Skip to main content
Global Challenges logoLink to Global Challenges
. 2025 May 6;9(6):2500011. doi: 10.1002/gch2.202500011

Synthesis, Characterization, Photo/Electrocatalytic Activity, and Corrosion Behavior of Ag2S/PANI Composite Layered TiO2 Nanotubes Coated on Ti Foil

Derya Birhan 1, Derya Tekin 1,, Burak Dikici 2, Taner Tekin 3, Hakan Kızıltaş 3
PMCID: PMC12151799  PMID: 40510650

Abstract

In this study, the photo/electrocatalytic and electrochemical properties of the composite structure formed by coating silver sulfide (Ag2S) and polyaniline (PANI) on titanium nanotubes are investigated. Titanium nanotubes (TiO2 NTs) are synthesized on the surface of titanium (Ti) sheets using the anodization method. Ag2S is deposited on the surface of TiO2 NTs through the successive ionic layer adsorption and reaction (SILAR) method, while PANI is coated via the chemical oxidative polymerization technique. SEM‐EDS, XRD, FTIR, and electrochemical measurements are used to characterize the prepared nanotubes. The corrosion inhibitor performance of TiO2 NTs is investigated by electrochemical impedance spectroscopy (EIS) and potentiodynamic polarization (PDS) measurements. EIS measurements showed that the PANI layer decreased the resistance of the photoanodes, while the Ag2S coating increased the resistance compared to PANI. It is determined that the semicircles in the Nyquist curves drawn using EIS measurements are formed due to the Warburg resistance in the low‐frequency region and the high‐frequency region caused by the charge transfer resistance. Corrosion parameters are measured using Tafel fitting curves. Overall, TiO2 NTs grown by anodization effectively prevented the corrosion of the Ti photoelectrode, and the corrosion rate is calculated at 0.013 mm/year (mpy). In electrocatalytic (EC), photocatalytic (PC), and photoelectrocatalytic (PEC) activity experiments, PANI/Ag2S/TiO2 NT photoelectrode removed 13.775%, 37.71%, and 61.075% of methylene blue (MB) within 120 min under UV light irradiation.

Keywords: Ag2S NPs, anodization, conductive polymers, corrosion, photo/electrocatalytic, TiO2 nanotube


This study explores the corrosion resistance and photoelectrocatalytic performance of TiO2 nanotubes (TiO2 NTs) coated with silver sulfide (Ag2S) and polyaniline (PANI). Electrochemical impedance spectroscopy (EIS) and potentiodynamic polarization techniques are used to evaluate the corrosion behavior of the samples in a simulated corrosive environment. The synergistic effect of Ag2S and PANI coatings significantly enhanced the corrosion resistance compared to bare TiO2 NTs. Additionally, the photoelectrocatalytic activity is analyzed under simulated solar illumination, demonstrating improved charge separation and photocurrent response. The results indicate that Ag2S/PANI‐modified TiO2 NTs exhibit superior anti‐corrosion and photoelectrocatalytic properties, making them promising candidates for sustainable energy and protective coating applications.

graphic file with name GCH2-9-2500011-g006.jpg

1. Introduction

With the emergence of green electricity production in recent years, electrochemical technologies are gaining importance. Hybrid processes developed by combining electrochemical steps with biological, chemical, or physical methods are now widely used. As an example of these hybrid processes, we can give the photoelectrocatalytic (PEC) process, which occurs by combining electrochemical and photochemical events. The emergence of the PEC process is important for photocatalytic (PC) activity and increases PC activity.[ 1 ] The use of photocatalytic technology to solve environmental pollution and energy crises is developing thanks to its high efficiency features, and clean and economical features. By this time, visible and UV light‐sensitive materials have been widely studied and excellent materials have been developed.[ 2 ] Titanium dioxide (TiO2) is a typical semiconductor used for reasons such as low cost, high stability, wide band gap, nontoxic, and high quantum efficiency of light‐induced charge.[ 3 ] However, the charge separation rate and low quantum efficiency of TiO2 greatly limit its photocatalytic activity.[ 4 ] Recently, various forms of TiO2, including hollow spheres,[ 5 ] nanowires,[ 6 ] nanoparticles,[ 7 ] nanosheets,[ 8 ] and nanotubes,[ 9 ] have been developed to solve the above‐mentioned problems. In particular, ordered structures of TiO2, such as nanotubes, exhibit a higher surface area/volume ratio compared to other forms. Among these forms, TiO2 nanotubes 10 ] are unique as they have excellent properties such as large specific surface area, short electron transfer paths, highly vertically oriented ordered structures, and charge transfer channels.[ 11 , 12 , 13 ] Although the morphology of TiO2 NTs has expanded the light absorption regions and increased the number of reaction sites of the photocatalyst, this structure alone does not satisfactorily improve the photocatalytic performance of TiO2. A high surface area/volume ratio reduces the recombination rate by increasing charge separation efficiency and electron transport rate. Additionally, these materials make UV light or sunlight ideal for photoelectronic applications. These features offer a wide range of applications in many technological fields such as solar cells, battery electrodes, and sensors. The most important of these areas are photocatalytic (PC) and photoelectrocatalytic (PEC) application areas.[ 1 , 14 ] TiO2 used in PEC activity is the most preferred photocatalyst in PEC separations due to reasons such as being environmentally friendly, photoactive, resistant to chemical corrosion and photo corrosion, and economical. However, TiO2 wide band gap of 3.2 eV absorbs only 4% of solar energy and releases 48%. In recent years, to increase photocatalytic activity, sulfide‐based materials with narrow band gaps (MoS2, CuS, NiS, CdS, SnS2, Bi2S3, and Ag2S)[ 15 ] have been doped. Among these materials, silver sulfide (Ag2S) is a direct band gap (0.9–1.05 eV)[ 16 ] highly stable semiconductor, has superior catalytic potential, is nontoxic, has unique electrical and optical properties, is economical, and has antibacterial properties.[ 17 , 18 ]

