Next Article in Journal
Correction: Ma, Y.; Song, T. A Bibliometric Review and Interdisciplinary Analysis of the Brahmaputra River. Water 2024, 16, 3115
Previous Article in Journal
Combined Effects of Carbon-to-Nitrogen (C/N) Ratio and Nitrate (NO3-N) Concentration on Partial Denitrification (PD) Performance at Low Temperature: Substrate Variation, Nitrite Accumulation, and Microbial Transformation
Previous Article in Special Issue
Addition of Heterotrophic Nitrification and Aerobic Denitrification Bacterial Agents to Enhance Bio-Nests Treating Low Carbon-to-Nitrogen Ratio Municipal Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of Visible-Light-Responsive TiO2/LDHs Heterostructures for Enhanced Photocatalytic Degradation Performance

Shandong Provincial Key Laboratory of Soil Conservation and Environmental Protection, College of Resources and Environment, Linyi University, Linyi 276000, China
*
Author to whom correspondence should be addressed.
Water 2025, 17(17), 2582; https://doi.org/10.3390/w17172582
Submission received: 31 July 2025 / Revised: 27 August 2025 / Accepted: 30 August 2025 / Published: 1 September 2025
(This article belongs to the Special Issue Science and Technology for Water Purification, 2nd Edition)

Abstract

A novel composite material comprising titanium dioxide and layered double hydroxides (TiO2/LDHs) was innovatively proposed and prepared using the co-precipitation method to overcome the shortcomings of titanium dioxide, such as low efficiency in separating electron–hole pairs induced by light and a low utilization rate of visible light. This material was used to study the visible-light-driven photocatalytic degradation of methylene blue. The experimental results show that by constructing efficient heterojunction structures through the alignment of interface band energies and regulating the interface charge transfer pathways, the recombination rate of photogenerated electron–hole pairs is significantly reduced, and the photocatalytic activity is greatly enhanced. Among the tested samples, the TiO2/LDHs composite material with an aluminum-to-titanium molar ratio of 1:1 (AT11) demonstrated the best photocatalytic performance. Within 70 min of simulated sunlight exposure, the degradation rate of methylene blue reached 98.2%, and the optimal concentration of the catalyst was 1 g/L. The photocatalytic process follows a first-order kinetic model. After four cycles of use, the degradation efficiency of methylene blue by the AT11 composite material was 78.93%, demonstrating good stability. The free radical capture experiments indicated that the main active substances for the photocatalytic degradation of methylene blue were h+ and ·OH. The constructed TiO2/LDHs heterostructure system significantly enhanced the photocatalytic performance of TiO2 materials, which was conducive to the efficient utilization of solar energy.

1. Introduction

Industrial advancement drives the widespread use of synthetic dyes in textiles, food processing, paper manufacturing, and pharmaceuticals. These compounds persist as major wastewater pollutants due to their molecular stability that resists standard treatments. Their polar groups enable water solubility while their chromophores impart color, collectively posing severe ecological threats to aquatic systems through persistent contamination [1,2,3,4]. These dye-derived organic pollutants exhibit carcinogenic, mutagenic, and toxic properties. When these compounds remain untreated, they jeopardize human health and ecological integrity. The effective elimination of such compounds from wastewater is thus critical for pollution control. Recent decades have seen advanced treatment technologies emerge, including adsorption [5,6], biological degradation [7,8], electrocatalysis [9], and photocatalysis [10,11]. Heterogeneous photocatalysis has emerged as a leading strategy for wastewater remediation, leveraging its potent oxidation power to mineralize diverse contaminants without secondary pollution. Its broad-spectrum efficacy against recalcitrant pollutants drives sustained research interest.
Conventional photocatalysts like TiO2 demonstrate robust photocatalytic activity. Under light excitation, they generate electron–hole pairs whose oxidation potential drives dye degradation, specifically oxidizing adsorbed molecules on the catalyst surface [12,13,14]. However, TiO2’s practical utility is constrained by its inherent limitations. Its wide bandgap (e.g., 3.2 eV in the anatase phase) confines its excitation at UV wavelengths, yielding negligible visible-light response and inefficient solar energy harvesting. Furthermore, TiO2 exhibits compromised photocatalytic efficiency due to synergistic limitations: weak surface adsorption capacity and rapid recombination of photogenerated charge carriers. These limitations reduce its performance under operational conditions [15,16,17,18]. To enhance the photocatalytic performance of TiO2, researchers employ three primary modification strategies: elemental doping [19,20], co-catalyst loading [21,22], and heterojunction engineering [23,24].
Doping modification can be classified into two types: metal doping and non-metal doping. Doping atoms are incorporated into the TiO2 crystal structure either by substitution (replacing Ti4+ atoms due to similar ionic radii) or by interstitial means (inserted or highly dispersed on the TiO2 surface in the form of single-nucleus complexes or clusters) [25]. The introduced ions affect the carrier recombination rate or bandgap width inside TiO2 by forming defects or altering the lattice type and ultimately influence the catalytic performance of the semiconductor [26]. Metal doping suppresses the recombination of photogenerated carriers on the surface of TiO2 and is currently the most widely used surface modification method. Metal ion doping includes transition metal ions, rare earth metal ions, and noble metal ions. Among them, the research on transition metal elements such as Pt [27], Ni [28], Zn [29], and Cr [30] is mainly focused on. Due to the multiple valence states of transition metals, the introduction of a small amount of transition metal ions can induce shallow trap energy levels, thereby enhancing electron capture and reducing carrier recombination. However, studies have shown that the doping of transition metal ions should be controlled within an appropriate range. When the amount of metal ions exceeds a certain limit, they become recombination centers, leading to an increase in the recombination rate of carriers [31,32]. The principle of non-metal doping is to mix the p orbitals of non-metals with the O 2p orbitals, thereby introducing impurity energy levels in the bandgap of TiO2 [25]. The valence band edge shifts upward, the bandgap narrows, and thus, the light response range of TiO2 redshifts, responding in the visible light range. Common non-metal doping elements include N [33], C [34], S [35], halogens [36]. They generally improve the photocatalytic performance of TiO2 by generating oxygen vacancies and narrowing the bandgap width.
The supported catalyst loads TiO2 onto various carriers (including carbon nanofibers, ceramics, etc.) [37,38] to solve problems such as the easy aggregation and difficult sedimentation of nanoparticles in a solution. Researchers have found that preparing magnetic composite photocatalysts by loading TiO2 onto magnetic carriers is expected to perfectly solve the problem of difficult separation and recovery [39]. Currently, the magnetic carriers that have been studied more extensively include Fe3O4, γ-Fe2O3, and spinel-type materials [40]. However, studies have found that direct contact between TiO2 and magnetic cores can reduce its photocatalytic activity. For instance, direct contact between Fe3O4 magnetic cores and TiO2 can trigger a “photo-dissolution phenomenon”, where Fe3O4 acts as an electron–hole recombination center, significantly reducing the photocatalytic activity and recyclability of the catalyst [39,40,41]. In addition, the performance of catalysts declines to varying degrees when loaded alone due to the significant reduction in the specific surface area, light absorption efficiency, and adsorption capacity of the composite materials after loading. Further research is still needed on the design, tunability, and photocatalytic activity of TiO2-based supported photocatalysts.
Constructing heterojunctions involves the combination of TiO2 with other semiconductors to achieve effective spatial separation and transfer of photogenerated carriers, thereby enhancing the overall photocatalytic activity of the catalyst. Park et al. [42] prepared In2S3/TiO2 composite materials. The formation of a heterojunction significantly enhanced the rapid separation of charges, resulting in the photocurrent density of this material being approximately 3.5 times that of pure TiO2. Yang et al. [43] constructed a heterojunction between TiO2 and Co3O4 and found that it had interface defects caused by a lattice mismatch. Through experiments and theoretical calculations, they discovered that the built-in electric field generated by electron transfer at the interface effectively promoted the migration of electron–hole pairs in opposite directions, thereby significantly enhancing the separation of photogenerated charges.
Among the aforementioned approaches, constructing heterojunctions with semiconductors, particularly 2D layered materials possessing tailored band alignments, proves highly effective for enhancing TiO2 photocatalysis. This strategy promotes charge separation while suppressing carrier recombination [44,45,46]. Layered double hydroxides (LDHs) are a class of two-dimensional lamellar crystalline compounds classified as anionic clays. Their structure features hexagonal or octahedral crystalline frameworks defining characteristic interlayer galleries [47]. The general chemical formula of layered double hydroxides (LDHs) is given by [M(1−x)2+Mx3+(OH)2]x+(An−)x/n·mH2O, where M2+ represents divalent cations (e.g., Ca2+, Mg2+, Fe2+, Co2+, Mn2+, Cu2+, Zn2+, Ni2+, etc.), M3+ denotes trivalent cations (e.g., Al3+, Cr3+, Fe3+, Mn3+, Co3+, etc.), An− signifies interlayer anions (e.g., CO32−, NO3, SO42−, Cl, etc.), and x represents the molar ratio of M2+ to M3+ [48,49]. LDHs feature expansive layered frameworks where bandgap engineering (2.0–3.4 eV) is achievable by varying M2+/M3+ cations. This tunability enables the effective construction of TiO2-based photocatalytic heterojunctions, enhancing visible-light absorption while offering abundant surface active sites. LDHs shorten migration pathways for photogenerated carriers, suppressing charge recombination. This collective effect markedly enhances overall photocatalytic efficiency [50,51,52,53].
We synthesized TiO2/LDHs composites with varied TiO2 loading via MgAl-LDHs layer reconstruction. The materials’ structure and morphology were characterized, while photocatalytic activity was assessed through methylene blue degradation under simulated sunlight. Reaction mechanisms were analyzed based on photocatalytic performance and radical trapping experiments.

