Next Article in Journal
Enabling High-Level Worker-Centric Semantic Understanding of Onsite Images Using Visual Language Models with Attention Mechanism and Beam Search Strategy
Previous Article in Journal
Using Near-Surface-Mounted Small-Diameter Steel Wires to Improve Construction Efficiency in Strengthening Substandard Lapped Spliced Reinforced Concrete Beams
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Design of a One-Dimensional Zn3In2S6/NiFe2O4 Composite Material and Its Photocathodic Protection Mechanism Against Corrosion

College of Civil Engineering, Qingdao University of Technology, Qingdao 266520, China
*
Author to whom correspondence should be addressed.
Buildings 2025, 15(6), 958; https://doi.org/10.3390/buildings15060958
Submission received: 1 February 2025 / Revised: 4 March 2025 / Accepted: 14 March 2025 / Published: 18 March 2025
(This article belongs to the Special Issue Application of Nanotechnology in Buildings Material)

Abstract

:
Z-scheme Zn3In2S6/NiFe2O4 nanocomposites were prepared by electrospinning and hydrothermal methods, and their photocathodic protection performance was studied on 304 SS and Q235 CS in NaCl solution (3.5 wt.%). The two-dimensional Zn3In2S6 loaded on the one-dimensional NiFe2O4 resulted in faster electron migration and enhanced light absorption capability. Moreover, it had been observed through electrochemical testing that the assembly of Zn3In2S6/NiFe2O4 heterojunctions improves the efficacy of photocathodic protection. Following illumination, the self-corrosion potentials of 304 SS and Q235 CS coupled with Zn3In2S6/NiFe2O4 nanocomposites decreased by 1040 mV and 560 mV, and the photoinduced current densities were 1.2 times and 3.9 times greater than the value of Zn3In2S6. Furthermore, the mechanism of enhanced photocathodic protection performance for Zn3In2S6/NiFe2O4 heterojunctions was systematically discussed. XPS and ESR analysis indicated that Zn3In2S6/NiFe2O4 composites follow the Z-scheme electron migration path and retain the stronger reduction and oxidation capacity of Zn3In2S6/NiFe2O4. Therefore, the Z-scheme heterostructures are responsible for the realization of cathodic protection for carbon steel.

1. Introduction

Steel structures are extensively employed in marine and coastal engineering projects owing to their advantages, such as light self-weights, straightforward and rapid construction processes, and recyclability. Consequently, they are a staple in modern construction engineering [1]. Nevertheless, the prevalent use of low-alloy and low-carbon steels in steel structures makes them susceptible to corrosion, leading to a substantial diminution of the steel structures and their load-bearing capacity, as well as compromising their overall stability and reliability and shortening their service lives [2]. Consequently, the development of innovative anti-corrosion technologies is paramount to effectively prolong the lifespan of equipment, minimize maintenance expenditures, ensure operational safety, and mitigate energy wastage. Photocathodic protection technology is a novel protection method for steel structures [3,4].
As we all know, metal corrosion involves the process by which metal atoms lose electrons and transform into metal compounds. The core principle of cathodic protection lies in furnishing negative electrons to the metal under protection, enabling it to undergo oxidation by oxides and thereby substituting for metal corrosion [5]. Photoelectric cathodic protection technology harnesses the unique photoelectric conversion capabilities of semiconductor materials [6]. By absorbing sunlight, a clean energy source, the technology generates photoelectrons to supply electrons to the protected metal. Furthermore, the semiconductor material can be used repeatedly without causing harm, addressing the issues of energy consumption and environmental pollution associated with traditional cathodic protection techniques. This technology represents an economical, energy-efficient, and environmentally friendly approach to corrosion prevention. The efficacy of this technology primarily hinges on the energy level structure and photoelectric conversion efficiency of the photoanode material. In this realm, TiO2 [7,8] and ZnO [9,10], widely used photoanode materials, can play a pivotal role in providing photoelectric cathodic protection for metals with lower corrosion rates, such as stainless steel [11,12]. However, due to their insufficiently negative conduction band positions and weak reduction activity, TiO2 and ZnO photoanodes fail to offer effective cathodic protection to metals prone to corrosion, such as carbon steel [13,14]. Additionally, they belong to the category of wide-bandgap semiconductors that primarily absorb ultraviolet light [15], which is scarce in sunlight. This results in a low utilization rate and photoelectrochemical efficiency of sunlight.
To tackle this challenge, extensive research has been undertaken, encompassing elemental doping [16], morphology control [17,18,19], and heterojunction modification [20,21,22]. Among these strategies, the construction of Z-scheme heterostructures stands out as a particularly effective approach, as it not only enhances the photoelectric conversion efficiency but also boosts the redox activity of the photoanode. Cao, W. et al. [23] fabricated a Bi/BiOBr/TiO2 nanocomposite material and observed that, under visible light illumination, this photoanode could shift the potential of coupled 316L stainless steel negatively to −430 mV versus SCE. Furthermore, based on electron spin resonance characterization of the Bi/BiOBr/TiO2 photoanode, they proposed a charge transfer mechanism for the Z-scheme heterojunction. Chang, Y. et al. [24] successfully synthesized TiO2/In2S3 heterojunction thin films using a one-step hydrothermal method. They discovered that, compared to pure In2S3 and TiO2, the composite photoelectrode composed of TiO2 and In2S3 significantly improved the photocathodic protection performance for 304 SS. This enhancement was attributed to the formation of a Z-scheme heterojunction structure within the In2S3@TiO2/In2S3 composite. Jin, Z.Q. et al. [25] prepared a Z-scheme polydopamine (CdS/PDA) heterostructure through the sample mixing method. Their study demonstrated that this Z-scheme CdS/PDA heterostructure exhibited excellent photoelectrochemical cathodic protection for 304 SS in a NaCl solution. Notably, when PDA was grown for 10 min, it effectively reduced the open-circuit potential of 304 SS to −0.5V. Xu, S.W. et al. [26] developed a series of TiO2/Au/CdS nanostructured coatings. Through active species trapping experiments, they revealed that these coatings exhibited Z-scheme heterojunction properties, enabling efficient charge separation and possessing robust reduction and oxidation capabilities. Zinc indium sulfide (ZnmIn2S3+m) is a typical visible-light-responsive catalyst type among ternary metal sulfides, featuring a similar layered structure and excellent photoelectric conversion performance [27]. Specifically, Zn3In2S6 is a direct bandgap semiconductor that crystallizes in a tetragonal structure with a space group of P3m1. In this structure, Zn atoms occupy tetrahedral sites, while In atoms connect tetrahedra and octahedra. Along the c-axis, it consists of a series of S-In-S-Zn-S-In-S-Zn-S-Zn-S atomic layers, which enhance light utilization and conductivity. Zn3In2S6 is a suitable material for constructing heterojunction structures, with a bandgap width of 2.8 eV and a conduction band position close to −0.9 eV, making it promising for providing protection to metals with lower self-corrosion potentials in photocathodic protection [28]. She et al. [29] combined Zn3In2S6 with a traditional TiO2 material to construct a Z-scheme Zn3In2S6/TiO2 heterojunction material. Fluorescence emission spectroscopy tests showed that the composite material significantly enhanced the transfer ability of photogenerated carriers. Xu et al. [28] prepared a Zn3In2S6/TiO2 nanocomposite material for the photocathodic protection of carbon steel. By loading Zn3In2S6, the light absorption range of TiO2 was extended to the visible light region. The composite material reduced the open-circuit potential of carbon steel by 490 mV under illumination and generated a photogenerated current density of 1.91 mA·cm−2. In addition, there are studies on the construction of other highly reductive heterostructures, such as Zn3In2S6/g-C3N4 [30], Zn3In2S6/BN [31], and polypyrrole/Zn3In2S6 [32]. Additionally, our research found that Zn2In2S5 [33] can serve as a reducing material for constructing a Z-scheme heterojunction. By incorporating Zn2In2S5 into TiO2, we successfully fabricated a zinc indium sulfide/titanium dioxide (Zn2In2S5/TiO2) photoelectrochemical cathodic protection coating. Through theoretical calculations and electrochemical experiments, we discovered that the Z-scheme heterojunction formed between Zn2In2S5 and TiO2 facilitates the separation of photogenerated electrons and holes. Compared to pure Zn2In2S5 and TiO2, the Zn2In2S5/TiO2 composite demonstrates a superior photoelectrochemical protection capability for 304 stainless steel.
NiFe2O4 has become one of the most popular ferrites in the photocatalytic field due to its excellent photostability and energy band structure. With a relatively narrow bandgap (approximately 2.5 eV [34]), NiFe2O4 has a certain light absorption ability in the visible range and good photochemical stability, which enables the broad application of NiFe2O4 in the fields of photocatalysis and photocathodic protection. Consequently, NiFe2O4 has been widely studied in organic dye degradation and visible photocatalysis [35]. However, NiFe2O4 on its own also has the problems of low conductivity [36] and a tendency to agglomerate, which results in a high photogenerated electron–hole pair recombination rate and low photoelectrochemical efficiency in visible light, thereby compromising its photocathodic protection performance.
In this work, the structure of NiFe2O4 was modified into one-dimensional (1D) nanofibers via electrospinning to enhance its electron transport capability and charge separation ability. Subsequently, Zn3In2S6 nanosheets were deposited onto 1D NiFe2O4 fibers through the hydrothermal method to obtain Zn3In2S6/NiFe2O4 heterojunctions. In a 3.5 wt.% NaCl solution, the variations of the open-circuit potential and photocurrent density of the Zn3In2S6/NiFe2O4 composites under light illumination were investigated in comparison with individual Zn3In2S6 and NiFe2O4. The photoelectrochemical cathodic protection effect of the Zn3In2S6/NiFe2O4 composites on 304 SS and Q235 CS was examined. The cathodic protection principle of the Zn3In2S6/NiFe2O4 composites for metals was elucidated, and the formation of the Z-scheme heterojunction between Zn3In2S6 and NiFe2O4 was demonstrated. This provides research ideas for the construction of stable Z-scheme heterojunction materials and their application in the photoelectrochemical cathodic protection of various metals.