Polyaniline (PANI) is the most extensively researched type of conductive polymer due to its environmental stability, easy synthesis, unique reversible protonic capability, variable electrical conductivity, and excellent redox reversibility.[ 19 , 20 ] It is possible to synthesize PANI through chemical/electrochemical, acid/base treatment, and oxidation/reduction. K. Lim et al synthesized conductive PANI/TiO2 nanotube rods using chemical oxidation polymerization.[ 21 ] Although there are studies on metal sulfides and conductive polymers, photoelectrodes with Ag2S and PANI coated on nanotubes grown on TiO2 foil are unavailable in the literature. Chen et al, who synthesized PANI, an excellent conductive polymer, through chemical oxidation, successfully synthesized the CuS/PANI composite structure with a core‐shell structure.[ 22 ] In this study, they achieved a specific capacitance of 308.1 F g−1 at 0.5 A g−1 of this composite structure. In addition, PANI photoelectrodes attract attention for reasons such as being easy to synthesize, environmentally stable, good electrical conductivity, and economical. They are used to make TiO2 nanotubes photosensitive to significantly increase the photocatalytic activity efficiency under visible light due to their delocalized conjugated structure during electron transfer.[ 23 ] The PANI‐modified metal sulfide structure was synthesized by Xiu Fang Wang et al.[ 24 ] This synthesized structure was used in the photocatalytic degradation of Rhodamine B dye and it was determined that more than 90% of the dye was removed.

In this study, TiO2 nanotube photoelectrodes were synthesized using the anodization method, and Ag2S/TiO2 nanotubes were synthesized using the SILAR method. By taking advantage of the strong interaction of Ag2S/TiO2 nanotubes, the photoelectrocatalytic activity of photoelectrodes can be improved, so that Ag2S/TiO2 NTs can be efficiently used in the removal of methylene blue. Finally, polyaniline synthesized by the chemical oxidation polymerization method was coated on the surface of Ag2S/TiO2 NTs, and the electrochemical properties of the nanotubes were examined. Characterization of PANI/Ag2S/TiO2 NT photoelectrodes was performed by SEM‐EDS, XRD, and FTIR analyses. Electrochemical measurements were made using a three‐electrode system in 1M H2SO4 solution, and the photoelectrocatalytic activities of PANI/Ag2S/TiO2 NT photoelectrodes were examined using methylene blue dye.

2. Results and Discussion

2.1. SEM‐EDS Analyzes

Figure 1 shows SEM‐EDS images of TiO2, PANI/TiO2, Ag2S/TiO2 and PANI/Ag2S/TiO2 NTs. In the SEM image of TiO2 NTs in Figure 1a, it is seen that the nanotubes synthesized by anodization are self‐organized and vertically oriented. The O/Ti ratio of TiO2 NTs was calculated as ≈1.49. The images show that the nanotube arrays are tightly aligned, and the average nanotube diameter is 100 nm.

Figure 1.

Figure 1

SEM‐EDS mapping images of a) TiO2 NT, b) PANI/TiO2 NT, c) Ag2S/TiO2 NT and d) PANI/Ag2S/TiO2 NTs.

In the SEM image of PANI/TiO2 NTs shown in Figure 1b, it is seen that the nanotubes prepared by the anodization process at 20 V for 1 h under optimized conditions have a highly ordered structure and the average diameter is 100 nm. PANI solution with a concentration of 0.01 m coated on the surface of TiO2 NTs adheres to the surface of the nanotubes like a gel, and the diameter of the nanotubes is observed to decrease.

Figure 1c shows the SEM image of Ag2S/TiO2 NT. It is seen that Ag2S nanoparticles coated using the SILAR method have a relatively uniform distribution on the surface of TiO2 NTs. After five repeated SILAR cycles, it is understood from the SEM images that Ag2S nanoparticles densely accumulate around the tubes.[ 25 ] The average diameter of Ag2S nanoparticles coated on TiO2 nanotubes by the SILAR method is between 10 and 25 nm.