2. Materials and Methods

2.1. Materials and Instruments

Butyl titanate (analytical grade, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), absolute ethanol (analytical grade, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), magnesium nitrate (analytical grade, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), and aluminum nitrate nonahydrate (analytical grade, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) were directly used for the preparation of composite materials. All aqueous solutions were prepared with deionized water.
The instruments used in this study are as follows: SX2-25-12 muffle furnace (Tianjin Zhonghuan Experimental Electric Furnace Co., Ltd., Tianjin, China), DS-20 vacuum drying oven (Tianjin Zhonghuan Experimental Electric Furnace Co., Ltd., Tianjin, China), SGY-1-type photochemical reactor (Nanjing Sidongke Electrical Appliance Co., Ltd., Nanjing, China), and CL19-1 Magnetic Stirrer (Shanghai SiLe Instrument Co., Ltd., Shanghai, China).

2.2. Synthesis of TiO2/LDHs

2.2.1. Preparation of TiO2 Sol

At room temperature, 4.4 mL of butyl titanate was dissolved in 26.4 mL of anhydrous ethanol (Beaker A) with stirring. Separately, Beaker B was prepared, containing 1.32 mL of triple-distilled water, 17.6 mL of anhydrous ethanol, and 0.62 mL of nitric acid. Using a peristaltic pump, Solution B was added to Beaker A at a rate of 30 drops/min with vigorous stirring. After complete addition, the mixture was continuously stirred for 30 min to yield a milky TiO2 sol.

2.2.2. Synthesis of LDHs Precursors

Under ambient temperature conditions, 12.80 g of magnesium nitrate hexahydrate (Mg(NO3)2 ·6H2O) and 9.38 g of aluminum nitrate nonahydrate (Al(NO3)3 ·9H2O) were weighed and dissolved in an appropriate amount of triple-distilled water in Beaker A. In this experiment, a mixed solution of NaOH and Na2CO3 was used to adjust the pH value of the system. According to the molar ratios of n(OH)/[n(Mg2+) + n(Al3+)] = 2.2 and n(CO32−)/[n(Mg2+) + n(Al3+)] = 0.667, 6.75 g of sodium hydroxide (NaOH) and 5.29 g of sodium carbonate (Na2CO3) were weighed and dissolved in Beaker B. The solutions in Beakers A and B were simultaneously added dropwise into a beaker containing triple-distilled water at a flow rate of 30 drops per minute. At this point, the pH of the reaction system was stably maintained at 8–9. Using a low-temperature saturated co-precipitation method in a 60 °C water bath, a hydrotalcite with a Mg:Al molar ratio of 2:1 was synthesized. The reaction mixture was continuously stirred during the process to obtain the desired LDHs precursor solution. Upon completion of the reaction, the resulting precipitate was collected through filtration, washed until a neutral pH was achieved [54,55].

2.2.3. Synthesis of TiO2/LDHs Nanocomposites

The LDHs precursor dispersion was stirred for 20 min before dropwise infusion into the TiO2 sol. The continuous stirring (12 h) of this mixture initiated crystallization, followed by a 6 h static aging stage. After the mixture was centrifuged and washed to a neutral pH, the obtained precipitate was oven-dried at 80 °C. Subsequent grinding and calcination (500 °C, 2 h, air) [56] of the precipitate yielded the TiO2/LDHs nanomaterial with Al:Ti = 1:1, labeled AT11. Identical procedures produced the following Al:Ti variants: 1:2 (AT12), 1:3 (AT13), 2:1 (AT21), and 3:1 (AT31).

2.3. Characterization

The phase and structural composition of the materials were analyzed through X-ray diffraction (XRD, D/MAX2500, Rigaku Corporation, Tokyo, Japan) and Raman spectroscopy (Horiba LabRAM HR Evolution, HORIBA Ltd., Kyoto, Japan), Shimadzu Corporation, Kyoto, Japan). The microstructural morphology of the samples was characterized using scanning electron microscopy (SEM, ZEISS Sigma 360, Carl Zeiss AG, Oberkochen, Germany) and transmission electron microscopy (TEM, JEOL JEM-F200, JEOL Ltd., Tokyo, Japan). The optical absorption properties and bandgap structure of the materials were evaluated via ultraviolet–visible diffuse reflectance spectroscopy (Shimadzu UV-3600i Plus, Shimadzu Corporation, Kyoto, Japan).

