Thi Huyen Nguyenac,
Tien Dung Caoa,
Thi Phuong Maia,
Van Hau Trana,
Van Trinh Pham
ac,
Van Chuc Nguyen
ac,
Van Nhat Phamb,
Viet Tiep Phungc,
Ngoc Minh Phanc,
Van Tan Trand,
Van Hao Nguyen
e,
Huy Tiep Nguyenf and
Van Tu Nguyen
*a
aInstitute of Materials Science, Vietnam Academy of Science and Technology, Hanoi, Vietnam. E-mail: tunv@ims.vast.vn
bUniversity of Science and Technology of Hanoi, Vietnam Academy of Science and Technology, Hanoi, Vietnam
cGraduate University of Science and Technology, Vietnam Academy of Science and Technology, Hanoi, Vietnam
dUniversity of Science, Vietnam National University, Hanoi, Vietnam
eInstitute of Science and Technology, TNU-University of Sciences, Thai Nguyen, Vietnam
fUniversity of Engineering and Technology, Vietnam National University, Hanoi, Vietnam
First published on 18th July 2025
We present a facile synthesis method for a C–TiO2–MoS2-based photocatalytic nanocomposite designed for the efficient degradation of pollutant dyes, specifically methylene blue (MB) and rhodamine B (RhB). Few-layered MoS2 nanosheets were exfoliated from natural bulk MoS2 via urea-assisted ball milling. These nanosheets were then mixed with TiO2 nanopowder and sodium deoxycholate (SDC) surfactant, followed by thermal annealing at 600 °C in argon (Ar) gas. The resulting TiO2–MoS2 mixture was found to be coated and interconnected by a carbon layer, which acted as an electron-conducting bridge and simultaneously prevented the aggregation of TiO2 nanoparticles and MoS2 nanosheets. The photocatalytic performance of the nanocomposite was evaluated through the degradation of MB and RhB. Remarkably, the nanocomposite exhibited excellent degradation of a high-concentration MB (RhB) solution (10−4 M), achieving near-complete degradation within 45 minutes under UV light and 50 minutes under visible light. Furthermore, under UV irradiation, the nanocomposite reached a degradation efficiency of 99.6% with a rate constant of 0.128 min−1, while slightly lower values were observed under visible light. A possible mechanism responsible for the enhanced MB degradation was also proposed. This strategy provides a promising pathway for the development of effective photocatalysts for practical photocatalytic applications.
To overcome these issues, many research groups have developed many strategies to improve further the photocatalytic performance of TiO2, such as doping with non-metallic ions, metal ions,7–10 depositing noble metals.11–13 Recently, a heterostructure formation of TiO2 with carbonaceous nanomaterials or narrow bandgap-semiconductors is an effective approach. Numerous previous studies have reported that combining TiO2 with carbonaceous nanomaterials, for instance, carbon nanotubes,14,15 C60,16,17 graphene oxide,18–20 and carbon dots21,22 can enhance the light absorption of TiO2 across the entire visible spectrum, compared to bare TiO2. The enhanced light absorption was ascribed to the formation of chemical bonds (Ti–O–C) between TiO2 and the carbon-based materials.23 Crucially, carbonaceous materials act as effective electron acceptors, facilitating the transfer of electrons from the photogenerated electron–hole pairs in TiO2, thereby enhancing charge separation efficiency.24,25 However, the low visible light utilization still restricts their wide applications in environmental management.
