Next Article in Journal
Correction: Şeker et al. Comparative Analysis of Phenolic, Carotenoid, and Elemental Profiles in Three Crataegus Species from Şebinkarahisar, Türkiye: Implications for Nutritional Value and Safety. Molecules 2025, 30, 2934
Previous Article in Journal
Recent Advances in the Synthesis and Applications of Nitrogen-Containing Macrocyclic Arenes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Construction of Hollow TiO2/ZnS Heterojunction Photocatalysts for Highly Enhanced Photodegradation of Tetracycline Hydrochloride

Anhui Provincial Key Laboratory of Green Carbon Chemistry, School of Chemistry and Material Engineering, Fuyang Normal University, Fuyang 236037, China
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(17), 3644; https://doi.org/10.3390/molecules30173644
Submission received: 3 August 2025 / Revised: 1 September 2025 / Accepted: 5 September 2025 / Published: 7 September 2025

Abstract

TiO2 photocatalysts exhibit great potential in solar fuel production and environmental remediation, yet their practical applications are often hindered by high electron-hole recombination rates. This study presents a novel strategy for fabricating hollow anatase TiO2-modified ZnS heterostructures (TiO2/ZnS) via a simple hydrothermal method. The heterostructure effectively combines the high electron mobility of ZnS, which facilitates rapid photogenerated electron transfer, with the high specific surface area of hollow TiO2, which enhances pollutant adsorption. As a result, TiO2/ZnS demonstrates superior tetracycline degradation efficiency due to optimized charge separation and improved accessibility to reactive sites, compared to pristine TiO2 and ZnS. Furthermore, the enhanced photocatalytic activity is attributed to efficient charge separation facilitated by Type-II heterojunctions between ZnS and anatase TiO2. Cycling tests reveal that TiO2/ZnS retains over 94% of its activity after 5 cycles. This work offers a versatile approach for stabilizing metal oxides through heterostructure engineering, with significant implications for scalable environmental catalysis.

1. Introduction

Tetracycline hydrochloride is one of the most fundamental and widely utilized antibiotics among the tetracycline class of broad-spectrum antibiotics [1]. However, only a fraction of tetracycline is absorbed by humans or animals, while the remaining active pharmaceutical ingredients persist in aquatic and terrestrial ecosystems, leading to substantial ecological harm through the induction of antibiotic resistance and disruption of microbial communities [2]. Traditionally, titanium dioxide (TiO2) has been extensively used as a photocatalytic material for tetracycline degradation [3]. Nevertheless, the TiO2-based systems that rely on physical adsorption suffer from rapid shifts in adsorption–desorption equilibrium and high regeneration costs, limiting their operational sustainability [4,5]. As a result, advanced photocatalytic technologies have emerged as promising alternatives, offering the dual benefits of solar energy utilization and the complete mineralization of organic pollutants, thus enhancing both green economy metrics and reaction step efficiency [6].
Most reported photocatalytic systems rely predominantly on ultraviolet light irradiation, which constitutes only about 5% of the solar spectrum, leading to significant underutilization of solar energy [7]. In contrast, visible light-driven TiO2-based photocatalysts have garnered significant attention as sustainable alternatives, operating under visible light conditions with reduced energy input [8]. However, the intrinsic limitations of TiO2—specifically its wide bandgap (anatase phase: 3.2 eV) and suboptimal quantum efficiency (<5%)—severely restrict its photocatalytic activity under visible-light irradiation, resulting in performance that is far from satisfactory [9,10].
To overcome these limitations, considerable efforts have been devoted to developing visible light-responsive TiO2-based photocatalysts. Among various strategies, constructing heterojunctions with semiconductors has proven particularly effective [11]. Indeed, highly active co-catalysts, such as metal compounds (e.g., Bi2O3, ZnO, and ZnS), can form heterojunctions with TiO2, thereby generating built-in electric fields that reduce carrier recombination rates and prolong hole lifetimes, thus enhancing the overall photocatalytic efficiency [12,13,14,15,16,17,18]. Therefore, to achieve high efficiency in tetracycline degradation, it is crucial to systematically investigate the electronic band structures of co-catalysts and the interfacial degradation mechanisms between co-catalysts and photocatalysts [19]. However, this remains a significant challenge. Notably, ZnS has emerged as an ideal co-catalyst candidate due to its appropriate band gap (3.7 eV), high conduction band potential, excellent electron mobility, good chemical stability, stable anchoring on support surfaces, and ability to optimize interfacial charge transfer [20,21]. When coupled with ZnS, TiO2 can form efficient heterojunctions that not only extend light absorption [22] but also significantly promote the separation of photogenerated charge carriers through built-in electric fields [23]. This synergistic effect addresses the core challenge of rapid electron-hole recombination in pure TiO2 while maintaining strong redox capability.
Regretfully, TiO2/ZnS heterojunction systems still face critical challenges in photocatalytic antibiotic degradation, including rapid photogenerated hole recombination and unresolved mechanistic details regarding the tetracycline degradation pathway [24]. Moreover, achieving simultaneously enhanced light absorption and charge separation in a rationally designed architecture continues to be a significant materials challenge [25,26].
In this study, we designed and synthesized a novel TiO2/ZnS heterojunction photocatalyst with a hollow spherical structure specifically for efficient tetracycline degradation under visible light irradiation. The TiO2/ZnS heterojunction catalyst extends the light absorption range of wide-bandgap TiO2 while simultaneously suppressing charge carrier recombination in ZnS. Characterization via transmission electron microscopy (TEM) and X-ray photoelectron spectroscopy (XPS) revealed that TiO2 exhibited a hollow porous structure, and a distinct heterojunction interface was formed between ZnS and TiO2, confirming the successful fabrication of the composite. Performance tests showed that the average carrier lifetimes of photogenerated holes in TZNS were 1.85 and 1.42 times longer than those of pure TiO2 and ZnS, respectively, indicating significantly enhanced charge separation efficiency. In situ electron paramagnetic resonance (EPR) analysis revealed that ZnS optimized the driving force of photogenerated holes, promoting hydroxyl radical generation and accelerating tetracycline degradation. Under visible light irradiation (100 min), the optimal sample TZNS-100 achieved 99% tetracycline degradation efficiency, demonstrating excellent photocatalytic activity. This work provides fundamental insight into the design principles of efficient heterojunction photocatalysts for antibiotic wastewater treatment.