2.4. Photocatalytic Experiments

The photocatalytic experiments were carried out in an SGY-1-type photochemical reactor, using a 300 W xenon lamp (full spectrum, 320–2500 nm) to simulate a visible light source. The photocatalytic degradation of the target pollutant was conducted using a 15 mg/L methylene blue solution (150 mL), and the catalyst concentration was 1 g/L. Prior to initiating the reaction, the mixture was stirred at 200 rpm in the dark for 20 min to achieve an adsorption equilibrium between the catalyst and the methylene blue solution. Throughout the photocatalytic experiment, magnetic stirring was maintained (200 rpm), and samples were collected every 10 min. After centrifugation, the absorbance of the supernatant was analyzed. A spectrophotometer was employed to measure the absorbance of methylene blue solutions at various concentrations at 665 nm and the maximum absorption wavelength of methylene blue. The absorbance–concentration standard curve was plotted, and through linear fitting, the standard curve equation was obtained as follows:
y = 0.0716 x + 0.0542 ,   R 2 = 0.9992

3. Results and Discussion

3.1. Material Characterization and Analysis

3.1.1. X-Ray Powder Diffraction

To determine the crystal structure and characteristics of the composite materials, XRD analysis was conducted on TiO2, LDHs, and AT11. The detection was carried out using Cu Kα radiation. The diffracted beam was filtered using a Ni monochromator with a wavelength (λ) of 0.15418 nm. The accelerating voltage and current were 40 kV and 200 mA, respectively. The scanning range of the diffraction angle (2θ) was from 10° to 80° with a scan step size of 0.02°. The characterization results are shown in Figure 1. The pure TiO2 material exhibited diffraction peaks at 2θ values of 25.30°, 37.79°, 48.04°, 53.88°, 55.07°, 62.69°, 68.30°, 70.03°, and 74.62°, corresponding to the (101), (004), (200), (105), (211), (204), (116), (220), and (215) crystal planes of anatase TiO2 (ICSD #9852), respectively [57]. LDHs showed prominent diffraction peaks at 2θ values of 35.89°, 46.99°, and 62.13°, which align with the characteristic peaks of the standard card Mg0.67Al0.33(OH)2(CO3)0.165(H2O)0.48 (ICSD #6296), corresponding to the (012), (018) and (113) crystal planes of LDHs [57]. In the XRD pattern of the AT11 composite, the primary diffraction peaks observed for TiO2 were associated with the (101) and (200) planes, while those for LDHs were linked to the (012), (018), and (113) planes. Additionally, two new diffraction peaks appeared in the AT11 composite at 2θ values of 18.18°, 32.55°, and 49.19°, consistent with the characteristic peaks of MgTi2O5 (ICSD #37232), corresponding to its (200), (203), and (250) crystal planes. This indicates that the AT11 composite not only contains the TiO2 and LDHs phases but also forms a new crystalline phase structure of MgTi2O5. The crystallite sizes of the samples were calculated using the Scherrer formula. The particle size of TiO2(101) is 11.1 nm, the average particle size of LDHs(018) is 21.9 nm, and the average particle size of AT11(101) is 45.1 nm. The larger particle size observed for AT11 may be attributed to the formation of a composite structure, where TiO2 nanoparticles are embedded within the interlayers of LDHs during the synthesis process.

3.1.2. Morphological Analysis of TiO2/LDHs

The microstructure and morphology of the TiO2/LDHs composite were investigated using SEM and TEM, as illustrated in Figure 2. Figure 2a,b reveal the layered structure of LDHs and the lattice fringes of TiO2. The particle distribution is relatively uniform, although significant agglomeration is observed. Figure 2c demonstrates that the composite exhibits an irregular polygonal structure, with TiO2 nanoparticles dispersed within the LDHs lamellar structure [57]. This indicates that the layered architecture provides abundant loading sites for TiO2 particles [58]. Such structural modifications are hypothesized to enhance the photocatalytic activity of the composite, a supposition subsequently validated by the results of photocatalytic experiments.

3.1.3. Raman Spectroscopy

The Raman spectrum of the AT11 composite material is illustrated in Figure 3, where the absorption peak at 1065 cm−1 is attributed to the characteristic stretching vibration of CO32− within the interlayer of LDHs [19]. The absorption peak at 846 cm−1 corresponds to NO3, which is a residue obtained from the decomposition of the precursor. The peak at 777 cm−1 is assigned to the Ti-O-Ti vibration, indicating the successful loading of TiO2 on the surface of LDHs [59]. The peak at 655 cm−1 reflects the interfacial coupling vibration of Ti-O-M (M = Mg/Al), demonstrating the bonding interaction within the heterostructure. In conjunction with the characteristic peaks at 18.18° and 32.55° observed in the XRD characterization of AT11, this peak is identified as the interfacial bonding vibration of Ti-O-Mg [60]. The peaks at 487, 428, and 385 cm−1 represent the v3/v4 vibrational characteristics of the tetrahedral [MO4] groups, indicating the presence of a defect structure with coexisting tetrahedral and octahedral configurations in the composite material. This structure, induced by anion intercalation or synthesis conditions, can optimize the material’s photocatalytic, adsorption, and ion exchange properties [44,61]. The peak at 256 cm−1 is attributed to the vibration of Ti-O-Al or Ti-O-Mg. The peak at 161 cm−1 is identified as the B1g vibrational mode of anatase TiO2, reflecting the symmetric stretching vibration of the Ti-O bond in the TiO2 crystal. Although the typical B1g peak of anatase TiO2 is located near 144 cm−1, the shift to 161 cm−1 may be due to the specific microstructural features of the sample, such as tensile stress introduced by heterointerfaces or surface defects. In summary, these findings demonstrate that TiO2 has been successfully loaded onto the surface of LDHs, forming a heterostructure that enhances the material’s photocatalytic, adsorption, and ion exchange properties.

3.1.4. UV–Vis Diffuse Reflectance Spectroscopy

Ultraviolet–visible diffuse reflectance spectroscopy was performed on TiO2 and AT11 to investigate the optical properties of composite materials, with the results presented in Figure 4. The data reveal that TiO2 exclusively exhibits a UV response, with an absorption edge at 400 nm. In contrast, AT11 demonstrates an extended absorption edge reaching 500 nm. This indicates that the composite formation between TiO2 and LDHs establishes an efficient heterojunction structure through interfacial band alignment. The LDHs layer consists of tunable metal elements, typically possessing a conduction band position higher than that of TiO2 and a valence band position significantly lower than that of TiO2. This band arrangement forms a type II heterojunction at the interface, facilitating the migration of photogenerated electrons from the LDHs conduction band to the TiO2 conduction band, while simultaneously transferring holes from the TiO2 valence band to the LDHs valence band. This process not only effectively suppresses carrier recombination but also extends the photoresponse range of the composite material into the visible light region through the narrow bandgap characteristics of LDHs [62].

3.1.5. Photoluminescence Spectroscopy

The photoluminescence behavior formed by the recombination of electron–hole pairs can reflect the separation, migration, and transfer of photogenerated carriers in semiconductors. Generally, the separation efficiency of electron–hole pairs is measured by photoluminescence spectroscopy (PL). The lower the PL spectrum intensity, the higher the separation efficiency of photogenerated carriers, and the better the photocatalytic activity of the material [63]. The PL spectra were used to study the electron–hole pair separation efficiency of the materials to explain the superior photocatalytic activity of TiO2/LDHs composites, as shown in Figure 5. It is evident from Figure 5 that both TiO2 and TiO2/LDHs are effectively excited at 320 nm. After TiO2 was combined with LDHs, the luminescence intensity decreased significantly. This change is attributed to the capture of photogenerated electrons on TiO2 by the positive charges on the surface of LDHs. This indicates that the TiO2/LDHs composites can effectively suppress the recombination of photogenerated carriers, reduce the recombination rate of photo-induced electron–hole pairs, and thereby enhance the overall photocatalytic activity of the materials.

