanal. Chem. 2021, 893, 115350. [Google Scholar] [CrossRef]
  41. Kesir, M.; Yildiz, I.; Bilgen, S.; Sokmen, M. Role of a novel cationic gemini surfactant (CGS) on a one-step sol-gel process and photocatalytic properties of TiO2 powders. J. Water Health 2022, 20, 1629–1643. [Google Scholar] [CrossRef]
  42. Wang, X.; Chen, X.; Thomas, A.; Fu, X.; Antonietti, M. Metal-containing carbon nitride compounds: A new functional organic-metal hybrid material. Adv. Mater. 2009, 21, 1609–1612. [Google Scholar] [CrossRef]
  43. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef] [PubMed]
  44. Jo, W.; Natarajan, T. Influence of TiO2 morphology on the photocatalytic efficiency of direct Z-scheme g-C3N4/TiO2 photocatalysts for isoniazid degradation. Chem. Eng. J. 2015, 281, 549–565. [Google Scholar] [CrossRef]
  45. Guo, F.; Huang, X.; Chen, Z.; Cao, L.; Cheng, X.; Chen, L.; Shi, W. Construction of Cu3P-ZnSnO3-g-C3N4 p-n-n heterojunction with multiple built-in electric fields for effectively boosting visible-light photocatalytic degradation of broad-spectrum antibiotics. Sep. Purif. Technol. 2021, 265, 118477. [Google Scholar] [CrossRef]
  46. Zhang, W.; Shi, W.; Sun, H.; Shi, Y.; Luo, H.; Jing, S.; Fan, Y.; Guo, F.; Lu, C. Fabrication of ternary CoO/g-C3N4/Co3O4 nanocomposite with p-n-p type heterojunction for boosted visible-light photocatalytic performance. J. Chem. Technol. Biotechnol. 2021, 96, 1854–1863. [Google Scholar] [CrossRef]
  47. Ni, T.; Yang, Z.; Cao, Y.; Lv, H.; Liu, Y. Rational design of MoS2/g-C3N4/ZnIn2S4 hierarchical heterostructures with efficient charge transfer for significantly enhanced photocatalytic H2 production. Ceram. Int. 2021, 47, 22985–22993. [Google Scholar] [CrossRef]
  48. Ye, X.; Zhu, T.; Hui, Z.; Wang, X.; Wei, J.; Chen, S. Revealing the transfer mechanisms of photogenerated charge carriers over g-C3N4/ZnIn2S4 composite: A model study for photocatalytic oxidation of aromatic alcohols with visible light. J. Catal. 2021, 401, 149–159. [Google Scholar] [CrossRef]
  49. Fu, D.; Han, G.; Liu, F.; Xiao, Y.; Wang, H.; Liu, R.; Liu, C. Visible-light enhancement of methylene blue photodegradation by graphitic carbon nitride-titania composites. Mater. Sci. Semicond. Process. 2014, 27, 966–974. [Google Scholar] [CrossRef]
  50. Wang, F.; Zhao, Y.; Zhang, M.; Wu, J.; Liu, G.; He, P.; Qi, Y.; Li, X.; Zhou, Y.; Li, J. Bimetallic sulfides ZnIn2S4 modified g-C3N4 adsorbent with wide temperature range for rapid elemental mercury uptake from coal-fired flue gas. Chem. Eng. J. 2021, 426, 131343. [Google Scholar] [CrossRef]
  51. Yang, L.; Zhao, J.; Wang, Z.; Wang, L.; Zhao, Z.; Li, S.; Li, G.; Cai, Z. Facile construction of g-C3N4/ZnIn2S4 nanocomposites for enhance Cr(VI) photocatalytic reduction. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2022, 276, 121184. [Google Scholar] [CrossRef] [PubMed]
  52. Zheng, H.; Liu, Y.; Zhou, Y.; Zhao, D.; Wang, D.; Yun, L.; Zhang, D.; Zhang, L. Improved photocathodic protection performance of g-C3N4/rGO/ZnS for 304 stainless steel. J. Phys. Chem. Solids 2021, 148, 109672. [Google Scholar] [CrossRef]
  53. Mishra, N.; Kuila, A.; Saravanan, P.; Bahnemann, D.; Jang, M.; Routu, S. Simultaneous S-scheme promoted Ag@AgVO3/g-C3N4/CeVO4 heterojunction with enhanced charge separation and photo redox ability towards solar photocatalysis. Chemosphere 2023, 326, 138496. [Google Scholar] [CrossRef]
  54. Rajendran, R.; Vignesh, S.; Sasireka, A.; Priya, P.; Suganthi, S.; Raj, V.; Sundar, J.K.; Srinivasan, M.; Shkir, M.; AlFaify, S. Investigation on novel Cu2O modified g-C3N4/ZnO heterostructures for efficient photocatalytic dye degradation performance under visible-light exposure. Colloid Interface Sci. Commun. 2021, 44, 100480. [Google Scholar] [CrossRef]