Figure 1d shows the SEM image of PANI/Ag2S/TiO2 NTs. PANI synthesized by chemical oxidation polymerization method was coated on the Ag2S/TiO2 NTs surface like a gel. From the SEM image, it can be seen that Ag2S nanoparticles adhere to the walls of TiO2 NTs and PANI covers this structure. It can be seen that the diameters of the nanotubes are larger than those of Ag2S/TiO2 NTs. The reason for this is that the agglomeration of Ag2S decreases as a result of the electrostatic interaction between Ag2S nanoparticles and PANI due to the positive charge of PANI. Although some tubes are seen to be closed due to doping, TiO2 nanotubes synthesized by the anodization method have a size of 100 nm, and it is understood from the SEM images that the doped Ag2S nanoparticles and PANI coating cause the thickness of the nanotube walls to increase.

EDS mapping was performed and the results are shown in Figure 1. The presence of N, O, Ti, S, and Ag elements confirmed the formation of the PANI/Ag2S/TiO2 NT structure. EDS analysis showed that all elements in the structure were well distributed.

2.2. XRD Analysis

The crystal lattice structures of TiO2, PANI/TiO2, Ag2S/TiO2, and PANI/Ag2S/TiO2 NTs were examined by XRD in the diffraction peaks seen in Figure 2 .

Figure 2.

Figure 2

XRD diagram of TiO2, PANI/TiO2, Ag2S/TiO2, and PANI/Ag2S/TiO2 NTs.

The 2θ peaks of TNTs at 25.27°, 35.14°, 38.34°, 40.16°, 53.11°, 55.07°, 62.96°, 70.61°, 76.23°, and 82.41° correspond to the (101), (103), (112), (004), (105), (105), (211), and (201) planes. The diffraction peaks of TiO2 NTs (JCPDS No. 21‐1272) are in perfect harmony with the cards. At the 2θ peaks, diffraction peaks belonging to Ag2S nanoparticles were seen and were compatible with (JCPDS No 14‐0072) cards.[ 26 ] The small peak seen at 48.07 in the XRD diagram of Ag2S/TiO2 and PANI/Ag2S/TiO2 NT corresponds to the (200) plane and it is understood that this peak belongs to Ag2S. Ag2S nanoparticles with monoclinic structure show diffraction peaks in the same places as TiO2. No obvious diffraction peak exists in samples containing sulfur due to its small size and amount.[ 27 ] It is understood that the peaks belonging to the anatase TiO2 phase are found in a dispersed form of recrystallized Ag2S nanoparticles at the observed angles. The small peak seen at 48.07 belongs to Ag2S and is seen in the XRD diagram. The Ag2S content in the Ag2S/TiO2 and PANI/Ag2S/TiO2 composite structures is 4.33% and 4.05% by weight. The small peak seen in the range of 30°‐35° in the XRD graph of PANI/Ag2S/TiO2 NT is also observed in the Ag2S/TiO2 NT graph. The peak is a reflection of the Ag2S crystal phase. It is thought that the PANI coating causes the intensity of the XRD peak to increase since it stabilizes the crystal structure of Ag2S.

The reaction mechanism of metallic Ag formation is as follows.

AgNO3Ag++NO3 (1)
Na2S2N2++S2 (2)
2Ag++S2Ag2S (3)

Metallic Ag and SO2 are produced during the calcination process conducted at 500 °C.

2Ag2S+3O24Ag+2SO2 (4)

2.3. FTIR analysis

The FTIR spectra of TiO2, PANI/TiO2, Ag2S/TiO2 and PANI/Ag2S/TiO2 NTs are shown in Figure 3 . In the FTIR spectrum of TiO2 nanotubes, the band seen at 3343 cm−1 corresponds to O‐H stretching vibration, while the band seen at 1725 cm−1 is attributed to the presence of hydroxyl groups (─OH). The band at 683 cm−1 corresponds to Ti─O─Ti stretching vibrations. The peak seen at 2360 cm−1 is due to the Ag─S bonds in the structure of Ag2S. PANI/Ag2S/TiO2 and Ag2S/TiO2 NTs show the absorption band of Ag‐S ≈1062 cm−1.[ 28 ] The low prominence of the peak at 2360 cm−1 is because the denser the Ag2S coating, the lower the density of functional groups containing the peak. The chemical bond structure of TiO2, PANI/TiO2, Ag2S/TiO2, and PANI/Ag2S/TiO2 NTs was confirmed by FTIR analysis. In the FTIR spectrum given in Figure 3, the peak of PANI/TiO2 and PANI/Ag2S/TiO2 NTs at 3236 cm−1 corresponds to the symmetric N─H stretching band. The absorption peaks at 1401 and 1575 cm−1 correspond to the C=C bond formed by aromatic ring stretching vibrations of quinoid and benzenoid rings. While the low‐frequency bands at 1297 cm−1 correspond to C‐N vibration, the peak at 1136 cm−1 corresponds to C─H in‐plane bending, and the peaks at 825 and 504 cm−1 correspond to C─H out‐of‐plane bending.[ 29 ]

Figure 3.