3.2. Photocatalytic Performance of Photocatalysts

3.2.1. Impact of Diverse Composite Materials on Photocatalytic Performance

This study investigated the photocatalytic degradation efficiency of TiO2/LDHs composite materials (AT11, AT12, AT13, AT21, and AT31) with varying content ratios under simulated sunlight, using methylene blue as the target pollutant, as illustrated in Figure 6. The experimental procedure involved a 20 min dark reaction to achieve an adsorption–desorption equilibrium of the catalyst prior to photocatalytic testing. The results demonstrate that the pure TiO2 photocatalyst exhibits minimal adsorption capacity for methylene blue, while the composite photocatalysts show progressively enhanced adsorption with an increase in the LDHs content. However, the adsorption rate significantly decelerates when the Al:Ti molar ratio exceeds 1:1. Photocatalytic reaction analysis reveals that all TiO2/LDHs composites exhibit substantially higher photocatalytic activity compared with pure TiO2. The photocatalytic degradation efficiency initially increases and subsequently decreases with a rise in the LDHs content, reaching optimal performance at an Al:Ti molar ratio of 1:1 (AT11), thus achieving a remarkable 98.20% degradation rate of methylene blue within 70 min of reaction.
Based on the aforementioned results, the TiO2/LDHs composite material exhibits superior adsorption capabilities, which enhance the direct contact efficiency between the target degradation molecules, their photocatalytic oxidation intermediates, and the photogenerated holes on the composite surface. This synergistic effect of adsorption and photocatalysis significantly improves the overall photocatalytic efficiency. Furthermore, during the material compounding process, TiO2 is embedded into the LDHs layers and pores, ensuring its uniform dispersion on the LDHs carrier. This provides more active sites for photocatalytic reactions, reduces the agglomeration and shielding of photocatalytically active particles, and enhances the overall photocatalytic reaction efficiency. When the proportion of LDHs in the composite material is relatively low, TiO2 can be uniformly distributed on the surface and interlayer of LDHs. However, with the continuous increase in the LDHs content, the excessive LDHs lead to a relative reduction in active sites on the surface of the composite material, resulting in a decrease in its photocatalytic efficiency.

3.2.2. Kinetic Analysis of Photocatalytic Reactions

The photocatalytic degradation data were fitted using a pseudo-first-order kinetic model, with the results presented in Table 1 and Figure 7. The fitting results indicate that the linear correlation coefficients for all catalysts exceed 0.96, demonstrating that their photocatalytic degradation processes conform to the pseudo-first-order reaction kinetic model. Among them, AT11 exhibits the highest photocatalytic reaction rate constant of 0.0543 min−1, which is 7.5 times that of pure TiO2, thereby substantiating a significant enhancement in photocatalytic activity upon the combination of TiO2 with LDHs.
The comparison and analysis results of the catalytic degradation performance of the same type of photocatalytic materials on pollutants are shown in Table 2, where TiO2/LDHs not only show a high degradation rate for methylene blue but also a short period, demonstrating a promising application prospect for the photocatalytic degradation of pollutants.

3.2.3. Impact of Catalyst Concentration on Photocatalytic Performance

The experiment employed AT11 composite material as the catalyst, maintaining identical reaction conditions and procedures as the aforementioned photocatalytic process. The photocatalytic reaction was conducted for 70 min to investigate the impact of varying catalyst concentrations on the degradation of methylene blue. The results are illustrated in Figure 8.
The figure clearly illustrates that when the catalyst concentration is below 1 g/L, the degradation rate of methylene blue exhibits an increasing trend with the elevation of catalyst concentration. The augmentation of catalyst concentration facilitates the adsorption of more reactants on the catalyst surface, thereby enhancing light absorption and improving the generation rate of photogenerated holes, which consequently promotes the photocatalytic reaction. The maximum degradation rate of methylene blue is achieved at a catalyst concentration of 1 g/L. However, when the catalyst concentration exceeds 1 g/L, the degradation rate of methylene blue gradually decreases with the increase in catalyst dosage. This phenomenon can be attributed to the shielding effect of the excessive catalyst on incident light, which adversely affects the transmittance of the reaction solution and consequently reduces the generation rate of photogenerated holes. Therefore, under the experimental conditions of this study, the optimal catalyst concentration is determined to be 1 g/L.

3.2.4. Impact of Inorganic Anions on Photocatalytic Reactions

This study selected Cl, NO3, SO42−, and CO32− as representative inorganic anions to investigate their impact on the photocatalytic degradation of methylene blue. Using AT11 as the catalyst at a concentration of 1 g/L, the experiments were conducted with the addition of 5 mmol/L of NaCl, NaNO3, Na2SO4, and Na2CO3, respectively. The photocatalytic reactions were carried out for 70 min, and the results are presented in Figure 9, where Cl, NO3, and SO42− exhibit minimal impact on the degradation of methylene blue within the system, whereas CO32− demonstrates a significant inhibitory effect on the photodegradation of methylene blue. This phenomenon can be attributed to the weak acidity of CO32−, which, upon hydrolysis in an aqueous solution, readily combines with hydrogen ions to form HCO3 and OH. Subsequently, HCO3 reacts with active species such as ·OH generated during the photocatalytic process, resulting in the formation of carbonate compounds. These compounds further cover the active sites, thereby reducing the catalytic activity.

3.2.5. Analysis of Photocatalytic Degradation Mechanism

The primary active species in photocatalytic oxidation reactions include photogenerated holes (h+), hydroxyl radicals (·OH), and superoxide radicals (·O2−). To explore the photocatalytic oxidation mechanism of methylene blue by TiO2/LDHs composites, this study conducted a radical trapping experiment. By introducing selective radical inhibitors (quenchers), the activity of specific radicals was blocked, thereby analyzing the contribution rate of different radicals to the degradation of the target pollutant. We individually added sodium bicarbonate [69] (NaHCO3, h+ scavenger), isopropanol [70] (IPA, ·OH scavenger), and p-benzoquinone [71] (BQ, ·O2− scavenger) to the reaction system, and conducted photocatalytic experiments using AT11 as the catalyst under the same reaction conditions. The results are shown in Figure 10.
As illustrated in Figure 10, the degradation efficiency of AT11 towards methylene blue decreased to 54.52%, 68.10%, and 87.63% upon the addition of NaHCO3, IPA, and BQ, respectively. This indicates that all three reactive species play significant roles in the photocatalytic degradation of methylene blue. Specifically, h+ and ·OH are identified as the primary reactive species responsible for the degradation, while ·O2− exhibits a limited influence on the photocatalytic reaction process.
Based on the above analysis, the reaction mechanism of the photocatalytic oxidation of methylene blue by TiO2/LDHs composite materials is speculated to be as follows. The LDHs component in the composite material exerts its adsorption performance, adsorbing organic pollutants on the adsorption sites, achieving the enrichment of organic pollutants and significantly increasing the adsorption capacity of organic pollutants on the catalyst surface. Under simulated sunlight irradiation, the TiO2/LDHs composite materials absorb light energy and become excited. The valence band electrons of O’s 2p orbitals transition to the hybridized Ti 3d orbitals and Al 3p orbitals in the conduction band, forming photogenerated electrons (e), while photogenerated holes (h+) are formed in the valence band, generating electron–hole pairs. The type II heterojunction formed at the interface between TiO2 and LDHs drives the photogenerated electrons to migrate from the LDHs conduction band to the TiO2 conduction band, while the holes transfer from the TiO2 valence band to the LDHs valence band. This process promotes the transfer of charges, and the existence of intermediate energy levels in indirect bandgaps prolongs the recombination time of electron–hole pairs, resulting in a decrease in the recombination rate of photogenerated electrons and holes, and thereby enhancing the photocatalytic activity of the composite material. In the TiO2/LDHs composite material photocatalytic degradation of the methylene blue system, the main active groups involved in the reaction are photogenerated holes (h+) and hydroxyl radicals (·OH). h+ has strong oxidizing properties and can directly undergo redox reactions with methylene blue molecules. ·OH is generated by the combination of h+ with H2O and OH, and it cooperates in oxidizing and degrading methylene blue.