2. Results and Discussion

2.1. Structure Characterization

The fabrication of the TiO2/ZnS heterostructure (denoted TZnS) is systematically illustrated in Figure 1. Synthesis began with the preparation of uniform TiO2 nanospheres via controlled polystyrene sphere (CPS) template-directed self-assembly, followed by solvothermal crystallization. Subsequently, the TiO2 spheres were purified via centrifugation and vacuum drying, then reacted with zinc acetate dihydrate and thiourea under hydrothermal synthesis at 140 °C for 4 h. This sequential process enabled controlled sulfidation, yielding well-defined TiO2/ZnS core–shell heterostructures with hollow TiO2 cores and burr-like ZnS shells.
The morphologies of pristine TiO2 and ZnS were characterized by TEM (Figures S1 and S2). As shown in Figure S1a, TiO2 exhibits well-defined hollow spherical structures (Figures S1a and S2a,b), enhancing pollutant adsorption and providing abundant photocatalytic active sites. HR-TEM analysis (Figure S1e) [27] confirms the (101) lattice planes, consistent with the characteristic 25.6° (101) peak of anatase TiO2 (JCPDS: 21-1272) shown in the inset, validating the successful synthesis of hollow TiO2 spheres via template-directed self-assembly.
TEM characterization of ZnS (Figure S2a) reveals nanoparticulate aggregates with irregular surface morphology, consistent with the SEM image in Figure 2b. The lattice fringe spacing of 0.312 nm in HR-TEM (Figure S2e) corresponds to the (111) planes of cubic ZnS [28]. Furthermore, elemental mapping (Figure S2b–d) reveals uniform Zn and S distribution, confirming successful ZnS synthesis.
SEM and TEM analysis of the TZnS−100 photocatalysts (Figure 2) demonstrates intimate integration of the components. Figure 2d shows ZnS particles densely anchored on the TiO2 surface, with elemental mapping confirming uniform Zn and S distribution. HR-TEM (Figure 2f) reveals a lattice spacing of 0.347 nm [29] and 0.312 nm [28], corresponding to the (101) planes of anatase TiO2 and (111) planes of ZnS, respectively, confirming ZnS deposition on the TiO2 surface and the formation of a TiO2/ZnS N-N heterostructure.
X-ray diffraction (XRD) analysis was performed on the as-synthesized materials (Figure 3). The XRD pattern of pristine ZnS exhibits diffraction peaks at 28.89°, 47.91°, and 56.62°, corresponding to the (111), (222), and (420) planes of cubic ZnS (JCPDS: 05-0566) [30]. Pristine TiO2 displays characteristic peaks at 25.28°, 36.94°, 37.80°, 38.57°, 48.05°, 53.89°, 55.06°, 62.12° and 62.69°, corresponding to the (101), (103), (004), (112), (200), (105), (211), (213) and (204) planes of anatase TiO2 (JCPDS: 21-1272) [31], confirming the successful synthesis of pure ZnS and TiO2.
The XRD pattern of the TZnS−100 composite (Figure 3a) contains peaks from both anatase TiO2 and cubic ZnS, with reflections corresponding to the TiO2 (101) plane and ZnS (111), (222), and (420) planes (JCPDS: 21-1272 and 05-0566, respectively). Notably, the ZnS (111) peak slightly shifts to higher angles in the TZnS-X composite compared to pristine ZnS (Figure 3b), indicating lattice strain. This lattice distortion likely results from TiO2/ZnS heterostructure formation, potentially involving interfacial charge transfer, and is typical of core–shell structures with intimate interfacial contact [32].
X-ray photoelectron spectroscopy (XPS) was employed to determine the elemental composition and chemical state of the TZnS photocatalyst. The survey spectra (Figure S3a) confirm the presence of Ti, O, and C in TiO2; Zn, S, O, and C in ZnS; and Ti, O, Zn, S, and C in the TZnS composite. The C 1s peak at 284.80 eV (Figure S3b) was used for charge calibration, while other signals are attributed to adventitious carbon contamination from instrumental or sample handling.
Analysis of the Ti 2p spectrum of TZnS (Figure 4a) shows peaks at 458.64 eV (Ti 2p3/2) and 464.32 eV (Ti 2p1/2), with a spin–orbit splitting of 5.68 eV, confirming the Ti4+ oxidation state, consistent with TiO2 XPS references [33]. Compared to pure TiO2, the Ti 2p peaks of TZnS shift to lower binding energies, indicating partial Ti4+ to Ti3+ reduction upon ZnS coupling, generating oxygen vacancies and markedly enhancing photocatalytic oxidation activity [11,14].
In the O 1s spectra (Figure 4b), pure TiO2 exhibits two components: a primary peak at 529.96 eV (lattice oxygen, Ti–O) and a secondary peak at 531.81 eV (surface hydroxyl groups, –OH). In TZnS, both Ti–O and –OH peaks shift to lower binding energies compared to pure TiO2, suggesting electron transfer from ZnS to TiO2 at the heterojunction interface.
Further comparison of the Zn 2p spectrum in the TZnS composite (Figure 4c) shows a slight increase in binding energy compared to pure ZnS, with the Zn 2p3/2 peak shifting from 1021.92 eV (pure ZnS) to 1022.02 eV and the Zn 2p1/2 peak from 1044.89 eV (pure ZnS) to 1044.99 eV (0.1 eV shift). This positive shift indicates reduced electron density around Zn atoms, consistent with interfacial charge redistribution. Similarly, the S 2p spectrum (Figure 4d) shows a 0.1 eV positive shift relative to pure ZnS [34]. These shifts in Zn and S species indicate electron depletion within ZnS, supporting electron transfer from ZnS to the TiO2 at the heterojunction interface, confirming the formation of a TiO2/ZnS heterojunction.

2.2. Photocatalyst Activity

2.2.1. Visible Light Degradation of TC

To assess the photocatalytic performance of TiO2, ZnS, and TZnS−100, degradation experiments were conducted on a 20 mg/L TC solution under visible light irradiation. Dark adsorption tests confirmed all catalysts reached adsorption equilibrium within 30 min, with no TC self-degradation observed in control experiments (Figure S6). As shown in Figure 5a (−30 min as the dark adsorption equilibrium point and 0 min as the start of irradiation), TZnS−100 achieved 80% TC degradation within 20 min, demonstrating superior activity over pristine TiO2 (55% degradation, 1.45 times) and ZnS (50% degradation, 1.60 times). UV-vis spectra after 20 min of visible light irradiation (Figure 5b) show residual TC concentration: TZnS−100 < ZnS < TiO2 (at λ = 357 nm, relative to TC control). This enhancement stems from the TiO2/ZnS N-N heterojunction, which enhances charge separation kinetics. Interfacial synergy facilitates efficient photogenerated carrier migration through the heterostructure, while maintaining ample surface active sites, as evidenced by the BET surface area trend: TZnS−100 > ZnS > TiO2 (Figure S4 and Table S1).
To evaluate recoverability, TZnS−100 was separated after each degradation cycle, washed with deionized water, dried, and reused. After the fifth cycle, TZnS−100 retained over 94% of its initial activity (Figure 5c), for surpassing the 20% drop of pure ZnS (Figure S7a) and 30% drop of pure TiO2 (Figure S7b) within fifth cycles, thereby demonstrating outstanding recyclability despite a marginal 6% efficiency decrease, likely due to cumulative catalyst loss during recovery and mechanical wear. Structural stability was verified by XRD (Figure 5d), with unchanged diffraction patterns for TZnS−100 after cycling, indicating excellent phase stability during photocatalytic operation.