Figure 3

FTIR spectrum of TiO2, PANI/TiO2, Ag2S/TiO2 and PANI/Ag2S/TiO2 NTs.

2.4. Photocatalytic (PC), Electrocatalytic (EC) and Photoelectrocatalytic (PEC) Activity Properties of TiO2 NT Photoelectrodes

Photocatalytic, electrocatalytic, and photoelectrocatalytic experiments of TiO2, Ag2S/TiO2, and PANI/Ag2S/TiO2 NTs lasted 120 min and the results are shown in Figure 4 .

Figure 4.

Figure 4

Electrocatalytic, photocatalytic and photoelectrocatalytic degradation graphs of TiO2, Ag2S/TiO2 and PANI/Ag2S/TiO2 NT photoelectrodes on methylene blue dye.

The dye removal efficiency of TiO2 NTs was evaluated by examining their degradation in methylene blue aqueous solution. PEC activity of TiO2 photoelectrodes prepared at 20V anodization voltage and 500 °C annealing temperature was evaluated in MB solution prepared with an initial dye concentration of 20 ppm containing 0.05m Na2SO4. Studies in the literature have determined that nanotube arrays created by anodization at 20V lead to high PEC activity.[ 30 , 31 ] TiO2 nanotubes removed the MB solution by photocatalytically 15.74%, electrocatalytically 5.635%, and photoelectrocatalytically 28.02% of the MB solution within 120 min. The photocatalytic performance of Ag2S/TiO2 NTs on MB was higher than that of pure TiO2, and it removed 27.215% of the dye within 120 min. Ag2S/TiO2 NT shows better performance than pure TiO2. The main reason is that Ag2S nanoparticles expand the response range of TiO2 to light and reduce the recombination rate of photoinduced electron‐hole pairs. PC, EC, and PEC removal graphs of the synthesized samples are shown in Figure 4. In EC and PEC processes using Ag2S/TiO2 NT photoelectrode under UV irradiation of MB, 10.075% and 45.09% removal were achieved, respectively. Accordingly, since the EC removal of MB without UV light irradiation is less than the PEC and PC removals, it is understood that the degradation of MB into EC cannot be done quickly.[ 32 ] As a result of the bias potential of 0.6 V applied to Ag2S/TiO2 NTs, the PEC decay rate of MB is ≈1.6 times that of the PC degrade rate. It was determined that the PC and PEC decay rates are 1.7 and 1.6 times higher than pure TiO2 NTs, respectively. Ag2S/TiO2 NTs have the second‐best removal performance in terms of PEC. This is because the applied bias potential prevents the recombination of photogenerated electron‐hole pairs, thus extending the lifetime of photogenerated carriers. The EC, PC and PEC degradations of PANI/Ag2S/TiO2 NT photoelectrode on methylene blue are shown in Figure 4. As shown in Figure 4, PANI/Ag2S/TiO2 NT removed 37.71% of methylene blue as PC. In the EC experiment where 0.6V voltage was applied, it removed 13.775% of the dye and in the PEC experiment, it removed 61.075% of the dye. The PEC degradation activity of MB is higher than the other samples, 2.18 times higher than TiO2 NT and 1.35 times higher than Ag2S/TiO2 NT.

PANI/Ag2S/TiO2 NTs showed different degradation efficiencies of 13.78%, 37.71%, and 61.075% in EC, PC, and PEC activities for 2 h, respectively. When compared with pure TiO2 and Ag2S/TiO2 NTs, PC and PEC activities are pretty good except for EC degradation rates. It is understood from the degradation experiments that the modification of PANI can prevent the recombination of electron‐hole pairs of TiO2 NTs under UV light and thus increase the degradation performance of organic dyes.[ 33 ]

2.5. Electrochemical Impedance Spectroscopy (EIS) of TiO2 NT Photoelectrodes

Nyquist plots are shown in Figure 5 after taking open circuit potential (OCP) for 2 h in 1 m H2SO4 solution. According to electrochemical theory, a current diagram is obtained by charging and discharging the double layer at the interface and oxidation/reduction of chemicals on the electrode surface or in solution. The equivalent electrical simple model circuit for EIS measurements is also shown in Figure 5. The appropriate equivalent circuit for Nquist plots was determined as R1 + Q1/R2 + Q2/(R3 + W3). Nyquist plots and the corresponding equivalent circuit connection of PANI and Ag2S‐doped TiO2 nanotube photoelectrodes are shown in Figure 5a–d. In the impedance and corrosion tests performed on TiO2, PANI/TiO2, Ag2S/TiO2, and PANI/Ag2S/TiO2 NT samples, 3 repetitions were made to ensure that the results were fully compatible with each other. Figure 6 shows the equivalent circuit model suitable for impedance data suitable for all photoelectrodes. Table 1 shows the EIS results performed in 1m H2SO4 solution.