2. Materials and Methods

2.1. One-Dimensional NiFe2O4 Nanomaterial Preparation

One-dimensional NiFe2O4 nanofibers were prepared by wet electrostatic spinning. Firstly, 10 mmol/L nickel acetate (C4H6NiO4) (Sinopharm Chemical Reagent Co., Ltd, Shanghai, China) and 20 mmol/L ferric nitrate hydrate (Fe(NO3)3·9H2O) (Macklin, Shanghai, China) were weighed and mixed sequentially in a beaker. After 15min of stirring, a clarified solution was obtained. Then, 10 wt.% poly(vinylpyrrolidone) (Sinopharm Chemical Reagent Co., Ltd, Shanghai, China) was added, and the mixture was stirred continuously for 10 h to form a homogeneous viscous precursor. The precursor solution was loaded into a syringe with a flow rate of 0.5 mL/h. Electrospinning parameters included a voltage of 18 kV, a needle-to-collector distance of 20 cm, and a collector rotation speed of 500 rad/min. After 8 h of spinning, the NiFe2O4 fiber membrane was peeled from the collector and calcined at 600 °C (heating rate: 5 °C/min) for 2 h. This process ultimately yielded one-dimensional NiFe2O4 nanomaterials. The preparation flowchart is shown in Figure 1.

2.2. Zn3In2S6/NiFe2O4 Composite Preparation

Zn3In2S6 was grown on the surface of one-dimensional NiFe2O4 nanomaterials via a hydrothermal method. Firstly, 15 mmol/L zinc sulfate (ZnSO4), 60 mmol/L thioacetamide (C2H5NS) (Sinopharm Chemical Reagent Co., Ltd, Shanghai, China), and 10 mmol/L indium chloride (InCl3) (Macklin, Shanghai, China) were weighed and mixed in a beaker with stirring to obtain the hydrothermal precursor solution of Zn3In2S6. Subsequently, 0.2 g of NiFe2O4 nanofibers were weighed and placed in a 10 mL centrifuge tube, to which 5 mL of H2O was added. The mixture was then subjected to ultrasonic treatment for 30 min to obtain a suspension of NiFe2O4 nanofibers. Afterward, the NiFe2O4 nanofiber suspension was mixed with the hydrothermal precursor solution of Zn3In2S6 and transferred into a 50 mL polytetrafluoroethylene (PTFE) liner. The liner was then placed in a reaction kettle and subjected to hydrothermal treatment at 180 °C for 16 h. The resulting powder sample was dried at 80 °C for 2 h, and the Zn3In2S6/NiFe2O4 nanocomposite was obtained. The preparation flowchart is shown in Figure 1.

2.3. Metal Electrode Preparation

A 1 cm × 1 cm × 0.5 cm block of 304 stainless steel (304 SS) and Q235 carbon steel (Q235 CS) was used as the metal electrode. Each block was sonicated in anhydrous ethanol for 15 min to remove surface contaminants, then dried and welded with a 2 mm diameter copper wire. The welded block was placed in a mold and encapsulated with epoxy resin (mixed at a 3:1 mass ratio) to prevent oxidation. After 48 h of curing, the electrode surface was sequentially polished with 220, 800, 1500, and 2000 mesh sandpaper, followed by 0.5 µm diamond paste to achieve a mirror finish. Finally, the polished electrode was stored in anhydrous ethanol until testing.

2.4. Microscopic Characterization Tests

Scanning electron microscopy (SEM) was performed using a Hitachi-SU8220 scanning electron microscope (Hitachi, Tokyo, Japan) with a tungsten electron source at a 15 kV accelerating voltage. X-ray diffraction (XRD) analysis utilized a Rigaku D/max-2000 (Rigaku, Beijing, China) diffractometer in the range of 10–90° at a scanning rate of 5 °/min. Transmission electron microscopy (TEM, HRTEM) was performed using a JEOL JEM-2100 microscope (JEOL Ltd., Tokyo, Japan)to further analyze the microscopic morphology, material composition, and composite condition of the materials. X-ray photoelectron spectroscopy (XPS) was carried out using a Thermo Scientific ESCALAB 250Xi spectrometer(Thermo Fisher, Waltham, America) under ultra-high vacuum (<1 × 10⁻8 Pa) with K-α radiation (15 kV). Samples underwent broad-spectrum and narrow-spectrum scans, followed by CasaXPS software(Version 2.3.24PR1.0)-based peak deconvolution to determine their elemental composition and oxidation states. Super depth of field microscopy (SDM) tests were performed using a PZ-CS3500 microscope (Pinzhics, Beijing, China) to analyze the degree of coating scratches and flaking.

2.5. Optical Performance Testing

Ultraviolet–visible diffuse reflectance (UV–Vis DRS) was tested using a TU-1901 (Persee, Beijing, China) with the vertical axis set to absorbance intensity (a.u.) and the wavelength range set to 200–1000 nm, and a glass was used to perform a baseline calibration before testing. Photoluminescence spectroscopy (PL) was recorded using a Hitachi F-7000 model fluorescence spectrophotometer (Hitachi, Tokyo, Japan) with a 300 W xenon lamp. The excitation wavelength was fixed at 370 nm and the emission wavelengths were tested in the range of 380–800 nm. The existence of free radicals in the reaction system was captured by ESR testing to determine whether the redox reaction occurred, thereby determining the band potential information of the photoanode and the charge transfer situation and analyzing the reaction mechanism.

2.6. Optical Performance Test

The open-circuit potential test (OCP) was performed using a Gamry Interface 5000E electrochemical workstation (Gamry, Philadelphia, America) with a three-electrode system: an Ag/AgCl reference electrode, a platinum sheet counter electrode, and a working electrode comprising the photoanode connected in parallel to the protected metal to determine the voltage change in the system. The test environment was a 3.5 wt.% NaCl solution and the test light source was a 500 W xenon lamp that irradiated the surface of the photoanode material through the side window of the photodissociation cell in the double electrolysis cell. The protected metal, the platinum sheet electrode, and the reference electrode were placed in the electrolysis cell together, and the ion exchange between the photodissociation cell and the electrolysis cell was carried out through the Nafion proton exchange membrane, with the test light intensity set at 100 mW/cm2. The kinetic potential polarization curve and electrochemical ac impedance (EIS) were measured using the same three-electrode system as the open-circuit potential test, and the kinetic potential polarization curve was tested in the range of −0.8 V~0.8 V curves with respect to the open-circuit potential with a scan rate of 1 mV/s, while for the EIS test, the frequency range of the test was set to 105 Hz~10−2 Hz, and the amplitude of the ac voltage was set to 10 mV. The Mott–Schottky test (MS) was performed using a 3.5 wt.% NaCl solution with a test range of −1 V~0.6 V with respect to the open-circuit potential, and a step height of 10 mV. The connection device for the above test is shown in Figure 2. The photogenerated current density–time test (J-t) was performed in a double electrolytic cell in which the working electrode was connected to the photoanode material, the counter electrode and the ground were connected to the protected metal, and the reference electrodes were also used as the Ag/AgCl electrode. The photoanode material and the protected metal together form the circuit, and the connection device is shown in Figure 2b.