2.2.2. Effect of Different Reaction Conditions

Optimal TiO2 Loading for TZnS−100 Synthesis. Systematic optimization identified 100 mg of TiO2 as the optimal loading for TZnS synthesis, maximizing TC degradation efficiency (Figure 6a). This results from balanced TiO2/ZnS interfacial contact, enhancing heterojunction charge transfer. Excessive loading (300 mg) reduced efficiency by 9% within 100 min (Figure 6b) due to TiO2 nanoparticle aggregation, blocking active sites, and causing light scattering.
Concentration Effects. TZnS−100 maintained over 96% TC degradation at ≤20 mg/L (Figure 6c), but efficiency dropped above 30 mg/L. This degradation limitation stems from three interconnected mechanisms: (1) Mass transport limitations reduce TC diffusion; (2) Photon absorption saturation at high pollutant concentrations; (3) Radical scavenging by TC degradation intermediates.
Ionic Interference Resistance. Common ions ( N a + , K + , C l , and S O 4 2 ) caused less than 5% efficiency variations (Figure 6d), thereby demonstrating the exceptional ionic interference resistance of the TZnS−100 catalyst.
pH Dependence. Neutral conditions (pH = 7) maximized degradation (99.0%), while acidic (pH = 3) and alkaline (pH = 11) conditions reduced efficiency by 20–40% (Figure 6e) due to surface charge effects: protonation repels TC in acidic conditions, and hydroxide consumes photogenerated holes and · OH radicals in alkaline conditions.
Water Matrix Performance. TZnS−100 maintained over 96.0% TC removal within 90 min across complex water matrices (Figure 6f), including tap water (96.2%), mineral water (96.5%), and Fuyang Lake water (96.1%). Deionized water showed slightly higher efficiency (99.0%) due to the absence of natural organic matter competing for reactive species and light absorption, confirming resilience to carbonates and silicates.

2.3. Possible Photocatalytic Mechanism

2.3.1. Active Species Determination

Active species trapping experiments systematically elucidated the photocatalytic degradation mechanism of TZnS−100 for TC (Figure 7a,b). Quenching hydroxyl radicals ( · OH) with isopropyl alcohol (IPA) reduced degradation efficiency by 85.7%, indicating that · OH contributes to secondary oxidation via H-atom abstraction. Scavenging holes ( h + ) with ammonium oxalate (AO) caused an 81.6% efficiency loss, demonstrating hole directly oxidizes TC via nucleophilic attack. Quenching superoxide radicals ( · O 2 ) with benzoquinone (BQ) resulted in a 56.2% inhibition, confirming · O 2 as the primary reactive species due to electrophilic addition to TC’s conjugated systems, amplified by radical chain reactions [20].
ESR spectroscopy with DMPO spin-trapping agents confirmed the photocatalytic generation of reactive oxygen species in ZnS and TZnS−100 heterostructure [35]. Under visible-light irradiation from a xenon lamp, distinct DMPO- · O 2 and DMPO- · O H signals appeared in the 3455–3555 mT magnetic field range (Figure 7c,d), absent in dark conditions, establishing the photo-driven radical formation.
Importantly, the DMPO- · O 2 signal intensity in TZnS−100 significantly exceeded that observed for pristine ZnS, indicating enhanced superoxide radical ( · O 2 ) production resulting from interfacial charge separation [36]. Specifically, photogenerated holes ( h + ) preferentially migrate toward ZnS due to its higher valence band position (−0.81 eV vs. NHE vs. −0.63 eV for TiO2), thereby reducing electron-hole recombination in the TiO2 core. This prolonged electron lifetime (>3.2 ns, measured by time-resolved photoluminescence, Figure 8c) facilitates efficient oxygen reduction reactions (O2 +   e · O 2 ) at the TiO2 interface.
Additionally, a weaker ESR signal of DMPO- · O H over ZnS was observed compared with TZnS−100, suggesting a limited generation of hydroxyl radicals in the photocatalytic reaction. These results confirm the synergistic effect of carrier separation in the TiO2/ZnS heterojunction, promoting TC photocatalytic degradation, which is consistent with active species trapping results.

2.3.2. Charge Transfer Kinetics

To systematically elucidate interfacial charge dynamics in the TZnS−100 heterostructure, complementary optoelectronic characterizations were hierarchically conducted under visible-light irradiation, establishing a progressive evidence chain for enhanced interfacial charge transfer.
First, transient photocurrent analysis (Figure 8a) confirmed the overall charge separation capability of the heterostructure. Notably, TZnS−100 exhibited the highest transient photocurrent density, surpassing that of pure ZnS and pure TiO2, respectively, demonstrating superior charge generation and interfacial transfer efficiency.
Subsequently, steady-state photoluminescence (PL) spectroscopy, performed using a FluoroMax-4 fluorescence spectrophotometer equipped with a 500 W xenon lamp light source (Figure 8b; λ = 320 nm), provided direct evidence of recombination suppression. Specifically, the lower PL intensity in TZnS−100 corroborates reduced electron-hole recombination through interfacial charge separation, thereby reinforcing the photocurrent findings.
Furthermore, time-resolved PL (TRPL, Figure 8c) quantitatively validated recombination kinetics. TZnS−100 ( τ a v g 3.57   n s ) exhibited an average carrier lifetime 42% longer than ZnS ( τ a v g 2.51   n s ) and 86% longer than TiO2 (1.92 ns), indicating significantly prolonged carrier availability. Critically, these data not only align with PL quenching trends but also establish that the TiO2/ZnS heterojunction suppresses recombination.
Consequently, electrochemical impedance spectroscopy (EIS, Figure 8d) mechanistically explained the observed phenomena. The smaller Nyquist arc radius in TZnS−100 unequivocally indicates minimized charge-transfer resistance, originating from interfacial synergy in the TiO2/ZnS N-N heterojunction [37], which simultaneously: (1) facilitates rapid carrier transport; (2) reduces radiative recombination; and (3) ultimately yields amplified photocatalytic efficiency.
To elucidate the photocatalytic mechanism, the band structure of TiO2, ZnS, and TZnS−100 was systematically characterized. Tauc plot analysis of UV–vis absorption spectra (Figure S5a) revealed distinct band gap variations: pristine TiO2 exhibited a band gap of 3.15 eV, and pure ZnS showed 3.30 eV [38]. Correspondingly, Figure S5b demonstrates a 12 nm blue shift in the absorption edge of TZnS−100 relative to TiO2, primarily attributed to the quantum confinement effect of the deposited ZnS nanoparticles [39]. This shift reflects successful TiO2/ZnS heterojunction integration, elevating the conduction band minimum energy, consistent with density functional theory (DFT) calculations confirming increased effective masses at the heterojunction boundary [40].
Mott–Schottky analysis quantitatively determined the interfacial energetics of the semiconductor components. The characteristic positive slopes in Mott–Schottky plots (Figure 8e,f) conclusively verify n-type behavior for both anatase TiO2 and synthesized ZnS [41,42]. The flat band potentials of anatase TiO2 and ZnS are −0.63 V and −0.81 V (vs. NHE, pH = 7; vs. Ag/AgCl), respectively. For n-type semiconductors, the E f b resides approximately 0.2 V negative relative to the conduction band minimum ( E C B ) due to pH-dependent Fermi level pinning. Using the relationship E V B =   E g   +   E C B [43], the conduction band (CB) and valence band (VB) energies for anatase TiO2 were calculated as −0.83 eV/2.32 eV, and for ZnS as −1.01 eV/2.29 eV. The resulting band alignment diagram (Figure 9) reveals a type II heterojunction between TiO2 and ZnS, facilitating efficient separation of photogenerated charge carriers through their staggered band offsets [44].