Figure 5.

Figure 5

Nyquist plots of a) TiO2, b) PANI/TiO2, c) Ag2S/TiO2 and d) PANI/Ag2S/TiO2 NTs in 1M H2SO4 solution.

Figure 6.

Figure 6

An equivalent electrical circuit model used to fit all experimental impedance data.

Table 1.

EIS analysis results obtained in 1 m H2SO4 solution.

Samples R1 [Ω] Q1 R2[Ω] Q2 R3[Ω)] W3
TiO2 NT 0.1461 24.27e‐6 11.09 36.25e‐6 2.437e6 2796
PANI/TiO2 NT 0.4048 0.8039e‐3 0.125 72.73e‐6 467.8 104.2
Ag2S/TiO2 NT 0.1903 0.5302e‐3 102.6 0.248e‐3 3916 181.7
PANI/Ag2S/TiO2 NT 1.254 0.2324e‐3 6.259 0.3824e‐3 1323 52.86

In EIS analysis, if we need to explain each component in the R1 + Q1/R2 + Q2/(R3 + W3) equivalent circuit model used, R1: Represents the resistance due to ionic conductivity at the electrode‐electrolyte interface or in the solution. Q1 (Constant Phase Element‐CPE1): Used to model non‐ideal double‐layer capacitance. R2 (Interfacial Resistance or Polarization Resistance): Represents the charge transfer resistance between the electrode and the electrolyte. It is related to corrosion or electrochemical reaction processes. Q2 (Constant Phase Element‐CPE2) is used to represent the deviation from the second capacitance, while R3 (Diffusion Resistance or Film Resistance) represents the resistance of the passive film, thin layer, or reactive interface formed on the electrode surface. W3 (Warburg Impedance‐Diffusion Impedance): Describes diffusion processes due to mass transport. Table 1 shows that the series solution resistance (R1) increases with PANI doping but decreases with Ag2S doping. It is observed that the PANI doping of 0.01 m increases the solution resistance, which is responsible for enhancing the electrochemical supercapacitive performance of photoelectrodes. The impedance graph should be vertical and parallel to the imaginary impedance axis for an ideal photoelectrode. The charge transfer between the electrode and the electrolyte is represented by the semicircular arcs indicated by the electrode material in the Nyquist plots shown in Figure 5a–d. The semicircles here show that the system has an electrochemical interface. Warburg impedance (W3) refers to a process controlled by diffusion in the lower frequency range, that is, in a linear part. A smaller charge transfer resistance and a higher charge separation efficiency are represented by a smaller semicircle. The Nyquist curves shown in Figure 5 consist of a linear section in the low‐frequency region resulting from the Warburg resistance and a semicircular section in the high‐frequency region related to the charge transfer resistance. In comparative EIS studies of the samples, the semicircles of TiO2 NTs[ 34 ] are larger than those of other samples.[ 35 ] This shows that the resistance of the TiO2 sample is higher than other photoanodes. As the Ag2S film is coated on the surface, the resistance of the photoanode decreases. The coating polymer layer on the nanotube significantly reduces the resistance of the photoanode. It is understood that the PANI/Ag2S/TiO2 NT photoanode shows more resistance than the PANI/TiO2 NT. This shows that PANI reduces the resistance of photoanodes more than Ag2S. After the weight measurements were made after the coating process, the amount of Ag2S coated on the Ag2S/TiO2 NT surface is 0.03 g, and the PANI and Ag2S coating on the PANI/Ag2S/TiO2 NT surface is 0.09 g. In this study, the R2 value (69.93 Ω) of Ag2S/TiO2 NT may have shown a high value because Ag2S is a semiconductor and partially prevents charge transfer. This indicates that electron transitions are difficult at the interface and increase recombination. PANI coating on Ag2S/TiO2 NT significantly reduced R2 and improved charge transportability. The fact that the PANI/Ag2S/TiO2 NT (6.259 Ω) sample has the lowest W3 value indicates that ions are transported very quickly. Therefore, Ag2S/TiO2 NTs coated with PANI are the samples with the best electrochemical performance.

2.6. Potentiodynamic Polarization Study (PDS) of TiO2 Nanotubes

The potentiodynamic polarization study performed in 1M H2SO4 solution for 0.01 m Ag2S and PANI‐coated samples is shown in Figure 7 . Calculated polarization parameters for TiO2 NTs are given in Table 2 . The Icorr and Ecorr values ​​of each sample were calculated based on Tafel extrapolation and ranged from 0.046 to 166.802 µA cm−2, and 33.003 to 302.355 mV, respectively. TiO2 and Ag2S/TiO2 NTs showed good corrosion resistance compared to other samples.

Figure 7.

Figure 7

Polarization curves of TiO2 NTs coated with 0.01 m Ag2S and PANI in 1M H2SO4 solution.

Table 2.