3.2.6. Evaluation of Photocatalytic Stability in Composite Materials

The reusability of photocatalytic materials constitutes a critical factor in assessing the stability of catalyst performance. This study conducted four consecutive cycling experiments on AT11 to evaluate its photocatalytic degradation efficiency of methylene blue, with the results illustrated in Figure 11. The data demonstrate that, after four cycles of reuse, the degradation efficiency of AT11 for methylene blue decreased from 98.20% to 78.93%, indicating relatively favorable stability of the catalyst. The observed decline in activity may be attributed to two primary factors: minor sample loss during the recovery and washing processes, and the partial coverage of catalyst surface pores by reactants and products during the reaction, leading to the deactivation of active sites and subsequent reduction in degradation efficiency.

4. Conclusions

TiO2/LDHs composite materials with varying Al/Ti molar ratios were synthesized via the co-precipitation method. The formation of an efficient heterojunction structure through interfacial band alignment effectively suppressed carrier recombination and extended the photoresponse range of the composite materials into the visible light spectrum. The AT11 composite demonstrated optimal performance, achieving a 98.20% degradation rate of methylene blue (15 mg/L) under simulated sunlight irradiation for 70 min. The photocatalytic reaction followed first-order kinetics with a rate constant of 0.0543 min−1. After four consecutive cycles, the AT11 composite maintained a methylene blue degradation efficiency of 78.93%, exhibiting excellent stability.
The composite of TiO2 with LDHs demonstrates superior photocatalytic performance due to the synergistic effects of different catalyst components and modifications in the band structure. LDHs significantly enhance the adsorption capacity of the composite while mitigating the agglomeration of TiO2, thereby providing more effective active sites for the reaction and substantially improving the efficiency of photocatalytic degradation. Radical trapping experiments reveal that the primary active species responsible for the photocatalytic degradation of methylene blue by TiO2/LDHs are h+ and ·OH.