2.3.3. Photocatalytic Mechanism

The photocatalytic mechanism of the TZnS−100 heterojunction, consistent with experimental characterizations, is illustrated in Figure 9. The N-N type heterojunction formed between TiO2 and ZnS exhibits optimal band alignment. Under visible light irradiation, photogenerated electron—hole pairs are produced in both semiconductors through excitation from the valence band (VB) to the conduction band (CB), facilitating effective charge separation via the type—II heterojunction structure [45,46]. The favorable band alignment at the TiO2/ZnS heterojunction promotes efficient interfacial charge separation: photogenerated electrons transfer from the ZnS CB to the TiO2 CB, while holes remain in the ZnS VB. This spatial separation induces an interfacial electric field, significantly suppressing electron-hole recombination, as evidenced by the enhanced photocurrent response (Figure 8a) and reduced charge transfer resistance in EIS (Figure 8d). Within the reaction system, electrons accumulated in the TiO2 CB participate in oxygen reduction to form superoxide radical anions ( · O 2 ): TiO2 (CB) e + O 2     · O 2 . Meanwhile, holes in the ZnS VB drive water oxidation to generate hydroxyl radicals ( · O H ): ZnS (VB) h + + H 2 O · O H + O H . The synergistic action of · O 2 and · O H induces stepwise degradation of TC: the radical initially attacks the aromatic ring, disrupting the conjugated π systems, followed by ring-opening reactions to form intermediates, ultimately mineralizing into nontoxic inorganic products such as CO2 and H2O. Therefore, the constructed TiO2/ZnS heterojunction conforms to the conventional type—II charge transfer mode.

3. Experimental Section

3.1. Chemicals

Anhydrous ethanol (C2H5OH), tetrabutyl titanate (TBT), zinc acetate dihydrate (Zn(CH3COO)2·2H2O), thiourea (CH4N2S), hexadecyltrimethylammonium bromide (CTAB, 99%), nitorblue tetrazolium chloride (NBT), t-BuOH, tetracycline hydrochloride (TC), and triethanolamine (TEOA) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). All chemicals were used as received without further purification.

3.2. Synthesis of TiO2/ZnS

The hollow TiO2 sphere was synthesized via template-directed self-assembly, following our previously established methods [47,48]. Initially, 3 g of tetrabutyl titanate and 2 g of cationic polystyrene spheres (CPS) were added to 50 mL of ethanol and vigorously stirred for 24 h. Subsequently, a solution containing 1 mL of H2O and 9 mL of ethanol was introduced to the reaction system to initiate the hydrolysis reaction, which proceeded at an ice bath for 24 h. Finally, the samples were washed with deionized water and ethanol, dried at 60 °C for 3 h, and then calcined at 450 °C for 2 h to yield the hollow TiO2 spheres.
The TiO2/ZnS composite was synthesized using a hydrothermal reaction with pre-prepared TiO2 as the substrate. First, TiO2 (10, 30, 50, 100, or 300 mg) and 0.01 mol of Zn(CH3COO)2·2H2O were dispersed in 10 mL of ethanol using sonication for 30 min to prepare solution A. Secondly, 0.01 mol of CTAB and 0.01 mol of thiourea (CH4N2S) were added to 40 mL of ethanol and sonicated for 10 min to form solution B. Solution A was slowly introduced into solution B using an injection method (10 μL per second). The resulting mixture was vigorously stirred for 120 min, before being transferred to a 100 mL sealed autoclave and heated at 140 °C for 4 h.
Finally, the samples were washed three times with deionized water and ethanol, and then dried at 60 °C to remove residual solvent, yielding the TiO2/ZnS composite samples (denoted as TZnS-X, where X corresponds to the initial TiO2 mass: 10, 30, 50, 100, 300).
Pure ZnS catalysts were prepared using the same method without adding TiO2.

3.3. Photocatalyst Characterization

The microscopic morphology and structure of the TZnS photocatalysts were characterized using scanning electron microscopy (SEM, JEOL JSM-7001F, Tokyo, Japan) and transmission electron microscopy (TEM, JEOL JEM-210HR). The crystal composition and crystallization degree of the catalysts were determined by X-ray diffraction (XRD, Rigaku D/max2500V, Tokyo, Japan). The light absorption range of the TZnS photocatalyst was determined using a UV-visible diffuse reflectance spectrometer (DRS, Shimadzu UV-2600, Kyoto, Japan) with a scanning range of 200—800 nm. The surface elemental state of the photocatalyst was analyzed using X-ray photoelectron spectroscopy (XPS, AMICUS, Berkshire, UK). Finally, the photogenerated radical signals were assessed using an electron paramagnetic resonance instrument (EPR, Bruker ESR 300, Billerica, MA, USA).

3.4. Photocatalytic Activity Test

The photodegradation experiment utilized the TZnS photocatalyst to treat a tetracycline hydrochloride (TC) solution (20 mg/L) as the target antibiotic contaminant. Visible light irradiation (λ > 420 nm) was provided by a 300 W xenon lamp (Beijing Perfectlight PLS-SXE300, Beijing, China). Experimental procedure: Initially, 10 mg of TZnS catalyst was added to 50 mL of the TC solution and sonicated for 3 min. The suspension was then magnetically stirred in the dark for 30 min to establish adsorption—desorption equilibrium. Subsequently, the xenon lamp was activated to initiate visible—light—driven photodegradation. At predetermined intervals, 2 mL aliquots were sampled and immediately filtered through a 0.22 μm filter to obtain a clear solution. The TC degradation efficiency was measured by analyzing the residual solution’s absorbance at 357 nm using a UV-Vis spectrophotometer (Shimadzu UV-2600). After each experiment, the remaining catalyst was filtered, washed, and dried for subsequent cycle tests.

3.5. Electrochemical Testing

Electrochemical measurements were conducted in a 0.5 M Na2SO4 electrolyte using a CHI660B electrochemical workstation (CH Instruments, Austin, TX, USA) with a standard three-electrode system. A platinum sheet served as the counter electrode, an Ag/AgCl reference electrode (saturated KCl) was employed for potential measurement, and ITO-coated glass (1 cm2 active area) served as the working electrode.
The working electrode was prepared as follows: 5 mg of TZnS active material was uniformly dispersed in a mixed solvent of 0.1 mL 5 wt% Nafion solution and 0.5 mL ethylene glycol, followed by sonication for 30 min to ensure homogeneity. A 40 μL aliquot of the suspension was precisely dispensed onto the ITO substrate using a micropipette, followed by thermal annealing at 60 °C for 8 h to form a uniform thin film. The electrode was cooled to room temperature before electrochemical testing.