Different from carbonaceous materials, molybdenum sulfide (MoS2), a member of transition metal dichalcogenides (TMDCs), is a semiconductor with a narrow indirect bandgap of about 1.3 eV in bulk, while MoS2 monolayer possesses a large direct bandgap of 1.9 eV.26,27 This allows MoS2 in the nanocomposite acts as an effective photo-absorber over the solar spectrum. Moreover, the establishment of a type-II band alignment between MoS2 and TiO2 promotes efficient separation of photoinduced electron–hole pairs. Consequently, TiO2/MoS2 composites have garnered growing attention for a wide range of catalytic applications. For instance, Wang et al. developed a MoS2/TiO2 (P25) composite photocatalyst that demonstrated enhanced photocatalytic performance in degrading methylene blue (MB) and rhodamine B (RhB) under simulated sunlight.28 By using a two-step hydrothermal process, Chandrabose et al. created MoS2/TiO2 heterostructure nanocomposites, which demonstrated high photocatalytic activity in textile dye contaminants.29 Tang et al. used a straightforward and scalable one-step sol–gel process to successfully create brown-TiO2@MoS2 heterostructures using 2D MoS2 nanosheets exfoliated from natural bulk molybdenite. Under visible light irradiation, the maximum degradation efficiency of 94.80% and the maximum degradation rate of 0.01764 min−1 were obtained using the heterostructure catalyst. According to experimental findings, as compared to pure brown-TiO2, the creation of brown-TiO2 and 2D MoS2 heterostructures can improve visible light absorption, extra active sites, and the separation and transfer of photogenerated charge.30
Recently, Nguyen et al. synthesized a Z-scheme C–MoS2/TiO2 heterostructure via a hydrothermal method, utilizing a carbon layer as a connecting bridge. The resulting catalyst achieved 99% degradation of methylene blue within 60 minutes under visible light irradiation.31 Ali et al. reported a simple one-step solvothermal synthesis of a C–TiO2–MoS2 heterostructure with visible light activity, utilizing a novel MoS2 cluster compound using a novel MoS2 cluster compound [(NH4)2 Mo3S13·2H2O] as the precursor for the synthesis of MoS2. The composite demonstrates superior photocatalytic activity under visible light for the degradation of RhB and 4-ATP compared to pure TiO2. However, most MoS2 samples used in these studies were synthesized from costly molybdenum and sulfur compounds, which require strict reaction conditions.32
Inspired by the above research, this work aims to develop new photocatalyst material based on C–TiO2–MoS2 nanocomposite. The proposed nanocomposite was prepared using a simple method by mixing TiO2 and lab-made MoS2 nanosheets with SDC solution and followed by the thermal annealing process at 600 °C in Ar gas. The resulting microstructure of the nanocomposite exhibited several key advantages for photocatalytic activity, including enhanced charge carrier separation and reduced aggregation of TiO2 and MoS2. The nanocomposite demonstrated effective photodegradation of high-concentration solutions of the pollutant dyes under both UV and visible light irradiation.
![]() | ||
Fig. 2 (a) SEM, (b) HRTEM image, (c) XRD pattern, (d) Raman spectra, and (e) EDS elemental mapping of C–TiO2–MoS2 nanocomposite. |
XRD analysis was performed to determine the crystal structure and phase composition of the synthesized samples. As shown in Fig. 2c, the XRD pattern of TiO2 (red curve) exhibits distinct diffraction peaks at 25.3° (101), 37.8° (004), and 48.0° (200), which are characteristic of anatase TiO2 (JCPDS #97-002-4276). The pattern of exfoliated MoS2 is also included for comparison. In the case of the C–TiO2–MoS2 nanocomposites, the characteristic peaks corresponding to both anatase TiO2 and exfoliated MoS2 were clearly observed, while diffraction peaks of carbon layer could not be detected. In addition, weak diffraction peaks corresponding to rutile TiO2 were detected, which can be attributed to the partial phase transformation of anatase to rutile induced by thermal annealing at 600 °C. To further confirm the formation of the C–TiO2–MoS2 nanocomposite, Raman measurements also were conducted. As shown in Fig. 2d, we can identify Raman active modes for the anatase phase of TiO2 (black curve) at 145.1 cm−1 (Eg), 400 cm−1 (B1g), 517.3 cm−1 (A1g), and 638.0 cm−1 (Eg), as well as the prominent E1g and A1g peaks of thin MoS2, located at 381.6 cm−1 and 405.4 cm−1, respectively. Additionally, the C–TiO2–MoS2 nanocomposite exhibited a Raman peak at 448.3 cm−1, characteristic of the rutile phase of TiO2, which is attributed to thermal annealing at 600 °C.33 Additionally, peaks at 1346 cm−1 and 1597 cm−1 correspond to the D and G bands of the carbon layer (red curve) were observed.34 These results demonstrated the coexistence of TiO2, carbon, and MoS2 within the C–TiO2–MoS2 nanocomposite.
The EDS mapping of the C–TiO2–MoS2 nanocomposites showed that the distribution of the carbon (C), titanium (Ti), oxygen (O), molybdenum (Mo), and sulfur (S) elements was highly uniform (Fig. 2e).
The N2 adsorption–desorption isotherms and pore size distribution curves of MoS2, TiO2, and C–TiO2–MoS2 were shown in Fig. S3 (ESI†). As shown in Fig. S3,† all the samples exhibited typical type IV isotherm features and H3 hysteresis loops, confirming the presence of mesoporous structures.35 Notably, the C–TiO2–MoS2 nanocomposite demonstrated the significantly higher BET surface area of 62.48 m2 g−1 than that of bare TiO2 (16.3 m2 g−1) and MoS2 nanosheets (6.7 m2 g−1). The pore volumes are 0.8, 0.36, and 0.018 cm3 g−1 for C–TiO2–MoS2 nanocomposite, TiO2 nanopowder, and MoS2 nanosheets, respectively. These results confirm that the C–TiO2–MoS2 nanocomposite has a more developed mesoporous network, which is advantageous for applications requiring high surface area and accessible porosity.