Author Contributions

Conceptualization, J.W. and L.R.; methodology, J.W. and L.R.; formal analysis, J.W.; data curation, J.W. and L.R.; writing—original draft preparation, J.W.; writing—review and editing, L.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by “Key Research and Development Program Project of Shandong Province”, grant number 2024TSGC0884.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors acknowledge the financial support through the Key Research and Development Program Project of Shandong Province, number (2024TSGC0884), Shandong, China.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Guo, J.; Zhou, T.T.; Guo, H.; Ge, C.; Lu, J.J. Application of nano-TiO2@adsorbent composites in the treatment of dye wastewater: A review. J. Eng. Fibers Fabr. 2025, 20, 15589250251329450. [Google Scholar] [CrossRef]
  2. Chang, Q.; Jiang, G.D.; Hu, M.X.; Huang, J.; Tang, H.Q. Adsorption of Methylene Blue from Aqueous Solution onto Magnetic Fe3O4/Graphene Oxide Nanoparticles. Environ. Sci. 2014, 35, 1804–1809. (In Chinese) [Google Scholar] [CrossRef]
  3. Willison, E.O.C.; Lopes, A.S.C.; Monteiro, W.R.; Filho, G.N.R.; Nobre, F.X.; Luz, P.T.S.; Nascimento, L.A.S.; Costa, C.E.F.; Monteiro, W.F.; Vieira, M.O.; et al. Layered double hydroxides as heterostructure LDH@Bi2WO6 oriented toward visible-light-driven applications: Synthesis, characterization, and its photocatalytic properties. React. Kinet. Mech. Catal. 2020, 131, 505–524. [Google Scholar] [CrossRef]
  4. Fadzli, J.; Hamid, K.H.K.; Him, N.R.N.; Puasa, S.W. A critical review on the treatment of reactive dye wastewater. Desalin. Water Treat. 2022, 257, 185–203. [Google Scholar] [CrossRef]
  5. Li, Z.J.; Sun, Y.K.; Xing, J.; Meng, A. Fast removal of Methylene Blue by Fe3O4 magnetic nanoparticles and their cycling property. J. Nanosci. Nanotechnol. 2019, 19, 2116–2123. [Google Scholar] [CrossRef]
  6. Kang, J.Y.; Deng, S.; Han, X.; Wu, S.X.; Bai, Y.Y.; Jiang, Y.H.; Yang, Y. Adsorption Properties and Mechanism of Ciprofloxacin Enhanced by Mn-Doped Biochar. Res. Environ. Sci. 2024, 37, 2526–2536. (In Chinese) [Google Scholar] [CrossRef]
  7. Kim, Y.K.; Yoo, K.; Kim, M.S.; Han, I.; Lee, M.; Kang, B.R.; Lee, T.K.; Park, J. The capacity of wastewater treatment plants drives bacterial community structure and its assembly. Sci. Rep. 2019, 9, 14809. [Google Scholar] [CrossRef]
  8. Assress, H.A.; Selvarajan, R.; Nyoni, H.; Ntushelo, K.; Mamba, B.B.; Msagati, T.A.M. Diversity, co-occurrence and implications of fungal communities in wastewater treatment plants. Sci. Rep. 2019, 9, 14056. [Google Scholar] [CrossRef]
  9. Das, S.; Ghangrekar, M.M. Tungsten oxide as electrocatalyst for improved power generation and wastewater treatment in microbial fuel cell. Environ. Technol. 2020, 41, 2546–2553. [Google Scholar] [CrossRef] [PubMed]
  10. Nobre, F.X.; Pessoa, W.A.; Ruiz, Y.L.; Bentes, V.L.I.; Silva-Moraes, M.O.; Silva, T.M.C.; Rocco, M.L.M.; Larrudé, D.R.G.; de Matos, J.M.E.; Couceiro, P.R.D. Facile synthesis of nTiO2 phase mixture: Characterization and catalytic performance. Mater. Res. Bull. 2019, 109, 60–71. [Google Scholar] [CrossRef]
  11. Bodzek, M.; Konieczny, K.; Kwiecinska-Mydlak, A. Nano-photocatalysis in water and wastewater treatment. Desalin. Water Treat. 2021, 243, 51–74. [Google Scholar] [CrossRef]
  12. Kumar, S.; Chanana, A. TiO2 based nanomaterial: Synthesis, structure, photocatalytic properties, and removal of dyes from wastewater. Korean J. Chem. Eng. 2023, 40, 1822–1838. [Google Scholar] [CrossRef]
  13. Liang, S.X.; Zhang, Y.H.; Sun, Z.C.; Chang, Y.K. Laboratory study on the evolution of waves parameters due to wave breaking in deep water. Wave Motion 2017, 68, 31–42. [Google Scholar] [CrossRef]
  14. Njema, G.G.; Kibet, J.K. A review of novel materials for nano-photocatalytic and optoelectronic applications: Recent perspectives, water splitting and environmental remediation. Prog. Eng. Sci. 2024, 1, 100018. [Google Scholar] [CrossRef]
  15. Chen, J.R.; Qiu, F.X.; Xu, W.Z.; Cao, S.S.; Zhu, H.J. Recent progress in enhancing photocatalytic efficiency of TiO2-based materials. Appl. Catal. A-Gen. 2015, 495, 131–140. [Google Scholar] [CrossRef]
  16. Chen, S.Q.; Hu, Y.H. Color TiO2 Materials as Emerging Catalysts for Visible-NIR Light Photocatalysis, A Review. Catal. Rev.-Sci. Eng. 2023, 66, 1951–1991. [Google Scholar] [CrossRef]
  17. Guo, Q.; Zhou, C.Y.; Ma, Z.B.; Yang, X.M. Fundamentals of TiO2 Photocatalysis: Concepts, Mechanisms, and Challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef] [PubMed]
  18. Peiris, S.; de Silva, H.B.; Ranasinghe, K.N.; Bandara, S.V.; Perera, I.R. Recent development and future prospects of TiO2 photocatalysis. J. Chin. Chem. Soc. 2021, 68, 738–769. [Google Scholar] [CrossRef]
  19. Fernandes, E.; Gomes, J.; Martins, R.C. Semiconductors Application Forms and Doping Benefits to Wastewater Treatment: A Comparison of TiO2, WO3, and g-C3N4. Catalysts 2022, 12, 1218. [Google Scholar] [CrossRef]
  20. Sari, Y.; Gareso, P.L.; Tahir, D. Review: Influence of synthesis methods and performance of rare earth doped TiO2 photocatalysts in degrading dye effluents. Int. J. Environ. Sci. Technol. 2025, 22, 1975–1994. [Google Scholar] [CrossRef]
  21. Indira, A.C.; Muthaian, J.R.; Pandi, M.; Mohammad, F.; Al-Lohedan, H.A.; Soleiman, A.A. Photocatalytic Efficacy and Degradation Kinetics of Chitosan-Loaded Ce-TiO2 Nanocomposite towards for Rhodamine B Dye. Catalysts 2023, 13, 1506. [Google Scholar] [CrossRef]
  22. Tiwari, V.; Pal, B.; Kaur, S. Photocatalysis of Ag-loaded MgTiO3 for degradation of fuchsin dye and dyes present in textile wastewater under sunlight. Sol. Energy 2025, 296, 113587. [Google Scholar] [CrossRef]
  23. Li, Y.; Zhang, M.Q.; Liu, Y.F.; Sun, Y.X.; Zhao, Q.H.; Chen, T.L.; Chen, Y.F.; Wang, S.F. In Situ Construction of Bronze/Anatase TiO2 Homogeneous Heterojunctions and Their Photocatalytic Performances. Nanomaterials 2022, 12, 1122. [Google Scholar] [CrossRef]
  24. Fang, Q.; Sun, Q.; Zhong, R.; Wang, H.; Qi, J. Recent advances in doping engineering of heterogeneous catalyst for carbon dioxide hydrogenation. Mater. Today Chem. 2025, 46, 102770. [Google Scholar] [CrossRef]
  25. Khan, H.; Shah, M.U.H. Modification strategies of TiO2 based photocatalysts for enhanced visible light activity and energy storage ability: A review. J. Environ. Chem. Eng. 2023, 11, 11153. [Google Scholar] [CrossRef]
  26. Asahi, R.; Morikawa, T.; Irie, H. Nitrogen-doped titanium dioxide as visible-light sensitive photocatalyst: Designs, developments, and prospects. Chem. Rev. 2014, 114, 9824–9852. [Google Scholar] [CrossRef]
  27. Zhao, Y.; Yu, Y.; Li, J.; Xie, J.; He, C. Hydrodechlorination under O2 promotes catalytic oxidation of CVOCs over a Pt/TiO2 catalyst at low temperature. Chem. Commun. 2025, 61, 4710–4713. [Google Scholar] [CrossRef]
  28. Park, J.; Lam, S.S.; Park, Y.; Kim, B.J.; An, K.; Jung, S. Fabrication of Ni/TiO2 visible light responsive photocatalyst for decomposition of oxytetracycline. Environ. Res. 2023, 216, 114657. [Google Scholar] [CrossRef]