4. Conclusions

In summary, we developed a rationally designed TZnS−100 heterojunction photocatalyst optimized for enhanced TC degradation under visible light. The TZnS−100 composite demonstrated superior photocatalytic activity, achieving 80% TC degradation within 20 min, representing 1.55−fold and 1.50−fold efficiency improvements compared to pristine TiO2 and ZnS, respectively. In addition, the optimal photocatalytic performance was achieved by incorporating 100 mg of TiO2 (prepared via the hydrothermal method) into the TZnS−100 composite. Time-resolved photoluminescence spectroscopy confirmed that the TiO2/ZnS heterojunction significantly prolonged charge carrier lifetimes (TZnS−100   τ a v g 3.57   n s vs. TiO2:   τ a v g 1.92   n s ), indicating efficient separation of photo-generated carriers. Further evidence from ESR experiments revealed enhanced generation of superoxide radicals ( · O 2 ) under visible light irradiation (DMPO− · O 2 quartet signal), with negligible radical activity in dark conditions. Moreover, the photocatalytic mechanism was determined to proceed via a type−II heterojunction model. This work not only provides a promising strategy for designing hollow heterojunction photocatalysts but also offers in−depth mechanistic insights into the critical role of superoxide radical anions in the photocatalytic degradation, providing theoretical guidance for the design of efficient photocatalytic systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30173644/s1, Figure S1: (a–f) TEM, the corresponding element mappings, HRTEM images of TiO2. Figure S2: (a–f) TEM, the corresponding element mappings, HRTEM images of ZnS. Figure S3: HRXPS spectra of T, Z and TZ: (a) full spectra, (b) C1s. Figure S4: BET spectra of (a) TiO2, ZnS, TZnS, and (b) TZnS-X (X = 10, 30, 50, 100, 300). Table S1: Surface area TiO2, ZnS and TZnS-X (X = 10, 30, 50, 100, 300). Figure S5: (a) DRS spectra and (b) Band gad of TiO2, ZnS, TZnS. Figure S6: Adsorption studies for TiO2, ZnS and TZnS under dark irradiation: Catalytic conditions with a catalyst concentration of 200 mg/L and a target compound (TC) concentration of 20 mg/L. Figure S7: Recyclability test of (a) ZnS and (b) TiO2.