  55. Wu, B.; Shan, C.; Zhang, X.; Zhao, H.; Ma, S.; Shi, Y.; Yang, J.; Bai, H.; Liu, Q. CeO2/Co3O4 porous nanosheet prepared using rose petal as biotemplate for photo-catalytic degradation of organic contaminants. Appl. Surf. Sci. 2021, 543, 148677. [Google Scholar] [CrossRef]
  56. Sivanandan, V.; Prasad, A. Lactose monohydrate (C12H22O11 center dot H2O) mediated synthesis and spectral analysis of nanocrystalline Ni0.5Cu0.5Fe2O. Int. J. Mater. Res. 2021, 112, 980–984. [Google Scholar]
  57. Huang, X.; Xu, X.; Yang, R.; Fu, X. Synergetic adsorption and photocatalysis performance of g-C3N4/Ce-doped MgAl-LDH in degradation of organic dye under LED visible light. Colloid Surface A Physicochem. Eng. Asp. 2022, 643, 128738. [Google Scholar] [CrossRef]
  58. Hoang, T.; Nguyen, P.; Shin, E. Effect of morphological modification over g-C3N4 on photocatalytic hydrogen evolution performance of g-C3N4-Pt photocatalysts. Catalysts 2023, 13, 10092. [Google Scholar] [CrossRef]
  59. Zhang, T.; Liu, Y.; Liang, J.; Wang, D. Enhancement of photoelectrochemical and photocathodic protection properties of TiO2 nanotube arrays by simple surface UV treatment. Appl. Surf. Sci. 2017, 394, 440–445. [Google Scholar] [CrossRef]
  60. Tian, J.; Chen, Z.; Ma, L.; Hou, J.; Feng, C.; Jing, J.; Sun, M.; Chen, D. Fabrication of flower-like WO3/ZnIn2S4 composite with special electronic transmission channels to improve carrier separation for photoinduced cathodic protection and electron storage. Appl. Surf. Sci. 2023, 607, 155019. [Google Scholar] [CrossRef]
  61. Gong, D.; Xu, S.; Zhang, K.; Du, L.; Qiu, P. Enhancing photoelectrochemical cathodic protection performance by facile tuning sulfur redox state in sacrificial agents. Chem. Eng. J. 2023, 451, 138552. [Google Scholar] [CrossRef]
  62. Yang, Z.; Li, H.; Cui, X.; Zhu, J.; Li, Y.; Zhang, P.; Li, J. Direct Z-scheme nanoporous BiVO4/CdS quantum dots heterojunction composites as photoanodes for photocathodic protection of 316 stainless steel under visible light. Appl. Surf. Sci. 2022, 603, 154394. [Google Scholar] [CrossRef]
  63. Kathiravan, S.; Kaliaraj, G.; Kumar, R.; Kirubaharan, A. A novel experimental setup for in situ oxidation behavior study of Nb/Hf/Ti (C-103) alloy for high temperature environments. Mater. Lett. 2021, 302, 130336. [Google Scholar] [CrossRef]
  64. Igbari, F.; Essien, E.; Abdulwahab, K.; Nejo, A.; Adetona, A.; Adams, L.A. Bipolar conductivity in amorphous Cu–Al–O thin films prepared by r.f. magnetron sputtering. Mater. Sci. Semicond. Process. 2021, 123, 105557. [Google Scholar] [CrossRef]
  65. Duan, Z.; Zhao, X.; Chen, L. BiVO4/Cu0.4V2O5 composites as a novel Z-scheme photocatalyst for visible-light-driven CO2 conversion. J. Environ. Chem. Eng. 2021, 9, 104628. [Google Scholar] [CrossRef]
  66. Wang, W.; Feng, X.; Chen, L.; Zhang, F. Z-Scheme Cu2O/Bi/BiVO4 nanocomposite photocatalysts: Synthesis, characterization, and application for CO2 photoreduction. Ind. Eng. Chem. Res. 2021, 60, 18384–18396. [Google Scholar] [CrossRef]
Figure 1. Synthesis schematic of ZIS@CNNs composites.
Figure 1. Synthesis schematic of ZIS@CNNs composites.
Coatings 13 01479 g001
Figure 2. SEM image of (a) CNNs, (b) ZIS, (c) 30% ZIS@CNNs, (d) EDS spectrum.
Figure 2. SEM image of (a) CNNs, (b) ZIS, (c) 30% ZIS@CNNs, (d) EDS spectrum.
Coatings 13 01479 g002
Figure 3. TEM image of (a) CNNs, (b) ZIS, (c) 30% ZIS@CNNs. HRTEM images of (d) 30% ZIS@CNNs composite.
Figure 3. TEM image of (a) CNNs, (b) ZIS, (c) 30% ZIS@CNNs. HRTEM images of (d) 30% ZIS@CNNs composite.
Coatings 13 01479 g003
Figure 4. N2 adsorption-desorption curves and pore size distribution isotherms of (a) CNNs and (b) 30% ZIS@CNNs composite.
Figure 4. N2 adsorption-desorption curves and pore size distribution isotherms of (a) CNNs and (b) 30% ZIS@CNNs composite.
Coatings 13 01479 g004
Figure 5. (a) The XRD patterns (b) FT-IR spectrum of CNNs and three different ZIS@CNNs composites.