  29. Ramasubbu, V.; Kumar, P.R.; Chellapandi, T.; Madhumitha, G.; Mothi, E.M.; Shajan, X.S. Zn(II) porphyrin sensitized (TiO2@Cd-MOF) nanocomposite aerogel as novel photocatalyst for the effective degradation of methyl orange (MO) dye. Opt. Mater. 2022, 132, 112558. [Google Scholar] [CrossRef]
  30. Chen, A.; Chen, W.F.; Majidi, T.; Pudadera, B.; Koshy, P. Mo-doped, Cr-Doped, and Mo-Cr codoped TiO2 thin-film photocatalysts by comparative sol-gel spin coating and ion implantation. Int. J. Hydrogen Energy 2021, 46, 12961–12980. [Google Scholar] [CrossRef]
  31. Tasbihi, M.; Kocí, K.; Troppová, I.; Edelmannová, M.; Reli, M.; Capek, L.; Schomäcker, R. Photocatalytic reduction of carbon dioxide over Cu/TiO2 photocatalysts. Environ. Sci. Pollut. Res. 2018, 25, 34903–34911. [Google Scholar] [CrossRef]
  32. Zhu, X.F.; Cheng, B.; Yu, J.G.; Ho, W.K. Halogen poisoning effect of Pt-TiO2 for formaldehyde catalytic oxidation performance at room temperature. Appl. Surf. Sci. 2016, 364, 808–814. [Google Scholar] [CrossRef]
  33. Lim, C.; An, H.R.; Ha, S. Highly visible-light-responsive nanoporous nitrogen-doped TiO2 (N-TiO2) photocatalysts produced by underwater plasma technology for environmental and biomedical applications. Appl. Surf. Sci. 2023, 638, 158123. [Google Scholar] [CrossRef]
  34. Shuaib, S.S.A.; Lu, Y.; Wang, Q. Ti3C2 MXene-Derived TiO2/g-C3N4 Heterojunctions for Highly Efficient Photocatalytic H 2 Generation. Chem.-Catal. Chem. Eng. 2025, 7, 17. [Google Scholar] [CrossRef]
  35. Saqib, M.; Rahman, N.; Safeen, K. Structure phase-induced photodegradation properties of cobalt-sulfur co-doped TiO2 nanoparticles synthesized by hydrothermal route. J. Mater. Res. Technol. 2023, 26, 8048–8060. [Google Scholar] [CrossRef]
  36. Chen, Y.; Li, A.; Fu, X.; Peng, Z. Novel F-doped carbon nanotube@(N,F)-co-doped TiO2-δ nanocomposite: Highly active visible-light-driven photocatalysts for water decontamination. Appl. Surf. Sci. 2023, 609, 155460. [Google Scholar] [CrossRef]
  37. Pan, X.; Zhao, Y.; Liu, S.; Korzeniewski, C.L.; Wang, S.; Fan, Z. Comparing Graphene-TiO2 Nanowire and Graphene-TiO2 Nanoparticle Composite Photocatalysts. ACS Appl. Mater. Interfaces 2012, 4, 3944–3950. [Google Scholar] [CrossRef]
  38. Yu, D.; Bai, J.; Liang, H.; Li, C. Electrospinning, solvothermal, and self-assembly synthesis of recyclable and renewable AgBrTiO2/CNFs with excellent visible-light responsive photocatalysis. J. Alloys Compd. 2016, 683, 329–338. [Google Scholar] [CrossRef]
  39. Sibi, M.G.; Verma, D.; Kim, J. Magnetic core-shell nanocatalysts: Promising versatile catalysts for organic and photocatalytic reactions. Catal. Rev. Sci. Eng. 2020, 62, 163–311. [Google Scholar] [CrossRef]
  40. Attia, M.S.; Abdel-Wahed, M.S.; El-Kalliny, A.S.; Badawy, M.I.; Gad-Allah, T.A. Core Double-Shell MnFe2O4@rGO@TiO2 Superparamagnetic Photocatalyst for Wastewater Treatment under Solar Light. Chem. Eng. J. 2020, 382, 122936. [Google Scholar] [CrossRef]
  41. Shu, J.; Yao, Y.; Zhang, X.; Zhang, L.; Zhao, H.; Zhang, S. Template-free synthesis of core-shell Fe3O4@MoS2@mesoporous TiO2 magnetic photocatalyst for wastewater treatment. Int. J. Miner. Metall. Mater. 2023, 30, 177–191. [Google Scholar] [CrossRef]
  42. Park, J.; Lee, T.H.; Kim, C.; Lee, S.A.; Choi, M.J.; Kim, H.; Yang, J.W.; Lim, J.; Jang, H.W. Hydrothermally Obtained Type-Ⅱ Heterojunction Nanostructures of In2S3/TiO2 for Remarkably Enhanced Photoelectrochemical Water Splitting. Appl. Catal. B Environ. 2021, 295, 120276. [Google Scholar] [CrossRef]
  43. Yang, Y.; Zhao, S.; Bi, F.; Chen, J.; Wang, Y.; Cui, L.; Xu, J.; Zhang, X. Highly efficient photothermal catalysis of toluene over Co3O4/TiO2 p-n heterojunction: The crucial roles of interface defects and band structure. Appl. Catal. B Environ. 2022, 315, 121550. [Google Scholar] [CrossRef]
  44. Xu, S.W.; Li, Z.W.; Wen, J.H.; Qiu, P.; Xie, A.J.; Peng, H.P. Review of TiO2-Based Heterojunction Coatings in Photocathodic Protection. Acs Appl. Nano Mater. 2024, 7, 8464–8488. [Google Scholar] [CrossRef]
  45. Du, M.Q.; Cao, S.X.; Ye, X.Z.; Ye, J.F. Recent Advances in the Fabrication of All-Solid-State Nanostructured TiO2-Based Z-Scheme Heterojunctions for Environmental Remediation. J. Nanosci. Nanotechnol. 2020, 20, 5861–5873. [Google Scholar] [CrossRef]
  46. Wu, X.H.; Chen, G.Q.; Wang, J.; Li, J.M.; Wang, G.H. Review on S-Scheme Heterojunctions for Photocatalytic Hydrogen Evolution. Acta Phys.-Chim. Sin. 2023, 39, 2212016. [Google Scholar] [CrossRef]
  47. Ye, D.; Sun, L.; Feng, J.Y.; Gao, S.J.; Zhu, K.; Wu, K.; Guo, R.T. Progress on photocatalytic elimination of CO2 and gaseous pollutants over LDHs-based materials. J. Ind. Eng. Chem. 2025, 144, 175–191. [Google Scholar] [CrossRef]
  48. Wang, Y.Z.; Wang, S.; Li, X.L.; Bai, P.; Yan, W.F.; Yu, J.H. Layered Inorganic Cationic Frameworks beyond Layered Double Hydroxides (LDHs): Structures and Applications. Eur. J. Inorg. Chem. 2020, 43, 4055–4063. [Google Scholar] [CrossRef]
  49. Chang, Q.F.; Zhang, X.L.; Wang, B.; Niu, J.T.; Yang, Z.X.; Wang, W.C. Fundamental understanding of electrocatalysis over layered double hydroxides from the aspects of crystal and electronic structures. Nanoscale 2022, 14, 1107–1122. [Google Scholar] [CrossRef]
  50. Hadnadjev-Kostic, M.; Vulic, T.; Lukic, N.; Jokic, A.; Karanovic, D. Photocatalytic Performance of TiO2-ZnAl LDH Based Materials: Kinetics and Neural Networks Approach. Pol. J. Environ. Stud. 2022, 31, 4117–4125. [Google Scholar] [CrossRef]
  51. Mohapatra, L.; Parida, K. A Review on Recent Progress, Challenges and Perspective of Layered Double Hydroxides as Promising Photocatalysts. J. Mater. Chem. A 2016, 4, 10744–10766. [Google Scholar] [CrossRef]
  52. Wang, L.Y.; Gao, X.; Cheng, Y.Q.; Zhang, X.X.; Wang, G.Q.; Zhang, Q.Y.; Su, J.X. TiO2@MgAl-layered double hydroxide with enhanced photocatalytic activity towards degradation of gaseous toluene. J. Photochem. Photobiol. A-Chem. 2019, 369, 44–53. [Google Scholar] [CrossRef]
  53. Suh, M.J.; Shen, Y.; Chan, C.K.; Kim, J.H. Titanium Dioxide-Layered Double Hydroxide Composite Material for Adsorption-Photocatalysis of Water Pollutants. Langmuir 2019, 35, 8699–8708. [Google Scholar] [CrossRef]
  54. Yang, K.; Yan, L.G.; Yang, Y.M.; Yu, S.; Shan, R.; Yu, H.Q.; Zhu, B.; Du, B. Adsorptive removal of phosphate by Mg–Al and Zn–Al layered double hydroxides: Kinetics, isotherms and mechanisms. Sep. Purif. Technol. 2014, 124, 36–42. [Google Scholar] [CrossRef]
  55. Yan, L.G.; Yang, K.; Shan, R.; Yan, T.; Wei, J.; Yu, S.; Yu, H.; Du, B. Kinetic, isotherm and thermodynamic investigations of phosphate adsorption onto core–shell Fe3O4@LDHs composites with easy magnetic separation assistance. J. Colloid Interface Sci. 2015, 448, 508–516. [Google Scholar] [CrossRef] [PubMed]
  56. Wei, J. Photoreduction of CO2 Using TiO2-Based Nano-Particles; Tianjin University: Tianjin, China, 2011. (In Chinese) [Google Scholar]
  57. Gao, X.; Zheng, K.; Zhang, Q.; Cao, X.; Wu, S.; Su, J. Self-assembly TiO2-RGO/LDHs nanocomposite: Photocatalysis of VOCs degradation in simulation air. Appl. Surf. Sci. 2022, 586, 152882. [Google Scholar] [CrossRef]