The average value of polarization parameters for TiO2 NTs coated with 0.01 m Ag2S and PANI and anodized at 20 V.

Samples Ecorr [mV] icorr [µAcm−2] Βa [mV dec−1] Βc [mV dec−1]
TiO2 NT 302.355 0.046 562.4 210.2
PANI/TiO2 NT 161.859 116.802 9586.7 27.1
Ag2S/TiO2 NT 33.003 7.546 227.5 221.6
PANI/Ag2S/TiO2 NT 294.229 48.632 585.4 199.0

It was observed that due to the formation of a stable oxide layer at a potential value of >0.5 V, no further corrosion was observed in the samples and a passive zone that was highly effective against corrosion was formed. No further corrosion occurred due to this passive zone. Ag2S/TiO2 NT and PANI/Ag2S/TiO2 NT formed a passive layer that fractured and reformed in a three‐stage process. Coating the TiO2 NT surface with Ag2S caused an increase in corrosion potential and a decrease in corrosion current density. Ag2S reduced the corrosion rate by forming a protective barrier on the surface. However, the Ag2S coating alone cannot reach the superior protection performance provided by the double‐layer coating used with PANI. Samples coated with PANI and Ag2S showed the lowest corrosion current and the highest corrosion potential. PANI, which has conductive properties, contributes to forming an effective passive layer when combined with Ag2S, making the TiO2 surface resistant to corrosion.

Figure 8 shows the corroded images of pure, Ag2S, and PANI‐coated NTs. It is seen that especially corrosion marks on the surface of PANI/TiO2 NT are formed more and the surface is destroyed. For Ag2S/TiO2 NT, it is seen that the pores are opened and the Ag nanoparticles are corroded. As seen from the EDS images, a very dense oxide layer has formed on the structures.

Figure 8.

Figure 8

SEM‐EDS images of a) TiO2, b) PANI/TiO2, c) Ag2S/TiO2 and d) PANI/Ag2S/TiO2 NTs after corrosion tests in 1 m H2SO4.

In future studies, it is possible to change parameters such as anodization time, pH ratios, and metal doping amounts or to conduct studies in nanocomposite form. It is also used in supercapacitor studies due to PANI's electrical conductivity.

3. Conclusion

This study includes the easy production process of PANI/TiO2 NTs produced by chemical oxidation polymerization synthesis of Ag2S‐doped TiO2 NTs and aniline monomer. Characterization was performed using SEM‐EDS, XRD, and FTIR devices. Electrochemical tests were carried out in 1 m H2SO4 solution using a potentiostat. The results show that the nanotubes are composed of TiO2 NTs doped with Ag2S and PANI in the anatase phase. SEM results show that the nanotube arrays formed by anodization are quite regular and homogeneously distributed oxide layers are formed on them. XRD results showed that the produced TiO2 NTs were compatible with all peaks in the anatase phase and the contribution of Ag nanoparticles did not cause any change in the obtained peaks. FTIR results confirmed that TiO2, Ag2S, and PANI were successfully synthesized. In EC, PC, and PEC activity experiments on methylene blue, PANI/Ag2S/TiO2 NT photoelectrode showed the best degradation activity with 13.775%, 37.71%, and 61.075% degradation rates, respectively.

4. Experimental Section

Materials

Titanium, Foil, 2.0 mm Thickness, 99.7% Metals Basis, Aniline (C6H7N, ≥99.0%), Ortho‐Phosphoric acid (H3PO4, 85%), Sodium fluoride (NaF, 99.99%), Hydrochloric acid (HCl, 37%), Sodium sulfide (Na2S.9H2O, ≥98.0%), Ag/AgCl reference electrode, Sulfuric acid (H2SO4, 95‐97%) Ammonium persulfate (NH4)2S2O8, ≥98%), Silver nitrate (AgNO3, ≥99.0%) from Merck and Sigma Aldrich companies and methylene blue (C16H18ClN3S.xH2O) used in photoelectrocatalytic experiments. The dyestuff was obtained from Fluka company.

Synthesis of TiO2 Nanotube Photoelectrodes

Ti foils with a thickness of 2.0 mm were cleaned with ultrasound in acetone, ethanol, and methanol for 10 min, respectively. Then, Ti Foils were anodized in 400 mL solution containing 0.14 m NaF and 0.5 m H3PO4 at 20 V for 1 hour. After the anodization process, the Ti foils were thoroughly washed with ionized water and dried. Finally, it was calcined at 500 °C for 3 hours. The anodization process will be carried out using a two‐electrode system (Ti foil as anode and Pt wire as cathode, with a distance of 3 cm between anode and cathode).

Synthesis of Ag2S/TiO2 Nanotube Photoelectrodes

Ag2S/TiO2 NT was synthesized using the SILAR method. TiO2 NTs synthesized by anodization were dipped into 4 beakers respectively and the coating process was carried out. The coating process is as follows. The 1st beaker contains 0.01 m AgNO3, the 2nd beaker contains ionized water, the 3rd beaker contains 0.01 m Na2S and the 4th beaker contains ionized water.[ 27 ] TiO2 NTs were immersed in an AgNO3 solution and cleaned in deionized water, then immersed in a Na2S solution and cleaned in deionized water. The coating process was repeated five times. To obtain the anatase TiO2 phase, Ag2S/TiO2 NTs were dried and calcined at 500 °C in air for 3 h.