Author Contributions

Formal analysis, Y.D.; Investigation, A.S.; Data curation, Y.W.; Writing—original draft, Y.Z.; Writing—review & editing, Y.Z.; Project administration, Y.T.; Funding acquisition, J.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Natural Science Research Key Project of the Anhui Provincial Department of Education (Grant No. 2022AH051333); the Excellent Young Teacher Cultivation Project of the Anhui Provincial Department of Education (Grant No. YQZD2024027); the Scientific Research Project of Fuyang Normal University (Grant No. 2022FSKJ05ZD); the Key Program for Anhui Provincial Key Laboratory of Green Carbon Chemistry (Grant No. AHGC2025005 and AHGC2025004); the National College Student Innovation and Entrepreneurship Training Program (Anhui Province, Grant No. S202410371082, S202410371087, and S202510371130); the Natural Science Foundation of Anhui Province (Grant No. 2008085QB93); and the Horizontal Cooperation Project of Fuyang Municipal Government and Fuyang Normal University (Grant No. SXHZ202207).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ding, D.; Wang, B.; Zhang, X.; Zhang, J.; Zhang, H.; Liu, X.; Gao, Z.; Yu, Z. The spread of antibiotic resistance to humans and potential protection strategies. Ecotoxicol. Environ. Saf. 2023, 254, 114734. [Google Scholar] [CrossRef]
  2. Wu, X.D.; Zhang, X.Y.; Liao, H.Z.; Guo, J.; Ma, Z.H.; Fu, Z.L. Microplastics and tetracycline affecting apoptosis, enzyme activities and metabolism processes in the Aurelia aurita polyps: Insights into combined pollutant effects. Front. Mar. Sci. 2025, 12, 1545131. [Google Scholar] [CrossRef]
  3. Zhou, P.F.; Han, X.Y.; Wu, L.X.; Qi, M.Y.; Shen, Y.B.; Nie, J.N.; Liu, X.N.; Wang, F.; Bian, L.; Liang, J.S. Facile construction of novel BiOBr/black-TiO2/tourmaline composites for the synergistic degradation of tetracycline in aqueous. Sep. Purif. Technol. 2025, 362, 131976. [Google Scholar] [CrossRef]
  4. Li, M.; Liu, Y.; Yang, S.Y.; Zhang, Y.; Wei, L.; Zhu, B.C. Efficient charge carrier transfer in TiO2/CulnS2 S-scheme heterojunction to boost photocatalytic degradation of tetracycline hydrochloride. J. Mater. Sci. Technol. 2025, 224, 245–256. [Google Scholar] [CrossRef]
  5. Balayeva, N.O.; Mamiyev, Z.; Dillert, R.; Zheng, N.; Bahnemann, D.W. Rh/TiO2-photocatalyzed acceptorless dehydrogenation of N-heterocycles upon visible-light illumination. ACS Catal. 2020, 10, 5542–5553. [Google Scholar] [CrossRef]
  6. Xu, H.Y.; Zhang, L.; Huang, K.Q.; Sun, D.C.; Tong, X.; Li, X.T.; Yang, Y.; Ji, X. Novel integrating sphere structure solar photocatalytic reactor design: Efficient conversion of nitrate wastewater into high-value ammonia. Renew. Energy 2025, 243, 122611. [Google Scholar] [CrossRef]
  7. Wang, W.B.; Liu, J.; Cui, M.X.; Li, X.H.; Niu, L.Y.; Wu, X.; Chen, Y.N.; Li, X.W.; Zhu, H.C.; Wang, D.; et al. Z-type ZnIn2S4 homojunction for high performance photocatalytic hydrogen evolution. Chem. Eng. J. 2025, 507, 160370. [Google Scholar] [CrossRef]
  8. Haghighi, P.; Haghighat, F. TiO2-based photocatalytic oxidation process for indoor air VOCs removal: A comprehensive review. Build. Environ. 2024, 249, 111108. [Google Scholar] [CrossRef]
  9. Ferreira, O.; Gil, M.; Carapeto, A.; Barreto, A.; Pereira, M.; Rodrigues, M.S.; Habib, L.; Gomes, L.C.; Teixeira-Santos, R.; Mergulhao, F.J.M.; et al. An engineered, biocide-grafted TiO2-based nanohybrid material with enhanced photocatalytic and antimicrobial activity. J. Clean. Prod. 2025, 494, 144937. [Google Scholar] [CrossRef]
  10. Zhao, Y.; Linghu, X.; Shu, Y.; Zhang, J.; Chen, Z.; Wu, Y.; Shan, D.; Wang, B. Classification and catalytic mechanisms of heterojunction photocatalysts and the application of titanium dioxide TiO2-based heterojunctions in environmental remediation. J. Environ. Chem. Eng. 2022, 10, 108077. [Google Scholar] [CrossRef]
  11. Mamiyev, Z.; Balayeva, N.O. Metal sulfide photocatalysts for hydrogen generation: A review of recent advances. Catalysts 2022, 12, 1316. [Google Scholar] [CrossRef]
  12. Lu, R.; Zhao, S.Z.; Yang, Y.; Wang, Y.; Huang, H.L.; Hu, Y.D.; Rodriguez, R.D.; Chen, J.J. Efficient adsorption and photocatalytic degradation of organic contaminants using TiO2/(Bi2O3/Bi2O2.33) nanotubes. Mater. Today Chem. 2023, 32, 101641. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Bo, X.; Zhu, T.; Zhao, W.; Cui, Y.; Chang, J. Synthesis of TiO2-ZnO n-n heterojunction with excellent visible light-driven photodegradation of tetracycline. Nanomaterials 2024, 14, 1802. [Google Scholar] [CrossRef]
  14. Wang, Q.P.; Wang, G.H.; Wang, J.; Li, J.M.; Wang, K.; Zhou, S.; Su, Y.R. In situ hydrothermal synthesis of ZnS/TiO2 nanofibers S-scheme heterojunction for enhanced photocatalytic H2 evolution. Adv. Sustain. Syst. 2023, 7, 2200027. [Google Scholar] [CrossRef]
  15. Zhang, Y.; Zhang, Q.; Xing, K.; Xiao, Y.; Zhao, M.; Cun, Y. Construction of Z-scheme heterojunction LaCoO3 modified ZnO for photocatalytic degradation of tetracycline under visible light irradiation. Catal. Commun. 2024, 186, 106815. [Google Scholar] [CrossRef]
  16. Xie, Y.; Wang, T.W.; Li, H.Q.; Sun, C.Z.; Hu, J.; Cao, S.S. Generalized synthetic strategies toward oxygen vacancy-enriched ZnO-ZnS hollow porous spheres with enhanced photocatalytic hydrogen evolution. Inorg. Chem. 2025, 64, 3563–3571. [Google Scholar] [CrossRef]
  17. Ma, F.; Xu, X.Y.; Huo, C.; Sun, C.Z.; Li, Q.; Yin, Z.L.; Cao, S.S. Dual heterogeneous structures promote electrochemical properties and photocatalytic hydrogen evolution for inverse opal ZnO/ZnS/Co3O4 Crystals. Inorg. Chem. 2024, 63, 8782–8790. [Google Scholar] [CrossRef]
  18. Wu, D.; He, Y.X.; Lin, C.; Li, B.; Ma, J.P.; Ruan, L.J.; Feng, Y.J.; Ban, C.G.; Ding, J.J.; Wang, X.X.; et al. Atmosphere-driven metal-support synergy in ZnO/Au catalysts for efficient piezo-catalytic hydrogen evolution. J. Mater. 2025, 11, 100959. [Google Scholar] [CrossRef]
  19. Wang, D.; Zhang, H.Y.; Yang, P.; Zhang, X. SnO2 nanoparticles endowed Pt-TiO2 bowl nanoarraysenhanced redox ability towards efficient tetracycline hydrochlorideremoval and H2 generation. Sustain. Mater. Technol. 2025, 44, e01358. [Google Scholar] [CrossRef]
  20. Jiang, Y.C.; Zhu, K.R.; Zhang, S.Y.; Wang, D.Z. Hierarchically porous ZnS/C photocatalyst toward tetracycline hydrochloride degradation under visible light: Efficiency and mechanism. J. Water Process Eng. 2025, 70, 107074. [Google Scholar] [CrossRef]
  21. Chen, Q.Q.; Yang, X.B.; Fu, X.P.; Pan, J.J.; Zhou, J.M.; Zhou, G.W.; Cheng, K.J. Fabrication of a novel Ag2S-ZnS/ZIF-8 hollow heterojunction by in-situ formation of Ag2S and ZnS on ZIF-8 with enhanced tetracycline degradation performance. Chem. Eng. J. 2024, 499, 156242. [Google Scholar] [CrossRef]
  22. Chen, L.; Chen, X.; Xiao, J. Flower-like Ti3C2T-TiO2 modified with ZnS nanoparticles as adsorption-catalytic cathodic material for lithium-sulfur batteries. J. Energy Storage 2025, 118, 116241. [Google Scholar] [CrossRef]
  23. Li, M.; Liu, L.; Li, Y.; Li, Y. Tuning piezoelectric driven photocatalysis by TiO2/ZnS heterojunction for mineral processing wastewater degradation. J. Water Process Eng. 2024, 64, 105727. [Google Scholar] [CrossRef]
  24. Liu, D.D.; Liang, H.; Xu, T.; Bai, J.; Li, C.P. Construction of ternary hollow TiO2-ZnS@ZnO heterostructure with enhanced visible-light photoactivity. J. Mol. Struct. 2022, 1248, 131493. [Google Scholar] [CrossRef]
  25. Li, C.Y.; Wu, J.H.; Wang, X.Z.; Cai, Y.X.; Jia, R.R.; Wang, W.W.; Xia, S.S.; Li, L.; Chen, Z.; Jin, C.C. One-step preparation of Z-scheme g-C3N4 nanorods/TiO2 microsphere with improved photodegradation of antibiotic residue. Ceram. Int. 2024, 50, 42127–42134. [Google Scholar] [CrossRef]
  26. Malik, M.; Nazir, M.A.; Khan, H.M.; Shah, S.S.A.; Anum, A.; Rehman, A.U.; Hashem, A.; Avila-Quezada, G.D.; Abd Allah, E.F. Light harvesting S-scheme g-C3N4/TiO2/M photocatalysts for efficient removal of hazardous moxifloxacin. Diam. Relat. Mater. 2024, 150, 111673. [Google Scholar] [CrossRef]
  27. Zhang, H.; Sun, P.; Fei, X.; Wu, X.; Huang, Z.; Zhong, W.; Gong, Q.; Zheng, Y.; Zhang, Q.; Xie, S.; et al. Unusual facet and co-catalyst effects in TiO2-based photocatalytic coupling of methane. Nat. Commun. 2024, 15, 4453. [Google Scholar] [CrossRef]
  28. Cui, Y.; Ru, G.; Zhang, T.; Yang, K.; Liu, S.; Qi, W.; Ye, Q.; Liu, X.; Zhou, F. Schottky interface engineering in Ti3C2Tx/ZnS organic hydrogels for high-performance multifunctional flexible absorbers. Adv. Funct. Mater. 2024, 35, 2417346. [Google Scholar] [CrossRef]
  29. Wang, H.; Qi, H.F.; Sun, X.; Jia, S.Y.; Li, X.Y.; Miao, T.J.; Xiong, L.Q.; Wang, S.H.; Zhang, X.L.; Liu, X.Y.; et al. High quantum efficiency of hydrogen production from methanol aqueous solution with PtCu-TiO2 photocatalysts. Nat. Mater. 2023, 22, 619–626. [Google Scholar] [CrossRef] [PubMed]
  30. Wu, P.; Liu, H.; Xie, Z.; Xie, L.; Liu, G.; Xu, Y.; Chen, J.; Lu, C.Z. Excellent charge separation of NCQDs/ZnS nanocomposites for the promotion of photocatalytic H2 evolution. ACS Appl Mater Interfaces 2024, 16, 16601–16611. [Google Scholar] [CrossRef]