3. Results

3.1. Microscopic Morphology

SEM and HRTEM were used to investigate the microscopic morphologies and compositions of the photoanode materials (Zn3In2S6 (a), NiFe2O4 (b), and the Zn3In2S6/NiFe2O4 composite), as shown in Figure 3. As shown in Figure 3a, the individual Zn3In2S6 is arranged in a staggered manner with a lamellar structure, a sheet thickness of approximately 30 nm, and intricate porosity. Moreover, two-dimensional (2D) Zn3In2S6 has a smooth surface, which is conducive to light reflection, thereby enhancing its light absorption efficiency. Figure 3b illustrates the preparation of NiFe2O4 fibers. The resultant one-dimensional (1D) NiFe2O4 nanofibers possessed a uniform diameter of approximately 120 nm, exhibiting a slightly wrinkled surface. This is attributed to the volatilization of organic components during high-temperature calcination. Figure 3c subsequently presents the SEM image of the Zn3In2S6/NiFe2O4 composite material. As can be observed in the image, the layers of Zn3In2S6 synthesized using the hydrothermal method are lamellar and overlaid on the surface of one-dimensional NiFe2O4, growing in a staggered and compact manner. The Zn3In2S6/NiFe2O4 composite material possessed a diameter ranging from 170 nm to 230 nm. This observation serves as evidence that the Zn3In2S6/NiFe2O4 composites have been successfully fabricated. The large specific surface area of NiFe2O4 fibers offers abundant sites for the growth and attachment of lamellar Zn3In2S6 sheets. Both the 1D nanofibers and the 2D lamellar structures substantially enhanced the electrical conductivity of the composite material, facilitating the rapid transfer of photogenerated electrons. Figure 3d exhibits the HRTEM image of the Zn3In2S6/NiFe2O4 composite material. Within the image, two different lattice spacings were observed, 0.32 nm and 0.25 nm, which corresponded to the (102) crystallographic plane [37] of Zn3In2S6 and the (311) crystallographic plane [38] of NiFe2O4, respectively. This demonstrates the successful preparation of the Zn3In2S6/NiFe2O4 composite material and the intimate contact between Zn3In2S6 and NiFe2O4, a condition conducive to the formation of heterojunctions and the transfer of charges between the materials.

3.2. Material Composition

XRD analysis provides information regarding the crystal surface, as well as the degree of crystallinity of the photoanode material’s crystals. Figure 4a presents the XRD curves of Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4 composite materials, respectively. The sharp diffraction peaks observed in all materials indicated a high degree of crystallinity, while the absence of impurity peaks suggested high sample purity. Among the diffraction peaks observed for Zn3In2S6, there were diffraction peaks at 28.23°, 32.73°, 46.91°, and 55.66°, which correspond to the (102), (014), (110), and (022) crystal planes of Zn3In2S6 (JCPDS NO. 80-0835), respectively, indicating the successful preparation of Zn3In2S6. The diffraction peak curve of NiFe2O4 showed diffraction peaks at 30.30°, 35.69°, 37.33°, 43.38°, 53.83°, 57.38°, and 63.02°, which correspond to the (220), (311), (222), (400), (422), (511), and (440) crystal planes of NiFe2O4 (JCPDS NO. 86-2267), respectively, indicating the successful preparation of NiFe2O4. The XRD curves of the Zn3In2S6/NiFe2O4 composite material exhibit characteristic peaks that correspond to the crystal planes of both Zn3In2S6 and NiFe2O, thereby confirming the successful composition of Zn3In2S6 and NiFe2O4 in the same location. In summary, the X-ray diffraction results demonstrated the successful fabrication of Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4 photoanode materials with enhanced crystallinity and high purity through the experimental process. In addition, the grain sizes D of NiFe2O4 and Zn3In2S6 before and after the composition were estimated using the Debye–Scherrer equation [39,40]:
D = Kλ/βcosθ
where λ = 1.5418 Å for target Cu, K = 0.94, θ is the diffraction angle, and β is the full width at half maximum (FWHM). The calculation results showed that the grain sizes of NiFe2O4 and Zn3In2S6 before compositing were 46.16 nm and 34.65 nm, respectively. After compositing, the grain size of NiFe2O4 decreased to 40.87 nm, while the grain size of Zn3In2O6 increased to 43.99 nm.
XPS is capable of chemically analyzing the material surface to determine the elemental composition and chemical state of the material. Figure 4b shows the spectral peaks of the elements Zn, In, S, Ni, and Fe in Zn3In2S6, as well as in the Zn3In2S6/NiFe2O4 composite samples, demonstrating the successful formation of the composite of the two materials. In addition, the high-resolution XPS scans of the orbitals of the elements C, O, Fe, In, Zn, and S enable the determination of the states of the elements in the material, as well as the electron transfer information. The characteristic curve of C 1s is represented in Figure 4c, and the fitting results showed that the characteristic peak at 284.8 eV was the adventitious carbon reference, and the characteristic peaks at 286.0 eV and 288.5 eV were C-O and C=O [41], respectively. The characteristic curves of O 1s are shown in Figure 4d, where the characteristic peak at 529.7 eV corresponded to O2− ions in NiFe2O4, O-C at 530.8 eV, and O=C at 531.7 eV. The characteristic peaks at 531.7 eV correspond to O=C. Figure 4e shows NiFe2O4 and Zn3In2S6/NiFe2O4, where the peaks at 711 eV and 724 eV correspond to the Fe 2p3/2 and Fe 2p1/2 spin-orbit peaks, respectively. The characteristic peaks of In 3d are shown in Figure 4f, and the characteristic peaks at 444.9 eV and 452.6 eV correspond to In 3d5/2 and In 3d3/2, respectively. In Figure 4g, the Zn 2p3/2 and Zn 3p1/2 orbital peaks are observed at 1021.9 eV and 1045.1 eV, respectively. In Figure 4h, 162.5 eV corresponds to the S 2p3/2 orbital and 163.7 eV to the S 2p1/2 orbital. The combined XRD and XPS results further demonstrate the successful preparation of Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4 composites, and are in agreement with the results obtained for the SEM and HRTEM tests. In addition, from the fine high-resolution XPS spectra of elemental orbitals, it can be found that the binding energies of the elements Fe and In are significantly shifted, indicating the existence of strong electronic interactions and electron transfer between Zn3In2S6 and NiFe2O4, where the binding energy of the Fe element is elevated in Zn3In2S6/NiFe2O4 composites and the binding energy of the In element is decreased in Zn3In2S6/NiFe2O4 composites. This indicates that electron transfer from NiFe2O4 to Zn3In2S6 and aggregation on the surface of Zn3In2S6, suggesting a charge transfer mechanism for Zn3In2S6/NiFe2O4 materials following a Z-scheme heterojunction [42].