The degradation percentage and degradation rate of MB are two important parameters to estimate the photocatalytic efficiency of the photocatalyst. The decline in intensity of the distinctive absorption peak at 664 nm was used to calculate the degradation percentage.
To determine the degradation rate of MB on the samples, a pseudo-first-order kinetic model was used in this study. The rate of degradation (k) of MB was derived from the slope of the linear transform ln (Ct/C0), where Ct and C0 are the intensity values of the peak at 664 nm at time t and 0, respectively. As shown in Fig. 3g, the C–TiO2–MoS2 nanocomposite shows the highest k value (0.128 min−1), followed by TiO2–MoS2 (0.05 min−1), C–TiO2 (0.013 min−1), bare TiO2 (0.011 min−1), and MoS2 (0.0096 min−1). The obtained results showed that the photodegradation efficiencies of the TiO2–MoS2 and C–TiO2–MoS2 samples are significantly higher than those of bare TiO2 and bare MoS2, C–TiO2. It was evident that the combination of TiO2–MoS2 can provide more active sites and enhance the separation of photogenerated charge carriers due to the formation of type II band alignment between MoS2 and TiO2. Moreover, the introduction of the outer carbon layer created an effective medium for the transport of charge carriers, thereby improving the photocatalytic activity of C–TiO2–MoS2 nanocomposite. More discussion will be presented in the mechanism section.
The stability of the C–TiO2–MoS2 nanocomposite as a photocatalyst under UV irradiation was also thoroughly assessed over five cycles as seen in Fig. 4. The C–TiO2–MoS2 catalyst showed degradation percentages of approximately 99.6%, 97.4%, 96.3%, 94.5%, and 93.2% for MB molecules over the 1st to 5th cycles, respectively. After five cycles, the value of 93.2% was retained, indicating the good stability of C–TiO2–MoS2 nanocomposite.
The photocatalytic performance of the nanocomposite was also systematically evaluated by varying the catalyst dosage, initial pH. When the amount of catalyst was increased from 10 mg to 20 mg/50 mL of MB solution, the degradation efficiency improved as shown in Fig. 5. This is due to the increased number of active sites. However, increasing the nanocomposite loading to 30 mg resulted in a plateau in photocatalytic efficiency, which is attributed to excessive turbidity that hindered light penetration and limited the activation of deeper catalyst layers.
Fig. 6 shows that the initial pH of the solution notably influenced the photocatalytic efficiency, which is attributed to favorable electrostatic interactions between the photocatalyst surface and pollutant molecules. Under acidic condition (pH = 5), the photocatalytic efficiency was decreased while that increased at basic condition (pH = 11), due to catalyst surface charge repulsion or inhibition of reactive species generation.
For real applications, the C–MoS2–TiO2 nanocomposite was also expected to achieve a high degradation efficiency under visible light. As shown in Fig. 7, the MB solution was mostly degraded after 50 min, achieving a degradation efficiency of 99.4% with a corresponding rate constant of k = 0.091 min−1. These values are relatively close to those of the nanocomposite under UV irradiation. This implies that the nanocomposite can operate with high efficiency under both UV and visible light. We made a summary of the photodegradation efficiency of our samples and those of previously reported TiO2-based composites and MoS2-based composites, as shown in Table 1. It should be noted that direct quantitative comparisons are challenging, as experimental conditions such as dye concentration and light source intensity vary considerably among different studies. In this study, thus, we emphasize the ability of the synthesized photocatalyst to effectively degrade methylene blue at significantly higher concentrations (10−4 M) than those typically reported (often ∼10−5 M). This demonstrates the robustness and high catalytic capacity of the material under more practical and challenging conditions. While a full comparative study under identical experimental settings would provide further validation, the current findings highlight the uniqueness of our system in handling high dye concentration, which is relevant for real wastewater treatment applications.