Figure 5. (a) The XRD patterns (b) FT-IR spectrum of CNNs and three different ZIS@CNNs composites.
Coatings 13 01479 g005
Figure 6. Full scan survey XPS spectra (a) and high-resolution XPS spectra of (b) C 1s, (c) N 1s, (d) Zn 2p, (e) In 3d and (f) S 2p from CNNs, ZIS, and 30% ZIS@CNNs composite.
Figure 6. Full scan survey XPS spectra (a) and high-resolution XPS spectra of (b) C 1s, (c) N 1s, (d) Zn 2p, (e) In 3d and (f) S 2p from CNNs, ZIS, and 30% ZIS@CNNs composite.
Coatings 13 01479 g006
Figure 7. (a) UV-Vis DRS, (b) plot of (αhν)2 versus for the band gap energy of CNNs and ZIS@CNNs composites with different ZIS content.
Figure 7. (a) UV-Vis DRS, (b) plot of (αhν)2 versus for the band gap energy of CNNs and ZIS@CNNs composites with different ZIS content.
Coatings 13 01479 g007
Figure 8. Steady-state PL spectra of CNNs, and ZIS@CNNs composites with different ZIS content.
Figure 8. Steady-state PL spectra of CNNs, and ZIS@CNNs composites with different ZIS content.
Coatings 13 01479 g008
Figure 9. (a) OCP-t graphs, (b) The I-t charts, (c) Tafel plots, and (d) Nyquist diagrams of 316 SS coupled with CNNs and three different ZIS@CNNs composites. (e) Different potential drops of 304 SS and 316 SS coupled with CNNs and ZIS@CNNs, respectively. (f) XRD patterns of 30% ZIS@CNNs composite before and after the electrochemical test.
Figure 9. (a) OCP-t graphs, (b) The I-t charts, (c) Tafel plots, and (d) Nyquist diagrams of 316 SS coupled with CNNs and three different ZIS@CNNs composites. (e) Different potential drops of 304 SS and 316 SS coupled with CNNs and ZIS@CNNs, respectively. (f) XRD patterns of 30% ZIS@CNNs composite before and after the electrochemical test.
Coatings 13 01479 g009
Figure 10. CV curves of (a) CNNs, (b) 20% ZIS@CNNs, and (c) 30% ZIS@CNNs (d) 40% ZIS@CNNs composite (e) Plots of the current densities of the CV curves under different scanning speeds.
Figure 10. CV curves of (a) CNNs, (b) 20% ZIS@CNNs, and (c) 30% ZIS@CNNs (d) 40% ZIS@CNNs composite (e) Plots of the current densities of the CV curves under different scanning speeds.
Coatings 13 01479 g010
Figure 11. (ae) M-S curves of CNNs, ZIS, and ZIS@CNNs composites.
Figure 11. (ae) M-S curves of CNNs, ZIS, and ZIS@CNNs composites.
Coatings 13 01479 g011
Figure 12. Electron-hole pair migration mechanism of ZIS@CNNs composite.
Figure 12. Electron-hole pair migration mechanism of ZIS@CNNs composite.
Coatings 13 01479 g012
Table 1. Fitting values of Ecorr and icorr obtained from Figure 9c.
Table 1. Fitting values of Ecorr and icorr obtained from Figure 9c.
SamplesEcorr (V vs. SCE)icorr (μA cm−2)
CNNs−0.2714.08
20% ZIS@CNNs−0.3326.46
30% ZIS@CNNs−0.6213.54
40% ZIS@CNNs−0.43410.93
Table 2. Fitting values of EIS data.
Table 2. Fitting values of EIS data.
Rs (Ω cm−2)Qf (F·cm−2·10−2)Rf (Ω cm−2)Cdl (F·cm−2·10−2)Rct
(Ω cm−2)
CNNs4.6055.411 × 10−2948.70.01807934.1
20% ZIS@CNNs4.5744.983 × 10−2976.30.02874.2
30% ZIS@CNNs4.2781.803 × 10−315903.916 × 10−4426.5
40% ZIS@CNNs2.7094.235 × 10−31135.32.721 × 10−4556.6
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

Li, W.; Yang, Z.; Li, Y.; Zhang, P.; Li, H. Highly Efficient Photocathodic Protection Performance of ZIS@CNNs Composites under Visible Light. Coatings 2023, 13, 1479. https://doi.org/10.3390/coatings13091479

AMA Style

Li W, Yang Z, Li Y, Zhang P, Li H. Highly Efficient Photocathodic Protection Performance of ZIS@CNNs Composites under Visible Light. Coatings. 2023; 13(9):1479. https://doi.org/10.3390/coatings13091479

Chicago/Turabian Style

Li, Weitao, Zhanyuan Yang, Yanhui Li, Pengfei Zhang, and Hong Li. 2023. "Highly Efficient Photocathodic Protection Performance of ZIS@CNNs Composites under Visible Light" Coatings 13, no. 9: 1479. https://doi.org/10.3390/coatings13091479

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