3.3. Optical Properties

Most sunlight is visible light, but due to the wide bandgap width, some semiconductors have a weak ability to absorb sunlight, resulting in insufficient photoelectric conversion abilities. The UV–Vis DRS test was conducted to evaluate the light absorption capabilities of Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4. The results of the test are displayed in Figure 5a. From the figure, it was observed that the single materials Zn3In2S6 and NiFe2O4 exhibited superior absorption capabilities below 475 nm and 596 nm, respectively. After NiFe2O4 was composited onto Zn3In2S6, the optical absorption threshold of Zn3In2S6 is red-shifted and broadened to 630 nm. This indicates that the heterojunction resulting from the contact between the two materials has the ability to modify the band structure of the composite, narrowing the forbidden band width of Zn3In2S6 and enhancing its optical absorption performance. Consequently, photons with lower energy can be absorbed, sunlight can be better utilized, and photoelectric conversion performance can be further enhanced, all of which are beneficial for enhancing photoelectric cathodic protection performance.
Using the Kubelka–Munk equation [43], the value of the forbidden bandwidth corresponding to the photoanode can be calculated from the UV–visible diffuse reflectance spectrum:
( α h ν ) 1 / n = A ( h ν E g )
where α is the light absorption coefficient, h is Planck’s constant,   ν is the frequency of light, A is a constant, Eg is the band gap of the semiconductor, and n represents the type of electron transition in the semiconductor (n = 1/2 when the material is a direct bandgap semiconductor; n = 2 when the material is an indirect bandgap semiconductor). From Figure 5b, the band gaps of Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4 are obtained as 2.72 eV, 2.41 eV, and 2.32 eV, respectively.
The photoluminescence (PL) intensity of the photoanode material was measured to verify the effect of the Zn3In2S6/NiFe2O4 composite on the separation efficiency of photogenerated electron–hole pairs. Due to the inefficient separation of the photogenerated electron–hole pairs, the electrons and holes compounded on the semiconductor surface after excitation will produce the luminescence or heat phenomenon. Therefore, in the fluorescence spectrum, the peak intensity represents the photoluminescence intensity. The material exhibiting a lower photoluminescence intensity indicates that its photogenerated electron–hole composite rate is lower. From Figure 3c, it can be seen that the strongest fluorescence peaks of NiFe2O4 and Zn3In2S6/NiFe2O4 appear at 456 nm and 453 nm, respectively. The red-shift of the strongest peak position of Zn3In2S6/NiFe2O4 indicates a reduced band gap after material compositing. The highest photoluminescence intensity in NiFe2O4 indicates its poor charge separation efficiency, while limited electron transfer further restricts its protective performance. On the other hand, Zn3In2S6/NiFe2O4 has a lower photoluminescence intensity, indicating that the heterojunction formed between Zn3In2S6 and NiFe2O4 can promote the separation between photogenerated electron–holes pairs and enhance the photoelectric conversion performance of the material.
Based on the results of the Mott–Schottky curves, further analysis was conducted on the energy band structure of the photoanode materials and the information on the variation of the concentration of photogenerated carriers. As can be seen in Figure 5d, the tangent slopes of Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4 are all greater than 0 within the tested range, exhibiting the characteristics of n-type semiconductors. The intersection of the curve’s approximately straight segment with the abscissa axis denotes the flat band potentials [44] of the material. For n-type semiconductors, the flat band potential is close to its conduction band position. Generally, the conduction band potential is approximately 0.1 eV~0.3 eV more negative than the flat band potential. Table 1 shows the flat band potentials of Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4 are −0.80 eV, −0.28 eV, and −0.63 eV, respectively. The valence band positions can be calculated to be 1.92 eV, 2.13 eV, and 1.69 eV, respectively. The obtained energy band structure information is summarized in Table 1. The positions of the flat band potentials of Zn3In2S6/NiFe2O4 are close to Zn3In2S6, showing the characteristics of Z-scheme semiconductors. From the change in the flat band potential, it can be observed that the doping of Zn3In2S6 reduced the flat band potential of NiFe2O4, indicating that the Zn3In2S6/NiFe2O4 composite has a more negative conduction band potential compared to NiFe2O4 alone. The greater negative conduction band potential provides a stronger electronic driving force for electrons transferred to the protected metal. Thus, the negative shift of the flat band potential indicates that the doping of Zn3In2S6 can effectively improve the electron transfer performance of NiFe2O4. In addition, the carrier concentration can be determined from the slope of the tangent. The slope of the tangent is inversely proportional to the carrier concentration. A smaller tangent slope indicates a higher carrier density. In Figure 5d, the Zn3In2S6/NiFe2O4 composite has the smallest slope of the straight-line segment, and therefore the highest carrier concentration. This further illustrates that the Zn3In2S6/NiFe2O4 composite improves the separation efficiency of the photogenerated electron–hole pairs and can provide more electrons to the protected metal.