  58. Wang, Z.; Sun, Y.; Zhao, X.; Zhang, K.; Wang, X.; Liu, C.; Cui, J. Construction of two-dimensional nano-composite g-C3N4/LDHs and photocatalytic degradation of ciprofloxacin. Appl. Chem. Ind. 2025, 54, 881–886, 894. [Google Scholar] [CrossRef]
  59. Seftel, E.M.; Niarchos, M.; Mitropoulos, C.; Mertens, M.; Vansant, E.F.; Cool, P. Photocatalytic removal of phenol and methylene-blue in aqueous media using TiO2@LDH clay nanocomposites. Catal. Today 2015, 252, 120–127. [Google Scholar] [CrossRef]
  60. Singh, R.; Biswas, K. Synthesis optimization and investigation on electrical properties of Fe2+-doped Mg2TiO4 ceramics for energy storage. J. Mater. Sci. Mater. Electron. 2020, 31, 12434–12443. [Google Scholar] [CrossRef]
  61. Khan, M.A.; Fattah-alhosseini, A.; Kaseem, M. Recent advances in the design and surface modification of titanium-based LDH for photocatalytic applications. Inorg. Chem. Commun. 2023, 153, 110739. [Google Scholar] [CrossRef]
  62. Miljevic, B.; van der Bergh, J.M.; Vucetic, S.; Lazar, D.; Ranogajec, J. Molybdenum doped TiO2 nanocomposite coatings Visible light driven photocatalytic self-cleaning of mineral substrates. Ceram. Int. 2017, 43, 8214–8221. [Google Scholar] [CrossRef]
  63. Jia, W.; Wu, H.; Zheng, Y.; Liu, Z.; Cai, G.; Wen, J.; Hu, G.; Tang, T.; Li, X.; Jiang, L.; et al. Co/N Co-Doped MoS2 with high pseudocapacitive performance for solid-stateflexible supercapacitors. ACS Appl. Energy Mater. 2023, 6, 2570–2581. [Google Scholar] [CrossRef]
  64. Salehi, G.; Abazari, R.; Mahjoub, A.R. Visible-light-induced graphitic-C3N4@nickel-aluminum layered double hydroxide nanocomposites with enhanced photocatalyticactivity for removal of dyes in water. Inorg. Chem. 2018, 57, 8681–8691. [Google Scholar] [CrossRef]
  65. Ni, Z.M.; Xue, J.L. Synthesis of CuMgAl Layered Double Hydroxides for Efficient Photocatalysis of Rhodamine B. Chem. J. Chin. Univ. 2013, 34, 503–508. [Google Scholar] [CrossRef]
  66. Li, J.; Liu, Y.; Li, H.; Chao, C. Fabrication of g-C3N4/TiO2 composite photocatalyst with extended absorption wavelength range and enhanced photocatalytic performance. J. Photochem. Photobiol. A 2016, 317, 151–160. [Google Scholar] [CrossRef]
  67. Wei, J.; Gong, F.; Wang, S. Study on the Photocatalytic Degradation of Ibuprofen in Water by Load Modified TiO2. Technol. Water Treat. 2025, 51, 68–76. [Google Scholar] [CrossRef]
  68. Li, K.; Ding, Q.; Qin, L.; Li, K.; Ma, J. Preparation and photocatalytic performance of TiO2-SiO2 composites. Funct. Mater. 2024, 10, 10220–10236. [Google Scholar] [CrossRef]
  69. Hassani, A.; Eghbali, P.; Mahdipour, F.; Waclawek, S.; Lin, K.Y.A.; Ghanbari, F. Insights into the synergistic role of photocatalytic activation of peroxymonosulfate by UVA-LED irradiation over CoFe2O4-rGO nanocomposite towards effective Bisphenol A degradation: Performance, mineralization, and activation mechanism. Chem. Eng. J. 2022, 453, 139556. [Google Scholar] [CrossRef]
  70. Xie, W.J.; Zhang, Y.; Xu, L.; Xie, D.; Jiang, L.; Dong, Y.M.; Yuan, Y. Degradation of Organic Dyes by the UCNP/h-BN/TiO2 Ternary Photocatalyst. Acs Omega 2023, 8, 48662–48672. [Google Scholar] [CrossRef]
  71. Rao, L.J.; Yang, Y.F.; Chen, L.K.; Liu, X.D.; Chen, H.X.; Yao, Y.Y.; Wang, W.T. Highly efficient removal of organic pollutants via a green catalytic oxidation system based on sodium metaborate and peroxymonosulfate. Chemosphere 2020, 238, 124687. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of TiO2, LDHs, and AT11.
Figure 1. XRD patterns of TiO2, LDHs, and AT11.
Water 17 02582 g001
Figure 2. Microscopic morphology of materials: (a,b) AT11 (TEM) and (c) AT11 (SEM).
Figure 2. Microscopic morphology of materials: (a,b) AT11 (TEM) and (c) AT11 (SEM).
Water 17 02582 g002
Figure 3. Raman spectroscopy spectrum of AT11.
Figure 3. Raman spectroscopy spectrum of AT11.
Water 17 02582 g003
Figure 4. UV–Vis spectra of materials: (a) TiO2 and (b) AT11.
Figure 4. UV–Vis spectra of materials: (a) TiO2 and (b) AT11.
Water 17 02582 g004
Figure 5. Photoluminescence spectra of materials.
Figure 5. Photoluminescence spectra of materials.
Water 17 02582 g005
Figure 6. Degradation effect of different photocatalytic materials on methylene blue: (a) Comparison of degradation effects of TiO2/LDHs with different Al:Ti molar ratios; (b) Comparison of degradation effects of TiO2 and LDHs.
Figure 6. Degradation effect of different photocatalytic materials on methylene blue: (a) Comparison of degradation effects of TiO2/LDHs with different Al:Ti molar ratios; (b) Comparison of degradation effects of TiO2 and LDHs.
Water 17 02582 g006
Figure 7. Pseudo-first-order kinetic fitting curve for photocatalytic degradation of methylene blue.
Figure 7. Pseudo-first-order kinetic fitting curve for photocatalytic degradation of methylene blue.
Water 17 02582 g007
Figure 8. Impact of catalyst concentration on methylene blue degradation.
Figure 8. Impact of catalyst concentration on methylene blue degradation.
Water 17 02582 g008
Figure 9. Effect of coexisting anions on degradation of methylene blue by AT11.
Figure 9. Effect of coexisting anions on degradation of methylene blue by AT11.
Water 17 02582 g009
Figure 10. Free-radical-trapping experiment.
Figure 10. Free-radical-trapping experiment.
Water 17 02582 g010
Figure 11. Recycling of AT11 composites.
Figure 11. Recycling of AT11 composites.
Water 17 02582 g011
Table 1. Kinetic parameters of photocatalytic degradation of methylene blue.
Table 1. Kinetic parameters of photocatalytic degradation of methylene blue.
SampleK/min−1R2
TiO20.007 20.968 7
AT130.009 10.976 6
AT120.010 20.987 5
AT110.054 30.980 5
AT210.016 80.965 4
AT310.027 00.969 6
Table 2. Comparison of photocatalytic properties between similar materials.
Table 2. Comparison of photocatalytic properties between similar materials.
CatalystPollutantLight SourceC0 (mg/L)Time (min)Efficiency (%)Ref.
g-C3N4/NiAl-LDHRhBVisible2024093.00[64]
CuMgAl-LDHsRhBVisible1024085.20[65]
g-C3N4/TiO2Methyl orangeVisible518073.55[66]
Fe-S/TiO2/GFCIbuprofenVisible1012080.04[67]
TiO2-SiO2RhBVisible5018097.80[68]
TiO2/LDHsRhBVisible157098.20This research
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

Wei, J.; Ren, L. Synthesis and Characterization of Visible-Light-Responsive TiO2/LDHs Heterostructures for Enhanced Photocatalytic Degradation Performance. Water 2025, 17, 2582. https://doi.org/10.3390/w17172582

AMA Style

Wei J, Ren L. Synthesis and Characterization of Visible-Light-Responsive TiO2/LDHs Heterostructures for Enhanced Photocatalytic Degradation Performance. Water. 2025; 17(17):2582. https://doi.org/10.3390/w17172582

Chicago/Turabian Style

Wei, Jing, and Liying Ren. 2025. "Synthesis and Characterization of Visible-Light-Responsive TiO2/LDHs Heterostructures for Enhanced Photocatalytic Degradation Performance" Water 17, no. 17: 2582. https://doi.org/10.3390/w17172582

APA Style

Wei, J., & Ren, L. (2025). Synthesis and Characterization of Visible-Light-Responsive TiO2/LDHs Heterostructures for Enhanced Photocatalytic Degradation Performance. Water, 17(17), 2582. https://doi.org/10.3390/w17172582

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