  31. Wen, D.; Feng, J.; Deng, R.; Li, K.; Zhang, H. Zn/Pt dual-site single-atom driven difunctional superimposition-augmented sonosensitizer for sonodynamic therapy boosted ferroptosis of cancer. Nat. Commun. 2024, 15, 9359. [Google Scholar] [CrossRef]
  32. Han, C.; Zeng, Z.; Zhang, X.; Liang, Y.; Kundu, B.K.; Yuan, L.; Tan, C.-L.; Zhang, Y.; Xu, Y.-J. All-in-one: Plasmonic janus heterostructures for efficient cooperative photoredox catalysis. Angew. Chem. Int. Ed. 2024, 63, e202408527. [Google Scholar] [CrossRef]
  33. Ban, C.G.; Wang, Y.; Feng, Y.J.; Zhu, Z.H.; Duan, Y.Y.; Ma, J.P.; Zhang, X.; Liu, X.; Zhou, K.; Zou, H.J.; et al. Photochromic single atom Ag/TiO2 catalysts for selective CO2 reduction to CH4. Energy Environ. Sci. 2024, 17, 518–530. [Google Scholar] [CrossRef]
  34. Chen, J.; Jia, Y.; Zhao, H.; Wu, D.; Du, Y.; Ren, X.; Wei, Q. Ternary heterojunction ZnS-ZnIn2S4-In2S3 efficiently catalyzes triethylamine to enhance electrochemiluminescence imaging signals. Adv. Funct. Mater. 2024, 35, 2420714. [Google Scholar] [CrossRef]
  35. Li, X.; Li, C.; Xu, Y.; Liu, Q.; Bahri, M.; Zhang, L.; Browning, N.D.; Cowan, A.J.; Tang, J. Efficient hole abstraction for highly selective oxidative coupling of methane by Au-sputtered TiO2 photocatalysts. Nat. Energy 2023, 8, 1013–1022. [Google Scholar] [CrossRef]
  36. Sun, X.; Chen, X.; Fu, C.; Yu, Q.; Zheng, X.S.; Fang, F.; Liu, Y.; Zhu, J.; Zhang, W.; Huang, W. Molecular oxygen enhances H2O2 utilization for the photocatalytic conversion of methane to liquid-phase oxygenates. Nat. Commun. 2022, 13, 6677. [Google Scholar] [CrossRef]
  37. Zhu, Y.; Wang, C.; Ai, C.; Xiang, Y.; Mao, C.; Qiao, W.; Wu, J.; Kubi, J.A.; Liu, X.; Wu, S.; et al. Photocurrent-directed immunoregulation accelerates osseointegration through activating calcium influx in macrophages. Adv. Funct. Mater. 2024, 34, 2406095. [Google Scholar] [CrossRef]
  38. Huang, F.; Wang, F.; Liu, Y.; Guo, L. Cu-ZnS modulated multi-carbon coupling enables high selectivity photoreduction CO2 to CH3CH2COOH. Adv. Mater. 2025, 37, e2416708. [Google Scholar] [CrossRef]
  39. Mamiyev, Z.Q.; Balayeva, N.O. Optical and structural studies of ZnS nanoparticles synthesized via chemical in situ technique. Chem. Phys. Lett. 2016, 646, 69–74. [Google Scholar] [CrossRef]
  40. Lee, H.K.; Kim, T.; Jang, Y.A.; Jeong, Y.; Lee, S.W.; Park, C.H. Near-infrared emissive CuInS2/ZnS quantum dot-embedded polymer scaffolds for photon upconversion imaging. Adv. Mater. 2025, 37, e2502333. [Google Scholar] [CrossRef]
  41. Lin, J.; Liao, J.; Dagnaw, F.W.; Li, J.; Xie, L.; Zhou, M.; Liu, C.; Jian, J.; Tong, Q. Silicon-based heterojunction photoanodes modified with copper-bipyridine catalyst for enhanced solar-driven water splitting. Appl. Surf. Sci. 2024, 648, 159089. [Google Scholar] [CrossRef]
  42. Jin, Y.; Yu, W.; Chen, L.; Yuan, R.; Liu, J.; Fu, Y.; Chai, Y. Dual-sensitized heterojunction Ag2S/ZnS/NiS composites with entire visible-light region absorption for ultrasensitive photoelectrochemical detection of tobramycin. Biosens. Bioelectron. 2024, 260, 116459. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, B.; Liu, Y.; Zhao, Y.; Liu, E.; Li, Z.; Miao, Z. Bioinspired mortise-and-tenon stacking supramolecular engineering for efficient carrier separation in S-scheme heterojunctions. Adv. Funct. Mater. 2025, 2507648. [Google Scholar] [CrossRef]
  44. Wang, J.; Li, S.; Yang, Q.; Liu, Z.; Tang, F.; Yang, X. Synthesis of In2O3/ZnS type II heterojunctions with superior photocatalytic activity: Mechanism and application in tetracycline hydrochloride degradation. Colloids Surf. A Physicochem. Eng. Asp. 2025, 723, 137306. [Google Scholar] [CrossRef]
  45. Yu, S.; Li, C.; Lin, Y.; Zhang, J.; Liu, Y.; Yu, F. Interfacial N-Ti bond modulated COFs-TiO2 type-II heterojunctions with directional charge transfer for efficient photocatalytic uranium reduction. Sep. Purif. Technol. 2024, 341, 126888. [Google Scholar] [CrossRef]
  46. Shabbir, A.; Sardar, S.; Mumtaz, A. Mechanistic investigations of emerging type-II, Z-scheme and S-scheme heterojunctions for photocatalytic applications-a review. J. Alloys Compd. 2024, 1003, 175683. [Google Scholar] [CrossRef]
  47. Xiao, Y.G.; Wu, J.J.; Li, H.Q.; Liu, X.; Li, Q.; Zhang, Y.; Cao, S.S. In-situ regulation of hollow TiO2 with tunable anatase/rutile heterophase homojunction for promoting excellent photocatalytic degradation of pesticides. Ceram. Int. 2023, 49, 38070–38080. [Google Scholar] [CrossRef]
  48. Ma, J.J.; Xiao, Y.G.; Chen, J.R.; Shen, Y.; Xiao, S.S.; Cao, S.S. Dual-pathway charge transfer mechanism of (anatase/rutile TiO2-Ag3PO4 )2-Ag3PO4 hollow photocatalyst promotes efficient degradation of pesticides. J. Colloid Interface Sci. 2025, 678, 334–344. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of the synthesis of TiO2/ZnS heterostructures.
Figure 1. Schematic illustration of the synthesis of TiO2/ZnS heterostructures.
Molecules 30 03644 g001
Figure 2. Morphological and Structural Characterization of TiO2, ZnS, and TZnS−100. (ac) SEM images of (a) TiO2, (b) ZnS and (c) TZnS, showing evolution from smooth TiO2 spheres to ZnS-decorated TiO2 surfaces; (d) EDS elemental mapping (Ti, Zn, S, O) of TZnS−100, showing homogeneous ZnS distribution on the TiO2 scaffold; (e) TEM image of a single TZnS−100 particle; (f) HRTEM image of TZnS−100, showing lattice fringes of 0.35 nm (TiO2 (101)) and 0.31 nm (ZnS (111)), confirming intimate interfacial contact. All scale bars: (ac) 1 μm; (d) 100 nm; (f) 10 nm, and 1 nm.
Figure 2. Morphological and Structural Characterization of TiO2, ZnS, and TZnS−100. (ac) SEM images of (a) TiO2, (b) ZnS and (c) TZnS, showing evolution from smooth TiO2 spheres to ZnS-decorated TiO2 surfaces; (d) EDS elemental mapping (Ti, Zn, S, O) of TZnS−100, showing homogeneous ZnS distribution on the TiO2 scaffold; (e) TEM image of a single TZnS−100 particle; (f) HRTEM image of TZnS−100, showing lattice fringes of 0.35 nm (TiO2 (101)) and 0.31 nm (ZnS (111)), confirming intimate interfacial contact. All scale bars: (ac) 1 μm; (d) 100 nm; (f) 10 nm, and 1 nm.
Molecules 30 03644 g002
Figure 3. XRD patterns of TiO2, ZnS, and TZnS: (a) TZnS−100, TiO2, and ZnS; (b) TZnS-X with varying TiO2, mass percentage (X denotes the mass percentage of TiO2).
Figure 3. XRD patterns of TiO2, ZnS, and TZnS: (a) TZnS−100, TiO2, and ZnS; (b) TZnS-X with varying TiO2, mass percentage (X denotes the mass percentage of TiO2).
Molecules 30 03644 g003
Figure 4. HR-XPS spectra of TZnS−100, TiO2 and ZnS: (a) Zn 2p; (b) S 2p; (c) Ti 2p; (d) O 1s.
Figure 4. HR-XPS spectra of TZnS−100, TiO2 and ZnS: (a) Zn 2p; (b) S 2p; (c) Ti 2p; (d) O 1s.
Molecules 30 03644 g004
Figure 5. (a) Photocatalytic degradation of TC over TiO2, ZnS, TZnS−100; (b) UV−vis spectra; (c) Recyclability test of TZnS−100; (d) XRD patterns of TZnS−100 before and after recyclability test. Condition: catalyst concentration: 200 mg/L; TC concentration: 20 mg/L; total volume: 50 mL.
Figure 5. (a) Photocatalytic degradation of TC over TiO2, ZnS, TZnS−100; (b) UV−vis spectra; (c) Recyclability test of TZnS−100; (d) XRD patterns of TZnS−100 before and after recyclability test. Condition: catalyst concentration: 200 mg/L; TC concentration: 20 mg/L; total volume: 50 mL.
Molecules 30 03644 g005
Figure 6. Effects of Different Factors on TZnS−100 Photocatalytic Degradation Efficiency (a) As a function of TiO2 mass loading; (b) TC degradation after 100 min; (c) TC concentration; (d) Common cation and anion; (e) Solution pH; (f) Water source (Catalyst concentration: 200 mg/L; TC concentration: 20 mg/L; Reaction volume: 50 mL).
Figure 6. Effects of Different Factors on TZnS−100 Photocatalytic Degradation Efficiency (a) As a function of TiO2 mass loading; (b) TC degradation after 100 min; (c) TC concentration; (d) Common cation and anion; (e) Solution pH; (f) Water source (Catalyst concentration: 200 mg/L; TC concentration: 20 mg/L; Reaction volume: 50 mL).
Molecules 30 03644 g006
Figure 7. (a,b) Active species trapping experiments (TZnS−100 concentration: 200 mg/L; TC concentration: 20 mg/L; Total volume: 50 mL), (c) DMPO− · O 2 and (d) DMPO− · O H adduct (1:2:2:1 quartet) ESR spectra under dark and visible−light conditions.
Figure 7. (a,b) Active species trapping experiments (TZnS−100 concentration: 200 mg/L; TC concentration: 20 mg/L; Total volume: 50 mL), (c) DMPO− · O 2 and (d) DMPO− · O H adduct (1:2:2:1 quartet) ESR spectra under dark and visible−light conditions.
Molecules 30 03644 g007
Figure 8. (a) Photocurrent response; (b) Photoluminescence spectra; (c) Time−Resolved Photoluminescence; (d) Electrochemical Impedance Spectroscopy; (e,f) Mott−Schottky plots of TiO2, ZnS, and TZnS−100.
Figure 8. (a) Photocurrent response; (b) Photoluminescence spectra; (c) Time−Resolved Photoluminescence; (d) Electrochemical Impedance Spectroscopy; (e,f) Mott−Schottky plots of TiO2, ZnS, and TZnS−100.
Molecules 30 03644 g008
Figure 9. Schematic diagram of the photocatalytic degradation mechanism of TC by TZnS−100.
Figure 9. Schematic diagram of the photocatalytic degradation mechanism of TC by TZnS−100.
Molecules 30 03644 g009
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

