Next Article in Journal
Volatile Organic Components and MS-e-nose Profiles of Indonesian and Malaysian Palm Sugars from Different Plant Origins
Previous Article in Journal
Banana Flour Adulteration Key Marker Unravelled by Inductively Coupled Plasma Optical Emission Spectrometry Assisted by Chemometric Tools
Previous Article in Special Issue
Influence of P(V3D3-co-TFE) Copolymer Coverage on Hydrogen Detection Performance of a TiO2 Sensor at Different Relative Humidity for Industrial and Biomedical Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Construction of 2D TiO2@MoS2 Heterojunction Nanosheets for Efficient Toluene Gas Detection

School of Materials and Chemistry, University of Shanghai for Science & Technology, Shanghai 200093, China
*
Author to whom correspondence should be addressed.
Chemosensors 2025, 13(5), 154; https://doi.org/10.3390/chemosensors13050154
Submission received: 15 March 2025 / Revised: 14 April 2025 / Accepted: 18 April 2025 / Published: 22 April 2025
(This article belongs to the Special Issue Advanced Chemical Sensors for Gas Detection)

Abstract

:
Monitoring trace toluene exposure is critical for early-stage lung cancer screening via breath analysis, yet conventional chemiresistive sensors face fundamental limitations, including compromised selectivity in complex VOC matrices and humidity-induced signal drift, with prevailing p–n heterojunction architectures suffering from inherent charge recombination and environmental instability. Herein, we pioneer a 2D core–shell n–n heterojunction strategy through rational design of TiO2@MoS2 heterostructures, where vertically aligned MoS2 nanosheets are epitaxially grown on 2D TiO2 derived from graphene-templated synthesis, creating built-in electric fields at the heterojunction interface that dramatically enhance charge carrier separation efficiency. At 240 °C, the TiO2@MoS2 sensor exhibits a superior response (Ra/Rg = 9.8 to 10 ppm toluene), outperforming MoS2 (Ra/Rg = 2.8). Additionally, the sensor demonstrates rapid response/recovery kinetics (9 s/16 s), a low detection limit (50 ppb), and excellent selectivity against interfering gases and moisture. The enhanced performance is attributed to unidirectional electron transfer (TiO2 → MoS2) without hole recombination losses, methyl-specific adsorption through TiO2 oxygen vacancy alignment, and steric exclusion of non-target VOCs via size-selective MoS2 interlayers. This work establishes a transformative paradigm in gas sensor design by leveraging n–n heterojunction physics and 2D core–shell synergy, overcoming long-standing limitations of conventional architectures.