![]() | ||
Fig. 7 (a) Time-dependent UV-vis absorption spectra of 10−4 M MB degradation in the presence of C–TiO2–MoS2 under visible light, (b) MB degradation percentage and (c) degradation rate, respectively. |
Photocatalyst | Dyes | Concentration (mg L−1) | Catalyst loading (g L−1) | Light source | Time (min) | Degradation efficiency (%) | Ref. |
---|---|---|---|---|---|---|---|
MoS2 | MB | 20 | 0.15 | 200 W Xe lamp [λ > 420 nm] | 120 | 5.9 | 36 |
TiO2 | RhB | 5 | 2 | 500 W Xe lamp [λ > 420 nm] | 180 | 19 | 37 |
C–TiO2 | MB | 5 | 0.1 | Philips household CFL, 85W, visible light | 120 | 94 | 21 |
MoS2/TiO2 | MB | 10 | 1 | 500 W, Xenon lamp | 90 | 93.8 | 30 |
MoS2/TiO2 (P25) | MB | 8 | 0.35 | 300 W, Tungsten lamp, (λ = 400–660 nm) | 240 | 32 | 29 |
MoS2/TiO2 | RhB | 10 | — | 230 W UV lamp, [λ = 365 nm] | 150 | 85.3 | 38 |
MoS2/TiO2 | MB | 20 | 0.33 | UV lamp (λ = 360 nm) | 65 | 100 | 39 |
MoS2/graphene/TiO2 (P25) | RhB | 10 | 0.5 | 150 W solar simulator | 80 | 80 | 40 |
C–MoS2/TiO2 | MB | 100 | 0.2 | 150 W Xe lamp [λ = 300–1900 nm] | 60 | 99 | 31 |
MoS2 | MB | 100 | 0.2 | 3 W, UV lamp (λ = 365 nm) | 120 | 94.7 | This work |
TiO2 | MB | 100 | 0.2 | 3 W, UV lamp (λ = 365 nm) | 120 | 92.8 | |
C–TiO2–MoS2 | MB | 100 | 0.2 | 3 W, UV lamp (λ = 365 nm) | 45 | 99.6 | |
3 W, Xenon lamp (λ = 450 nm) | 50 | 99.4 | |||||
RhB | 100 | 0.2 | 3 W, UV lamp (λ = 365 nm) | 45 | 99.5 | ||
3 W, Xenon lamp (λ = 450 nm) | 50 | 99.1 |
To check photocatalytic activity of the nanocomposite with another pollutant, the C–TiO2–MoS2 nanocomposite was further applied to the decomposition of RhB under the same conditions. The same trends were recorded and shown in Fig. S6.†
We try to explain the mechanism behind the enhanced degradation efficiency of the pollutant dyes using our sample under both UV and visible light, we propose the following explanation, illustrated in Fig. 8a. Under UV irradiation, where the photon energy exceeds the bandgap energies of both MoS2 and TiO2, electron–hole pairs are generated in both semiconductors. A significant fraction of the photogenerated electrons from the conduction bands (CB) of MoS2 and TiO2 are rapidly transferred to the carbon layer, which acts as a conductive pathway. This facilitates electron migration and enhances their interaction with surface-adsorbed oxygen molecules. Simultaneously, some electrons from the CB of MoS2 transfer to the CB of TiO2 due to the formation of type-II band alignment. This transfer minimizes electron–hole recombination. The photoinduced holes interact with water molecules or hydroxyl ions on the catalyst surface, leading to the generation of hydroxyl radicals (˙OH), while the photogenerated electrons reduce molecular oxygen (O2) to form superoxide radicals (˙O2−). Consequently, both ˙OH and ˙O2− radicals decompose pollutant dyes into carbon dioxide (CO2) and H2O and intermediates. Under visible light irradiation, electron–hole pairs are primarily generated in the MoS2, while minor electrons are produced in the TiO2 due to the carbon impurity level.41 This explains why the photodegradation efficiency of the nanocomposite catalyst under visible light is slightly lower than that under UV light. To assess the key roles of the active species involved in the photocatalytic reaction, three distinct scavengers—namely isopropyl alcohol (IPA), ethylenediaminetetraacetic acid disodium salt (EDTA-2Na), and benzoquinone (BQ)—were introduced into the reaction system. These agents were employed to selectively quench hydroxyl radicals (˙OH), photogenerated holes (h+), and superoxide radicals (˙O2−), respectively.42,43 As illustrated in Fig. 8b, the degradation efficiency of methylene blue (MB) under UV irradiation decreased from 99.6% to 89.4%, 80.7%, and 57.8% in the presence of BQ, EDTA-2Na, and IPA, respectively. This result indicates that hydroxyl radicals (˙OH) are the dominant reactive species responsible for MB degradation, whereas holes (h+) and superoxide radicals (˙O2−) contribute to a lesser extent. In addition, the total organic carbon (TOC) analysis was also conducted to evaluate the extent of methylene blue (MB) degradation and its conversion into inorganic end-products. As shown in Fig. 8c, the TOC measurements were used to assess the degree of MB mineralization. The C–TiO2–MoS2 nanocomposite achieved a TOC removal efficiency of approximately 72.7%, indicating that a substantial portion of the MB dye was mineralized into CO2 and H2O under UV irradiation. These results demonstrate the strong mineralization capability of the C–TiO2–MoS2 photocatalyst.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra03330f |
This journal is © The Royal Society of Chemistry 2025 |