3.4. Photochemical Protection Performance and Mechanism Analysis

Since photocathodic protection is based on the transfer and accumulation of photogenerated electrons on the metal surface, OCP (open-circuit potential) and J-t (current–time) tests are used to analyze the potential changes on the surface of different coupled metal electrodes and the magnitude of the generated photogenerated current to evaluate the photocathodic protection performance of composite photoanode materials. Figure 6 shows the open-circuit potential changes on the metal electrode surface under illuminated and dark conditions for NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites with 304 SS and Q235 CS in a simulated marine environment. In Figure 6a, upon illumination, due to the formation of a Schottky barrier between the photoanode material and the metal, the electrons transfer and accumulate, causing polarization, which is manifested as a rapid negative shift of the open-circuit potential. It can be seen that the open-circuit potential of 304 SS coupled with NiFe2O4, Zn3In2S6, and Zn3In2S 6/NiFe2O4 composite materials under dark conditions is −40 mV, −160 mV, and −60 mV, respectively. After illumination, the open-circuit potentials of the coupled metals decrease to −50 mV, −1050 mV, and −1100 mV, respectively, with potential negative shifts of 10 mV, 890 mV, and 1040 mV, respectively. The potential of the coupling metal with the Zn3 In2S6/NiFe2O4 composite coupled with 304 SS exhibits the largest negative shift in open-circuit potential, indicating that more photogenerated electrons are accumulated on its surface. This suggests that when Zn3In2S6 is composited with NiFe2O4, the composite enhances the light absorption range and improves the material’s ability to utilize light due to the semiconductor’s composition. At the same time, the heterojunction formed between the two heterojunctions improves the separation ability of photogenerated electron–hole pairs, reduces the photogenerated electron–hole recombination of the semiconductor material itself, and enables more electrons to gather on the surface of the protected metal.
Figure 6b shows the photogenerated current density generated by the photoanode under illuminated and dark conditions, further illustrating the charge separation and transfer properties of the photoanode material. It can be observed that under dark conditions, at the equilibrium of the Fermi level between the coupled metal and photoanode, almost no current is generated between the electrodes. Upon illumination, the photocurrent density between the electrodes increases rapidly. However, due to the material’s own defects, the photogenerated current density of the photoanode material begins to decline, eventually stabilizing after a period of time and generating a sharp peak. When 304 SS is coupled with NiFe2O4, the photogenerated current density between the two is approximately 0.8 μA·cm−2, and when coupled with Zn3In2S6, the photogenerated current density reaches approximately 40 μA·cm−2, which suggests that under illumination, both materials are capable of generating photogenerated electron flows towards the metal surface, providing metal protection. In addition, when coupled with Zn3In2S6/NiFe2O4, a photogenerated current density of approximately 50 μA·cm−2 can be generated between the metal and the photoanode, which is 62.5 times and 1.2 times higher than that of the coupling with NiFe2O4 and Zn3In2S6 alone, respectively. This further demonstrates that the composite of the photoanode materials formed a heterojunction that can improve the separation efficiency of photogenerated electron–hole pairs and provide more protective electrons for metals.
Similarly, Figure 6c shows the OCP of Q235 CS coupled with NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites, respectively. Under dark conditions, the OCP of Q235CS surfaces coupled with NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites is −610 mV, −650 mV, and −650 mV, respectively. Upon illumination, the potentials change to −580 mV and −1030 mV. The potential drops are −30 mV, 380 mV, and 560 mV, respectively. Among them, the OCP of Q235 CS coupled with NiFe2O4 increases upon illumination, suggesting that the conduction band position is insufficient to provide protective properties to the Q235 CS with a more negative Fermi energy level. The Zn3In2S6/NiFe2O4 composites still have the largest potential drop, indicating that instead of forming a conventional Type II heterojunction between the heterojunction materials, a Z-scheme heterojunction with a higher redox capacity is formed, which allows the photogenerated electrons passing through the heterojunction to accumulate on the Zn3In2S6 conduction band with a stronger electronic driving force and transfer to the surface of the protected metal.
Figure 6b,d show the magnitude of photogenerated current densities between Q235 CS and NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 under illumination. Upon illumination, photogenerated currents of −3 μA·cm−2, 9 μA·cm−2, and 35 μA·cm−2 were generated between NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 and Q235 CS, respectively. Among them, NiFe2O4 does not transfer the photocurrent to the surface of Q235 CS. The photogenerated current density generated by Zn3In2S6/NiFe2O4 was 3.9 times that of Zn3In2S6, demonstrating that Zn3In2S6/NiFe2O4 for Q235 CS also has better cathodic protection performance.
The corrosion potential (Ecorr)/coupling potential (Ecoupler) and corrosion current density (Icorr)/coupling current density (Icoupler) of 304 SS or Q235 CS coupled with NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 can also be determined using dynamic polarization curves and can likewise determine the photocathodic protection capability of the photoanode material. From Figure 7a,b, it can be seen that under dark conditions, the coupling potentials of 304 SS with NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 in the 3.5 wt.% NaCl solution are all near the self-corrosion potentials of 304 SS, at −70 mV, −120 mV, and −210 mV, respectively, which has minimal effects on the corrosion rate of 304 SS. The small variation of the coupling potentials indicates the rebalancing of the Fermi energy levels after the metal is coupled with the photoanodes. Under illumination, the coupling potential of 304 SS with NiFe2O4 reaches −90 mV, decreasing by 20 mV, with a coupling current density that reaches 3.98 μA·cm−2; the coupling potential of 304 SS with Zn3In2S6 reaches −620 mV, decreasing by 410 mV, with a coupling current density that reaches 5.30 μA·cm−2; and the coupling potential of 304 SS with Zn3In2S6/NiFe2O4 reaches −780 mV, decreasing by 660 mV, with a coupling current density that reaches 10 μA·cm−2. Under the influence of light, the photogenerated electrons are enriched on the metal surface, resulting in a decrease in the coupling potential. During the electron migration process, the coupling current density represents the photocurrent’s magnitude. The increase in the coupling current density indicates an increase in the electron concentration during the charge transfer process, which corresponds to the results of the OCP and the photogenerated current densities. Due to limitations in their light utilization and carrier separation efficiencies, NiFe2O4 and Zn3In2S6 alone exhibit relatively weaker coupling potential drops and coupling current densities. The Zn3In2S6/NiFe2O4 composite photoanode with 304 SS shows the most negative coupling potential and the largest coupling current density, which indicates that its metal surface possesses the largest number of photogenerated electrons, improving its photogenerated carrier separation ability and providing the best photocathodic protection performance.
Figure 7c,d present the corrosion potential/coupling potential and the corrosion current density/coupling current density of Q235 CS coupled with NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 in a 3.5 wt.% NaCl solution. The coupling potential under dark conditions indicates that the photoanode material exhibited no significant protective effect on Q235 CS. Under illumination, the coupling potential of Q235 CS with NiFe2O4 reached −670 mV, increasing by 10 mV, and the coupling current density reached 8.33 μA·cm−2; the coupling potential of Q235 CS with Zn3In2S6 reached −960 mV, with a decrease of 260 mV, and the coupling current density reached 16.81 μA·cm−2; the coupling potential of Q235 CS with Zn3In2S6/NiFe2O4 reached −1140 mV, with a decrease of 420 mV, and the coupling current density reached 32.48 μA·cm−2. The composite exhibited the most negative coupling potential and the highest coupling current density when coupled with Q235 CS, which demonstrated that the composite had the best cathodic protection performance, consistent with the OCP and photogenerated current density. All corrosion potential/coupling potential and corrosion current density/coupling current density data are listed in Table 2 and Table 3.
As one of the most commonly used electrochemical test methods, EIS is employed to investigate the charge transfer resistance of semiconductor materials, thereby demonstrating their charge transfer performance. In this context, the dotted and solid lines represent the original and fitted results, respectively. In the fitted circuit, Rc denotes the surface film resistance, with CPEc being its capacitance. Rct represents the charge transfer resistance, CPEdl represents the capacitance of the double layer, and Rs is the solution resistance. The charge transfer resistance can be judged by comparing the curve diameters; in general, as the curve diameter decreases, the charge transfer resistance also decreases accordingly.
Figure 8a,b illustrate the variation of charge transfer performance of Zn3In2S6 loaded with NiFe2O4 in a 3.5 wt.% NaCl solution under dark and illuminated conditions, respectively. From the figure, it can be observed that NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 all have higher impedance values under dark conditions, which indicates lower charge transfer capabilities. Under illumination, the impedance values of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 decrease significantly, indicating that the light excitation induces electrons to absorb energy at higher energy-excited states, enhancing their activity and facilitating charge transfer. Importantly, Zn3In2S6/NiFe2O4 has the smallest charge transfer resistance, thus indicating that the surface electrons of Zn3In2S6/NiFe2O4 are more prone to transfer and have the optimal electron transfer performance. This is because the Zn3In2S6/NiFe2O4 composite interface forms a built-in electric field, which inhibits the photogenerated electrons from the excited state to the ground state complexed with holes, thereby increasing the concentration of photogenerated electrons and making electron transfer more likely to occur. Thus, more electrons can be transferred to the surface of the protected metal when coupled with the metal, indicating that Zn3In 2S6/NiFe2O4 has better cathodic protection performance compared to a single material. The fitted values in the electrochemical impedance spectra are listed in Table 4.
In order to gain insight into the charge transfer mode between the materials, the heterojunction mechanism of Zn3In2S6/NiFe2O4 can be verified by discussing the relationship between the catalytic potentials of the radicals and the conduction band and valence bands of the photoanode in the ESR spectrum. As shown in Figure 9a, the characteristic curves of ·O2− appear in both Zn3In2S6 and Zn3In2S6/NiFe2O4 reactive systems, indicating the formation of ·O2− radicals in both systems. However, no ·O2− radicals were generated in the NiFe2O4 system. This is because the O2/·O2 potential in the reaction system is approximately −0.33 V (vs. NHE) [45], and the conduction band potentials of Zn3In2S6 and Zn3In2S6/NiFe2O4 are −0.80 and −0.63 eV, respectively, both of which are lower than the generation potential of ·O2−. Therefore, ·O2− is produced in the system. In contrast, the conduction band potential of NiFe2O4 is only −0.28 eV, which is insufficient to reduce O2 to generate ·O2−. In addition, the characteristic curve of Zn3In2S6/NiFe2O4 showed the highest intensity in the figure, indicating that due to the effect of the heterojunction, electrons accumulate on a certain material, thereby increasing the charge concentration. If the system conforms to the conventional Type II heterojunction, electrons will accumulate on the conduction band of NiFe2O4. However, the conduction band reduction of NiFe2O4 is insufficient, and the composite material is unlikely to produce the strongest peak in the characteristic curve. Therefore, the system can only follow the transfer law of the Z-scheme heterojunction, where electrons accumulating on Zn3In2S6 exhibit sufficient reducing ability.
In Figure 9b, it can be seen that ·OH radicals appear in both the NiFe2O4 and Zn3In2S6/NiFe2O4 reaction systems, but are absent in Zn3In2S6. This is because the valence band position of NiFe2O4 is 2.13 eV and exceeds that of ·OH/OH (1.99 eV vs. NHE) [46]. Therefore, the holes in the valence band of NiFe2O4 are able to oxidize OH in the environment to produce ·OH. The formation of the heterojunction between Zn3In2S6/NiFe2O4 leads to photogenerated holes accumulating in the valence band of NiFe2O4, which has a strong oxidizing capacity, leading to ·OH appearing in the reaction system. In contrast, due to the lower position of the valence band and the weaker oxidizing ability of Zn3In2S6, no ·OH is generated. The highest peak intensity in the Zn3In2S6/NiFe2O4 system indicates that the accumulation of photogenerated holes occurs in the Zn3In2S6/NiFe2O4 composite material. If the conventional Type II heterojunction structure was followed, the photogenerated holes would ultimately be accumulated in the valence band of Zn3In2S6, which is not consistent with the fact that Zn3In2S6/NiFe2O4 has the strongest peak intensity. Therefore, the Zn3In2S6/NiFe2O4 composite material should retain the photogenerated electrons in the conduction band of Zn3In2S6, and the photogenerated holes should be transferred to the valence band of NiFe2O4, which is consistent with the Z-scheme heterojunction’s charge transfer mechanism.
Figure 10 shows the charge transfer mechanism of the Zn3In2S6/NiFe2O4 Z-scheme heterojunction. Z-scheme heterojunctions differ from conventional Type II heterojunctions in that Z-scheme heterojunctions not only enhance the photogenerated electron–hole separation capability of the Zn3In2S6/NiFe2O4 composite material but also retain the electron-driving force of the semiconductor with a more negative conduction band potential and the valence band oxidation capacity of semiconductors with a higher valence band potential. This mechanism enables efficient photocathodic protection for various metals with lower self-corrosion potentials in simulated marine environments.

4. Discussion

This paper studies the construction of a Z-scheme heterojunction by modifying one-dimensional NiFe2O4 with Zn3In2S6 to enhance the photoelectrochemical cathodic protection ability of the composite material, enabling it to have a significant protective effect on Q235 carbon steel.
(1)
Two-dimensional Zn3In2S6 interlaces with and grows on the surface of one-dimensional NiFe2O4 nanofibers, and lattice information corresponding to the NiFe2O4 (311) crystal faces and Zn3In2S6 (102) crystal faces can be clearly identified. Furthermore, the diffraction peaks of NiFe2O4 and Zn3In2S6 were both found in the XRD curve of Zn3In2S6/NiFe2O4. The Zn3In2S6/NiFe2O4 composite material was successfully prepared.
(2)
The construction of the energy band of the Zn3In2S6/NiFe2O4 composite material led to optical absorption thresholds that had been broadened to 630 nm. Its valence band was 1.69 eV and its conduction band was −0.63 eV. Furthermore, the Zn3In2S6/NiFe2O4 composite material has the lowest photoluminescence intensity and charge transfer resistance. It possessed a favorable light utilization capability and an excellent capacity for the separation and transfer of photogenerated electrons.
(3)
Compared with Zn3In2S6 and NiFe2O4, the Zn3In2S6/NiFe2O4 composite material coupled with both of the metals demonstrates superior photocathodic protection performance. When coupled with 304 SS, the open-circuit potential dropped to 1040 mV, and the photogenerated current density was 50 μA·cm−2. As for Q235 CS, the open-circuit potential had dropped to 560 mV, and the photogenerated current density was 35 μA·cm−2.
(4)
The charge transfer pathway between Zn3In2S6 and NiFe2O4 was consistent with the mechanism of a Z-scheme heterojunction. The photogenerated electrons and holes were concentrated in the conduction band of Zn3In2S6 and the valence band of NiFe2O4, respectively. This charge transfer mechanism preserves the strong redox capability of photogenerated carriers, thereby providing photocathodic protection for metals with low self-corrosion potentials, such as Q235 CS.
The ferroelectric properties of NiFe2O4 may cause its conductivity to be greatly affected by temperature. In future work, temperature can be taken as an influencing factor to explore the effect of temperature differences on the photoelectrochemical protection performance of NiFe2O4-based Z-scheme heterojunction materials.