Synthesis of PANI/Ag2S/TiO2 Nanotube Photoelectrodes

PANI polymer was synthesized by the chemical oxidation method. 0.01 m aniline was mixed in ionized water in a beaker. On the other hand, 0.01 m APS was dissolved in ionized water in another beaker. After 15 min of mixing, the APS solution was added dropwise into the aniline solution. Then, the pH was adjusted to 2.3 with the help of HCl acid. Finally, the PANI polymer was coated on Ag2S/TiO2 NTs with the help of sol‐gel. During the coating process, the Ti foil was kept in the solution for 1 min and the coating process was repeated five times. Finally, it was air dried.

Characterization of TiO2 NT Photoelectrodes

Morphological characteristics and elemental analysis of photoelectrodes were investigated using scanning electron microscopy (SEM‐EDS, JEOL, JSM‐7001F). Fourier transform infrared spectroscopy (FTIR, Bruker, TENSOR 27) was used to demonstrate the completion of the reaction in synthesizing TiO2 nanotubes modified with metal sulfide and polymer. The crystalline phases of the samples were determined by an X‐ray diffraction device (Rigaku, SmartLab) containing Cu‐Kβ radiation (λ = 1.3923).

Electrochemical Measurements

Electrochemical impedance spectroscopy and potentiodynamic corrosion measurements were performed with a three‐electrode system connected to a Gamry potentiostat/galvanostat (model G‐300). Ti foil (1x1 cm) was used as the anode, Pt wire as the cathode, and Ag/AgCl as the reference. For measurements, three electrodes were immersed in 1 m H2SO4 solution.

Photoelectrocatalytic Degradation of Methylene Blue

The photocatalytic, electrocatalytic, and photoelectrocatalytic activities of TiO2 NTs were tested in the degradation of methylene blue dyestuff in an aqueous solution. As experimental parameters, 400 mL solution volume, 20 ppm dye concentration, 25 °C ambient temperature, 254 nm (44W/m2) UV light source, and 0.6 V operating voltage were selected. The solution was stirred in the dark for 30 min to ensure adsorption‐desorption balance before the photoreaction started. O2 support was provided to the solution medium by an air pump and 0.05 m Na2SO4 was added. Degradation experiments were carried out using two electrode systems. The concentration of samples taken at certain time intervals was determined at 668 nm on a UV spectrophotometer (Optizen α).

Conflict of Interest

The authors declare no conflict of interest.

Acknowledgements

This project (FDK‐2022‐11336) was supported by Atatürk University, Administration System of Scientific Research Project.

Birhan D., Tekin D., Dikici B., Tekin T., Kızıltaş H., Synthesis, Characterization, Photo/Electrocatalytic Activity, and Corrosion Behavior of Ag2S/PANI Composite Layered TiO2 Nanotubes Coated on Ti Foil. Global Challenges 2025, 9, 2500011. 10.1002/gch2.202500011