Zhang, Y.; Su, A.; Ding, Y.; Wu, Y.; Tan, Y.; Chang, J. Construction of Hollow TiO2/ZnS Heterojunction Photocatalysts for Highly Enhanced Photodegradation of Tetracycline Hydrochloride. Molecules 2025, 30, 3644. https://doi.org/10.3390/molecules30173644

AMA Style

Zhang Y, Su A, Ding Y, Wu Y, Tan Y, Chang J. Construction of Hollow TiO2/ZnS Heterojunction Photocatalysts for Highly Enhanced Photodegradation of Tetracycline Hydrochloride. Molecules. 2025; 30(17):3644. https://doi.org/10.3390/molecules30173644

Chicago/Turabian Style

Zhang, Ying, Anhui Su, Yuqin Ding, Yuhan Wu, Yapeng Tan, and Jianguo Chang. 2025. "Construction of Hollow TiO2/ZnS Heterojunction Photocatalysts for Highly Enhanced Photodegradation of Tetracycline Hydrochloride" Molecules 30, no. 17: 3644. https://doi.org/10.3390/molecules30173644

APA Style

Zhang, Y., Su, A., Ding, Y., Wu, Y., Tan, Y., & Chang, J. (2025). Construction of Hollow TiO2/ZnS Heterojunction Photocatalysts for Highly Enhanced Photodegradation of Tetracycline Hydrochloride. Molecules, 30(17), 3644. https://doi.org/10.3390/molecules30173644

Article Metrics

Back to TopTop