5. Conclusions

This paper focuses on metal protection in marine corrosion environments, constructing a stable Z-scheme Zn3In2S6/NiFe2O4 nanocomposite based on NiFe2O4 and exploring its protective performance and charge transfer mechanism for Q235 CS. In this study, the one-dimensional NiFe2O4 modified by the electrospinning method led to an improved charge transfer capability, promoting the formation of a stable Z-scheme heterostructure between Zn3In2S6 and NiFe2O4. The formation of the Z-scheme heterojunction effectively enhanced the light absorption capacity of NiFe2O4 and increased the separation efficiency of photogenerated carriers. In addition, the highly reductive photogenerated electrons in the Z-scheme heterojunction of the Zn3In2S6/NiFe2O4 composite could provide effective protection for Q235 CS. The conclusion of this paper provides a new example for the industrial application of stable Z-scheme heterostructure construction and photoelectrochemical cathodic protection.

Author Contributions

Conceptualization, X.Z. and Y.C.; methodology, X.Z.; validation, Y.C. and X.W.; formal analysis, X.W. and Y.C.; resources, X.Z.; data curation, Y.C.; writing-original draft preparation, X.W.; writing-review and editing, X.W. and Y.C.; supervision, X.Z.; project administration, X.Z.; funding acquisition, X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, with the grant number 52479126.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. El-Etre, A.Y. Khillah extract as inhibitor for acid corrosion of SX 316 steel. Appl. Surf. Sci. 2006, 252, 8521–8525. [Google Scholar] [CrossRef]
  2. Chen, W.; Wang, Z.; Xu, G.; Song, W.; Xie, Y.; Zhao, L.; Xia, M.-H.; Li, W. Friction and anti-corrosion characteristics of arc sprayed Al+Zn coatings on steel structures prepared in atmospheric environment. J. Mater. Res. Technol. 2021, 15, 6562–6573. [Google Scholar] [CrossRef]
  3. Momeni, M.M.; Zeinali, P. Fabrication of Ag electrodeposited-iron doped TiO2 nanotube composites for photoelectrochemical cathodic protection applications. J. Electroanal. Chem. 2021, 891, 115283. [Google Scholar] [CrossRef]
  4. Zhang, X.; Li, X.; Wang, X.; Jin, Z. High-efficiency dual protection of photoelectrochemical cathodic protection coating with In2O3/C3N4/CNT nanocomposites for 304 stainless steel in marine environment. Constr. Build. Mater. 2024, 453, 139096. [Google Scholar] [CrossRef]
  5. Kandi, D.; Martha, S.; Thirumurugan, A.; Parida, K.M. Modification of BiOI Microplates with CdS QDs for Enhancing Stability, Optical Property, Electronic Behavior toward Rhodamine B Decolorization, and Photocatalytic Hydrogen Evolution. J. Phys. Chem. C 2017, 121, 4834–4849. [Google Scholar] [CrossRef]
  6. Xu, S.; Zhang, Z.; Zhang, H.; Li, Z.; Peng, H.; Kawi, S.; Xie, A.; Qiu, P. In-situ fabrication of TiO2/ZIF-67 composite film with electron storage characteristics for significantly enhanced photoelectrochemical cathodic protection of 304 stainless steel. Chem. Eng. J. 2025, 503, 158693. [Google Scholar] [CrossRef]
  7. Hosseini, S.F.; Seyed Dorraji, M.S.; Rasoulifard, M.H. Boosting photo-charge transfer in 3D/2D TiO2@Ti3C2 MXene/Bi2S3 Schottky/Z-scheme heterojunction for photocatalytic antibiotic degradation and H2 evolution. Compos. Part B Eng. 2023, 262, 110820. [Google Scholar] [CrossRef]
  8. Jiang, A.; Di, Y.; Chen, S.; Zhang, D.; Chen, X.; Zhang, Z.; Zhang, X.; Dong, Q. Photocathodic protection of 304 stainless steel by coating muscovite/TiO2 heterostructure. Appl. Clay Sci. 2023, 240, 106974. [Google Scholar] [CrossRef]
  9. Liu, Y.; Zhu, Z.; Cheng, Y. An in-depth study of photocathodic protection of SS304 steel by electrodeposited layers of ZnO nanoparticles. Surf. Coat. Technol. 2020, 399, 126158. [Google Scholar] [CrossRef]
  10. Yang, Y.; Cheng, Y.F. One-step facile preparation of ZnO nanorods as high-performance photoanodes for photoelectrochemical cathodic protection. Electrochim. Acta 2018, 276, 311–318. [Google Scholar] [CrossRef]
  11. Bu, Y.; Chen, Z.; Ao, J.; Hou, J.; Sun, M. Study of the photoelectrochemical cathodic protection mechanism for steel based on the SrTiO3-TiO2 composite. J. Alloys Compd. 2018, 731, 1214–1224. [Google Scholar] [CrossRef]
  12. Cui, J.; Pei, Y. Enhanced photocathodic protection performance of Fe2O3/TiO2 heterojunction for carbon steel under simulated solar light. J. Alloys Compd. 2019, 779, 183–192. [Google Scholar] [CrossRef]
  13. Wang, W.; Ye, Y.; Li, G.; Yang, Z.; Duan, J.; Sun, J.; Yan, Y. High-efficiency photocathodic protection performance of novel MnIn2S4/TiO2 n-n heterojunction films for Q235 carbon steel in chloride-containing simulated concrete pore solution. J. Alloys Compd. 2023, 941, 168957. [Google Scholar] [CrossRef]
  14. Ren, J.; Qian, B.; Li, J.; Song, Z.; Hao, L.; Shi, J. Highly efficient polypyrrole sensitized TiO2 nanotube films for photocathodic protection of Q235 carbon steel. Corros. Sci. 2016, 111, 596–601. [Google Scholar] [CrossRef]
  15. Li, H.; Wang, X.; Liu, Y.; Hou, B. Ag and SnO2 co-sensitized TiO2 photoanodes for protection of 304 SS under visible light. Corros. Sci. 2014, 82, 145–153. [Google Scholar] [CrossRef]
  16. Wan, L.; Gao, Y.; Xia, X.-H.; Deng, Q.-R.; Shao, G. Phase selection and visible light photo-catalytic activity of Fe-doped TiO2 prepared by the hydrothermal method. Mater. Res. Bull. 2011, 46, 442–446. [Google Scholar] [CrossRef]
  17. Lei, C.-X.; Zhou, H.; Wang, C.; Feng, Z.-D. Self-assembly of ordered mesoporous TiO2 thin films as photoanodes for cathodic protection of stainless steel. Electrochim. Acta 2013, 87, 245–249. [Google Scholar] [CrossRef]
  18. Kong, C.; Su, X.; Qing, D.; Zhao, Y.; Wang, J.; Zeng, X. Controlled synthesis of various SrTiO3 morphologies and their effects on photoelectrochemical cathodic protection performance. Ceram. Int. 2022, 48, 20228–20236. [Google Scholar] [CrossRef]
  19. Deng, H.; Huang, M.-C.; Weng, W.-H.; Lin, J.-C. Photocathodic protection of iron oxide nanotube arrays fabricated on carbon steel. Surf. Coat. Technol. 2015, 266, 183–187. [Google Scholar] [CrossRef]
  20. Lei, J.; Shao, Q.; Wang, X.; Wei, Q.; Yang, L.; Li, H.; Huang, Y.; Hou, B. ZnFe2O4/TiO2 nanocomposite films for photocathodic protection of 304 stainless steel under visible light. Mater. Res. Bull. 2017, 95, 253–260. [Google Scholar] [CrossRef]
  21. 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]
  22. Li, Z.; Jin, Z.; Jiang, H.; Cai, J.; Cao, Y.; Zhang, X.; Hou, B. Manganese indium sulfide/indium oxide with S-scheme heterojunction for enhanced dual performance of metal in marine environment. J. Clean. Prod. 2025, 490, 144785. [Google Scholar] [CrossRef]
  23. Cao, W.; Wang, W.; Yang, Z.; Wang, W.; Chen, W.; Wu, K. Enhancing photocathodic protection performance by controlled synthesis of Bi/BiOBr/TiO2 NTAs Z-scheme heterojunction films. J. Alloys Compd. 2023, 960, 170675. [Google Scholar] [CrossRef]