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  • 1. Palmas S., Mais L., Mascia M., Vacca A., Curr Opin Electrochem 2021, 28, 100699. [Google Scholar]
  • 2. Liu T., Liu B., Yang L., Ma X., Li H., Yin S., Sato T., Sekino T., Wang Y., Appl Catal. B 2017, 204, 593. [Google Scholar]
  • 3. Guo Q., Zhou C., Ma Z., Yang X., Adv. Mater. 2019, 31, 1901997. [DOI] [PubMed] [Google Scholar]
  • 4. Gao B., Sun M., Ding W., Ding Z., Liu W., Appl. Catal. B 2021, 281, 119492. [Google Scholar]
  • 5. Wu X., Lu GQ(M)., Wang L., Energy Environ Sci 2011, 4, 3565. [Google Scholar]
  • 6. Rahman Md A, Bazargan S., Srivastava S., Wang X., Abd‐Ellah M., Thomas J P., Heinig N F., Pradhan D., Leung K. T., Energy. Environ. Sci. 2015, 8, 3363 [Google Scholar]
  • 7. Shen Q., Cao C., Huang R., Zhu L., Zhou X., Zhang Q., Gu L., Song W., Angew. Chem. Int. Ed. 2020, 59, 1216. [DOI] [PubMed] [Google Scholar]
  • 8. Han Q., Wu C., Jiao H., Xu R., Wang Y., Xie J., Guo Q., Tang J., Adv. Mater. 2021, 33, 2008180. [DOI] [PubMed] [Google Scholar]
  • 9. Li Z., Zhang Z., Dong Z., Wu Y., Zhu X., Cheng Z., Liu Y., Wang Y., Zheng Z., Cao X., Wang Y., Liu Y., J. Solid. State. Chem. 2021, 303, 122499. [Google Scholar]
  • 10. Acar M. T., Çomaklı O., Yazıcı M., Arslan M. E., Yetim A. F., Çelik A., Surf. Interfac. 2024, 50, 104472. [Google Scholar]
  • 11. Zhang S., Long M., Zhang P., Wang J., Lu H., Xie H., Tang A., Yang H., Chem. Eng. J. 2022, 429, 132091. [Google Scholar]
  • 12. Makarova M V., Amano F., Nomura S., Tateishi C., Fukuma T., Takahashi Y., Korchev Y E., ACS. Catal. 2022, 12, 1201. [Google Scholar]
  • 13. Wang K., Zhao K., Qin X., Chen S., Yu H., Quan X., J. Hazard. Mater. 2022, 424, 127747. [DOI] [PubMed] [Google Scholar]
  • 14. Palmas S., Da Pozzo A., Delogu F., Mascia M., Vacca A., Guisbiers G., J. Power. Sources 2012, 204, 265. [Google Scholar]
  • 15. Zhang Y., Zhu M., Zhang S., Cai Y., Lv Z., Fang M., Tan X., Wang X., Appl. Catal. B 2020, 279, 119390. [Google Scholar]
  • 16. Borovaya M., Naumenko A., Horiunova I., Plokhovska S., Blume Y., Yemets A., Appl. Nanosci. 2020, 10, 4931. [Google Scholar]
  • 17. Zhang L., Li P., Feng L., Chen Xi, Jiang J., Zhang S., Zhang C., Zhang A., Chen G., Wang H., J. Hazard. Mater. 2020, 387, 121715. [DOI] [PubMed] [Google Scholar]
  • 18. Offor P. O., Whyte G. M., Ezekoye V. A., Omah A. D., Ude S. N., Ocheri C., Ezukwoke N., Ezema I. C., Madiba I. G., Okorie B. A., Maaza M., Ezema F I., Optik 2019, 185, 519. [Google Scholar]
  • 19. Naik Y. V., Kariduraganavar M., Srinivasa H. T., Siddagangaiah P. B., Naik R., J. Energy. Storage 2024, 97, 112874. [Google Scholar]
  • 20. Altannyhi K. A., Elnaggar E. M., Elsayed B. A., Elsenety M. M., Egypt J. Pet. 2024, 33, 14. [Google Scholar]
  • 21. Karim M. R., Yeum J. H., Lee Mu S, Lim K. T., React. Funct. Polym. 2008, 68, 1371. [Google Scholar]
  • 22. Chen C., Zhang Q., Peng C., Polym. Polym. Compos. 2017, 25, 483. [Google Scholar]
  • 23. Palmas S., Mascia M., Vacca A., Llanos J., Mena E., RSC Adv. 2014, 4, 23957. [Google Scholar]
  • 24. Wang X., Chen S., Shuai Y. J. R. J. o. P. C. A., Phys. Chem. A 2016, 90, 2069. [Google Scholar]
  • 25. Dong W., Gao H., Lin S., Dong Z., Zheng Z., Wu Y., Zhu X., Cheng Z., Liu Y., Wang Y., Cao X., Wang Y., Zhang Z., Liu Y., J. Solid State Chem. 2022, 310, 123010. [Google Scholar]
  • 26. Wu X., Liao L., Du W., Qin A., Proc. Eng. 2015, 102, 273. [Google Scholar]
  • 27. Kiziltaş H., Tekin T., Chem. Eng. Commun. 2017, 204, 852. [Google Scholar]
  • 28. Zahedifar M., Seyedi N., Razavi R., Biomass. Convers. Biorefin. 2022, 14, 10011. [Google Scholar]
  • 29. Nada F., Atta, Galal A., Amin H. M. A., Int. J. Electrochem. Sci. 2012, 7, 3610. [Google Scholar]
  • 30. Zhang Z., Yuan Y., Shi G., Fang Y., Liang L., Ding H., Jin L., Environ. Sci. Technol. 2007, 41, 6259. [DOI] [PubMed] [Google Scholar]
  • 31. Kiziltas H., Opt. Mater. 2022, 123, 111926. [Google Scholar]
  • 32. Xie K., Sun L., Wang C., Lai Y., Wang M., Chen H., Lin C., Electrochim. Acta. 2010, 55, 7211. [Google Scholar]
  • 33. Zhao D., Jiang L., Yang R., Zhang Y., Zhou Q., Chemosphere 2022, 302, 134928. [DOI] [PubMed] [Google Scholar]
  • 34. Sivaprakash V., Natrayan L., Suryanarayanan R., Narayanan R., Paramasivam P., J. Nanomater. 2021, 2021, 1. [Google Scholar]
  • 35. Poudel M. B., Yu C., Kim H. J., Catalysts 2020, 10, 546. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.


Articles from Global Challenges are provided here courtesy of Wiley

RESOURCES