  24. Chang, Y.; Suo, K.; Wang, Y.; Ren, X.; Cao, J. In2S3@TiO2/In2S3 Z-Scheme Heterojunction with Synergistic Effect for Enhanced Photocathodic Protection of Steel. Molecules 2023, 28, 6554. [Google Scholar] [CrossRef]
  25. Jin, Z.; Liu, Y.; Jiang, H.; Zhang, X. Enhanced photocathodic protection on 304 SS in marine environment by Z-Scheme polydopamine-modified CdS. Constr. Build. Mater. 2024, 416, 135216. [Google Scholar] [CrossRef]
  26. Xu, S.; Li, Z.; Wen, J.; Qiu, P.; Xie, A.; Peng, H. Review of TiO2-Based Heterojunction Coatings in Photocathodic Protection. ACS Appl. Nano Mater. 2024, 7, 8464–8488. [Google Scholar] [CrossRef]
  27. Deng, F.; Peng, J.; Li, X.; Luo, X.; Ganguly, P.; Pillai, S.C.; Ren, B.; Ding, L.; Dionysiou, D.D. Metal sulfide-based Z-scheme heterojunctions in photocatalytic removal of contaminants, H2 evolution and CO2 reduction: Current status and future perspectives. J. Clean. Prod. 2023, 416, 137957. [Google Scholar] [CrossRef]
  28. Xu, S.; Zhang, K.; Du, L.; Gong, D.; Lu, G.; Qiu, P. Zn3In2S6/TiO2 Nanocomposites for Highly Efficient Photocathodic Protection to Carbon Steel. ACS Appl. Nano Mater. 2022, 5, 18297–18306. [Google Scholar] [CrossRef]
  29. She, H.; Wang, Y.; Zhou, H.; Li, Y.; Wang, L.; Huang, J.; Wang, Q. Preparation of Zn3In2S6/TiO2 for Enhanced CO2 Photocatalytic Reduction Activity Via Z-scheme Electron Transfer. ChemCatChem 2018, 11, 753–759. [Google Scholar] [CrossRef]
  30. Ni, L.; Xiao, Y.; Zhou, X.; Jiang, Y.; Liu, Y.; Zhang, W.; Zhang, J.; Liu, Z. Significantly Enhanced Photocatalytic Performance of the g-C3N4/Sulfur-Vacancy-Containing Zn3In2S6 Heterostructure for Photocatalytic H2 and H2O2 Generation by Coupling Defects with Heterojunction Engineering. Inorg. Chem. 2022, 61, 19552–19566. [Google Scholar] [CrossRef]
  31. Fan, Q.-Q.; Niu, C.-G.; Guo, H.; Huang, D.-W.; Dong, Z.-T.; Yang, Y.-Y.; Liu, H.-Y.; Li, L.; Qin, M.-Z. Insights into the role of reactive oxygen species in photocatalytic H2O2 generation and OTC removal over a novel BN/Zn3In2S6 heterojunction. J. Hazard. Mater. 2022, 438, 129483. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, D.; Xu, Y.; Jing, L.; Xie, M.; Song, Y.; Xu, H.; Li, H.; Xie, J. In situ construction efficient visible-light-driven three-dimensional Polypyrrole/Zn3In2S6 nanoflower to systematically explore the photoreduction of Cr(VI): Performance, factors and mechanism. J. Hazard. Mater. 2020, 384, 121480. [Google Scholar] [CrossRef] [PubMed]
  33. Li, Z.; Zhang, X.; Jin, Z.; Jiang, H.; Wang, X.; Chen, Y.; Zhang, Y.; Hou, B. Enhanced anticorrosion mechanism of photocathodic protection coatings with zinc indium sulfide/titanium oxide Z-scheme heterojunctions. Constr. Build. Mater. 2024, 453, 139109. [Google Scholar] [CrossRef]
  34. Tan, D.D.; Bi, D.F.; Shi, P.H.; Xu, S.H. Preparation and Photocatalytic Property of TiO2/NiFe2O4Composite Photocatalysts. Adv. Mater. Res. 2012, 518–523, 775–779. [Google Scholar] [CrossRef]
  35. Jiang, Z.; Chen, L.; Shi, N.; Ji, L. Synthesis and photocatalytic properties of BiOI/NiFe2O4 composites. Mater. Res. Express 2019, 6, 066207. [Google Scholar] [CrossRef]
  36. Hariharasuthan, R.; Chitradevi, S.; Radha, K.S.; Chithambaram, V. Characterization of NiFe2O4 (Nickel Ferrite) nanoparticles with very low magnetic saturation synthesized via co-precipitation method. Appl. Phys. A 2022, 128, 1045. [Google Scholar] [CrossRef]
  37. Xiong, S.; Liu, X.; Shen, Z.; Hu, Z.; Yang, H.; Hao, J.; Cai, J.; Yang, J. Selective Oxidation of 5-Hydroxymethyl Furfural Coupled with H2 Production over Surface Sulfur Vacancy-Rich Zn3In2S6/Bi2MoO6 Heterojunction Photocatalyst. Inorg. Chem. 2023, 62, 20120–20128. [Google Scholar] [CrossRef]
  38. Wang, J.; Yang, G.; Wang, L.; Yan, W. Synthesis of one-dimensional NiFe2O4 nanostructures: Tunable morphology and high-performance anode materials for Li ion batteries. J. Mater. Chem. A 2016, 4, 8620–8629. [Google Scholar] [CrossRef]
  39. Boubaker, K. Optimization of ZnO sheets dimension in terms of ductility, micro-indentation, mechanical resistance, Amlouk-Boubaker optothermal expansivity and crystallites size. Mater. Sci. Eng. A 2011, 528, 1455–1457. [Google Scholar] [CrossRef]
  40. Atkins, E. Elements of X-Ray Diffraction. Phys. Bull. 1978, 29, 572. [Google Scholar] [CrossRef]
  41. Thomas, A.; Fischer, A.; Goettmann, F.; Antonietti, M.; Müller, J.-O.; Schlögl, R.; Carlsson, J.M. Graphitic carbon nitride materials: Variation of structure and morphology and their use as metal-free catalysts. J. Mater. Chem. 2008, 18, 4893–4908. [Google Scholar] [CrossRef]
  42. Liu, H.; Sun, F.; Li, X.; Ma, Q.; Liu, G.; Yu, H.; Yu, W.; Dong, X.; Su, Z. g-C3N4/TiO2/ZnIn2S4 graphene aerogel photocatalysts with double S-scheme heterostructure for improving photocatalytic multifunctional performances. Compos. Part B Eng. 2023, 259, 110746. [Google Scholar] [CrossRef]
  43. Butler, M.A. Photoelectrolysis and physical properties of the semiconducting electrode WO2. J. Appl. Phys. 1977, 48, 1914–1920. [Google Scholar] [CrossRef]
  44. Jing, J.; Chen, Z.; Bu, Y.; Sun, M.; Zheng, W.; Li, W. Significantly enhanced photoelectrochemical cathodic protection performance of hydrogen treated Cr-doped SrTiO3 by Cr6+ reduction and oxygen vacancy modification. Electrochim. Acta 2019, 304, 386–395. [Google Scholar] [CrossRef]
  45. Li, B.; Lai, C.; Zeng, G.; Qin, L.; Yi, H.; Huang, D.; Zhou, C.; Liu, X.; Cheng, M.; Xu, P.; et al. Facile Hydrothermal Synthesis of Z-Scheme Bi2Fe4O9/Bi2WO6 Heterojunction Photocatalyst with Enhanced Visible Light Photocatalytic Activity. ACS Appl. Mater. Interfaces 2018, 10, 18824–18836. [Google Scholar] [CrossRef]
  46. Dong, F.; Wang, Z.; Li, Y.; Ho, W.-K.; Lee, S.C. Immobilization of Polymeric g-C3N4 on Structured Ceramic Foam for Efficient Visible Light Photocatalytic Air Purification with Real Indoor Illumination. Environ. Sci. Technol. 2014, 48, 10345–10353. [Google Scholar] [CrossRef]
Figure 1. NiFe2O4 and Zn3In2S6/NiFe2O4 preparation process.
Figure 1. NiFe2O4 and Zn3In2S6/NiFe2O4 preparation process.
Buildings 15 00958 g001
Figure 2. The connection device for the OCP, EIS, and MS tests (a). The connection device for the J-t test (b).
Figure 2. The connection device for the OCP, EIS, and MS tests (a). The connection device for the J-t test (b).
Buildings 15 00958 g002
Figure 3. Scanning electron microscope images of Zn3In2S6 (a), NiFe2O4 (b), and Zn3In2S6/NiFe2O4 (c). High-resolution transmission electron microscopy images of Zn3In2S6/NiFe2O4 (d).
Figure 3. Scanning electron microscope images of Zn3In2S6 (a), NiFe2O4 (b), and Zn3In2S6/NiFe2O4 (c). High-resolution transmission electron microscopy images of Zn3In2S6/NiFe2O4 (d).
Buildings 15 00958 g003
Figure 4. X-ray diffraction patterns (a) and X-ray photoelectron spectra (b) of Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4. The elemental orbital spectra of Fe (c) and Ni (d) in the composite material of NiFe2O4 vs. Zn3In2S6/NiFe2O4, and Zn3In2S6 vs. the elemental orbital fineness spectra of Zn (e), In (f), Zn 2p3/2 and Zn 3p1/2 orbital peaks (g), and S 2p3/2 and S 2p1/2 orbital peaks (h).
Figure 4. X-ray diffraction patterns (a) and X-ray photoelectron spectra (b) of Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4. The elemental orbital spectra of Fe (c) and Ni (d) in the composite material of NiFe2O4 vs. Zn3In2S6/NiFe2O4, and Zn3In2S6 vs. the elemental orbital fineness spectra of Zn (e), In (f), Zn 2p3/2 and Zn 3p1/2 orbital peaks (g), and S 2p3/2 and S 2p1/2 orbital peaks (h).
Buildings 15 00958 g004
Figure 5. UV−Vis diffuse reflectance patterns of Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4 photoanodes (a) and the corresponding calculated energy bandwidth information (b). Photoluminescence (PL) intensity (c) and Mott–Schottky curves (d) for Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4 photoanodes.
Figure 5. UV−Vis diffuse reflectance patterns of Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4 photoanodes (a) and the corresponding calculated energy bandwidth information (b). Photoluminescence (PL) intensity (c) and Mott–Schottky curves (d) for Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4 photoanodes.
Buildings 15 00958 g005
Figure 6. Open−circuit potential (a) and photogenerated current intensity (b) of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites coupled with 304 SS in a 3.5 wt.% NaCl solution. Open-circuit potential (c) and photogenerated current intensity (d) of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites coupled with Q235 CS.
Figure 6. Open−circuit potential (a) and photogenerated current intensity (b) of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites coupled with 304 SS in a 3.5 wt.% NaCl solution. Open-circuit potential (c) and photogenerated current intensity (d) of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites coupled with Q235 CS.
Buildings 15 00958 g006
Figure 7. Dynamic polarization curves of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites coupled with 304 SS in a 3.5 wt.% NaCl solution under closed light (a) and open light (b). Dynamic polarization curves of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites coupled with Q235 CS under closed light (c) and open light (d).
Figure 7. Dynamic polarization curves of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites coupled with 304 SS in a 3.5 wt.% NaCl solution under closed light (a) and open light (b). Dynamic polarization curves of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites coupled with Q235 CS under closed light (c) and open light (d).
Buildings 15 00958 g007
Figure 8. Nyquist curves of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites in a 3.5 wt.% NaCl solution under dark (a) and light (b) conditions.
Figure 8. Nyquist curves of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4 composites in a 3.5 wt.% NaCl solution under dark (a) and light (b) conditions.
Buildings 15 00958 g008
Figure 9. Spectra of superoxide radical (a) and Hydroxyl radical (b) with DMPO as trapping agent.
Figure 9. Spectra of superoxide radical (a) and Hydroxyl radical (b) with DMPO as trapping agent.
Buildings 15 00958 g009
Figure 10. Charge transfer mechanism of the Zn3In2S6/NiFe2O4 Z−scheme heterojunction.
Figure 10. Charge transfer mechanism of the Zn3In2S6/NiFe2O4 Z−scheme heterojunction.
Buildings 15 00958 g010
Table 1. Energy band structure information for Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4.
Table 1. Energy band structure information for Zn3In2S6, NiFe2O4, and Zn3In2S6/NiFe2O4.
PhotoanodeFlat Charged Level
(V vs. Ag/AgCl)
Conductive Tape
(eV)
Bandgap Width
(eV)
Zn3In2S6−0.80−0.802.72
NiFe2O4−0.28−0.282.41
Zn3In2S6/NiFe2O4−0.63−0.632.32
Table 2. Dynamic polarization curve parameters of 304 SS after coupling with NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4.
Table 2. Dynamic polarization curve parameters of 304 SS after coupling with NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4.
SampleEcorr or Ecouple
(V)
Icorr or Icouple
(μA·cm−2)
βa
(V decay−1)
βc
(V decay−1)
304 SS−0.160.923.29−8.52
NiFe2O4 dark−0.071.1915.54−5.21
NiFe2O4 illumination−0.903.9823.16−5.49
Zn3In2S6 dark−0.120.631.93−7.28
Zn3In2S6 illumination−0.625.308.79−5.49
Zn3In2S6/NiFe2O4 dark−0.211.341.93−7.28
Zn3In2S6/NiFe2O4 illumination−0.78102.03−6.18
Table 3. Dynamic polarization curve parameters of Q235 CS after coupling with NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4.
Table 3. Dynamic polarization curve parameters of Q235 CS after coupling with NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4.
SampleEcorr or Ecouple
(V)
Icorr or Icouple
(μA·cm−2)
βa
(V decay−1)
βc
(V decay−1)
Q235 CS−0.681.0413.71−13.68
NiFe2O4 dark−0.685.018.24−9.25
NiFe2O4 illumination−0.678.337.35−5.17
Zn3In2S6 dark−0.713.7120.14−4.92
Zn3In2S6 illumination−0.9616.8111.26−5.38
Zn3In2S6/NiFe2O4 dark−0.725.2417.25−6.84
Zn3In2S6/NiFe2O4
illumination
−1.1432.486.02−8.55
Table 4. Nyquist curve results for equivalent circuit fittings of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4.
Table 4. Nyquist curve results for equivalent circuit fittings of NiFe2O4, Zn3In2S6, and Zn3In2S6/NiFe2O4.
SampleRs
(kΩ·cm2)
Rct
(kΩ·cm2)
CPEcRf
(kΩ·cm2)
CPEdl
Yo(SnΩ−1cm−2)n1Ydl(SnΩ−1cm−2)n2
NiFe2O4 dark13.872.3 × 1055.5 × 10−615.11.2 × 10−50.90
Zn3In2S6 dark14.493.7 × 1061.1 × 10−6148.77.8 × 10−60.94
Zn3In2S6/NiFe2O4
dark
13.221.1 × 1077.1 × 10−60.951.7 × 1023.1 × 10−60.95
NiFe2O4 light15.561.0 × 1035.52 × 10−511.5 × 1048.8 × 10−40.92
Zn3In2S6 light17.031.9 × 1023.4 × 10−40.767.8 × 1022.9 × 10−30.96
Zn3In2S6/NiFe2O4 light15.151.7 × 1021.7 × 10−50.861.8 × 1041.4 × 10−41
13.872.3 × 1055.5 × 10−615.11.2 × 10−50.90
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

Wang, X.; Chen, Y.; Zhang, X. Design of a One-Dimensional Zn3In2S6/NiFe2O4 Composite Material and Its Photocathodic Protection Mechanism Against Corrosion. Buildings 2025, 15, 958. https://doi.org/10.3390/buildings15060958

AMA Style

Wang X, Chen Y, Zhang X. Design of a One-Dimensional Zn3In2S6/NiFe2O4 Composite Material and Its Photocathodic Protection Mechanism Against Corrosion. Buildings. 2025; 15(6):958. https://doi.org/10.3390/buildings15060958

Chicago/Turabian Style

Wang, Xiaotong, Yuehua Chen, and Xiaoying Zhang. 2025. "Design of a One-Dimensional Zn3In2S6/NiFe2O4 Composite Material and Its Photocathodic Protection Mechanism Against Corrosion" Buildings 15, no. 6: 958. https://doi.org/10.3390/buildings15060958

APA Style

Wang, X., Chen, Y., & Zhang, X. (2025). Design of a One-Dimensional Zn3In2S6/NiFe2O4 Composite Material and Its Photocathodic Protection Mechanism Against Corrosion. Buildings, 15(6), 958. https://doi.org/10.3390/buildings15060958

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