1. Introduction

Toluene is a common volatile organic compound (VOC) that can damage the central nervous system, causing headaches, dizziness, and liver and kidney toxicity with long-term exposure. High-concentration inhalation can lead to acute poisoning. Toluene also serves as a potential biomarker for diagnosing diseases like lung cancer and diabetes [1]. Traditional detection methods, such as gas chromatography–mass spectrometry (GC-MS), are accurate, but require expensive equipment and laboratory settings, preventing real-time monitoring. Electrochemical sensors face cross-interference issues, and photoionization detectors (PIDs) are costly to maintain. Therefore, developing high-sensitivity, low-cost, portable semiconductor toluene gas sensors through material innovation and intelligent design is crucial [2,3,4]. Such sensors could enable home health monitoring, industrial safety warnings, and early disease screening, promoting advancements in precision medicine and environmental governance [5].
Two-dimensional transition metal dichalcogenides (TMDs) have emerged as promising candidates for next-generation gas sensors, due to their atomic-scale thickness, tunable electronic properties, and high surface-to-volume ratios [6]. Molybdenum disulfide (MoS2), a graphene-like layered TMD, has garnered significant attention. In its thermodynamically stable 2H phase, the crystal structure consists of S-Mo-S trilayers stacked via van der Waals interactions, with each Mo atom in trigonal prismatic coordination [7,8,9]. The hydrothermal synthesis of MoS2 enables precise control over layer number (typically one to five layers) and defect density, yielding nanostructures with specific surface areas and direct bandgaps suitable for charge transfer-based gas sensors [10,11]. Notably, 2D MoS2 nanosheets exhibit superior gas adsorption kinetics compared to their 0D quantum dot or 1D nanobelt due to their exposed basal planes (90% surface accessibility) and abundant edge sulfur vacancies, which serve as preferential adsorption sites [12]. Additionally, in the case of ultrathin MoS2 layers, the sulfide terminations at MoS2 edges can achieve maximal exposure. Thus, constructing heterostructures to vertically grow such thin 2D MoS2 represents a viable strategy [13].
Heterojunction engineering effectively enhances sensing performance by lowering operating temperatures and improving sensitivity, selectivity, and humidity resistance in nanomaterials [14,15]. Due to the band structure mismatch between two sensing materials, energy bands bend to form electron- depletion layers (EDLs) or hole accumulation layers (HALs) when the Fermi levels equilibrate [16,17]. Furthermore, a hierarchical composite structure can provide additional active adsorption sites for target gases and exhibit superior catalytic activity compared to single metal oxides. Thus, heterostructures are critical for optimizing the gas-sensing properties of metal oxides [18,19]. Hermawan et al. [20] reported a p–n CuO@SnO2 heterojunction that achieved a remarkable toluene response (Ra/Rg = 540 at 75 ppm) with exceptional selectivity. The enhanced performance was attributed to the synergistic effects of the p–n heterojunction, high surface area, and porous architecture. Notably, in situ formation of metallic Cu under high toluene concentrations disrupted the p–n junction and established Ohmic contact with the n-type SnO2, further amplifying the sensing response. Liu et al. [21] fabricated a porous Co3O4–Fe3O4 composite with p–n heterojunctions by functionalizing Fe3O4 with ZIF-67-derived Co3O4. The Co3O4 derived from ZIF-67 exhibited a higher specific surface area than Fe3O4, leading to a 20-fold improvement in toluene response at 225 °C. Additionally, the sensor demonstrated a low detection limit (0.1 ppm, response = 1.68), excellent selectivity, and long-term stability. This enhancement primarily stemmed from the p–n heterojunction, increased surface area, and optimized oxygen vacancy (OV) concentration. Therefore, heterojunction construction provides a viable pathway for developing highly sensitive and practical gas sensors based on nanostructured metal oxides. The integration of vertically aligned ultrathin MoS2 nanosheets with a wide-bandgap metal oxide semiconductor (MOS) core constructs a 2D core–shell n–n heterojunction, where the MOS core acts as an electron donor and the MoS2 shell serves as an electron acceptor [14]. A built-in electric field forms at the heterointerface through type II band alignment, significantly enhancing carrier separation efficiency, while the core–shell structure achieves specific gas recognition through the synergistic interface coupling effect between MOS and MOS2 components [6].
Here, we report a unique 2D heterostructure TiO2@MoS2 nanosheet. Two-dimensional TiO2 nanosheets were synthesized via a graphene sacrificial template method, followed by the hydrothermal growth of fish scale-like MoS2 nanosheets on their surfaces. By adjusting the hydrothermal reaction time, MoS2 layers with controlled thicknesses were achieved, exhibiting an average lateral size of 193 ± 21 nm. Compared to pristine MoS2 nanosheets, the TiO2@MoS2 heterostructure demonstrated a superior toluene response (Ra/Rg = 9.8@10 ppm) at 240 °C, along with rapid response/recovery kinetics (9 s/16 s), excellent stability, durability, and high humidity resistance. The enhanced gas-sensing performance is attributed to the n–n heterojunction and high specific surface area. The synergistic interaction between the two components and the unique core–shell hierarchical architecture with self-assembled nanostructures collectively amplify the sensor’s capabilities. This study provides a rational material design strategy, offering a promising candidate for efficient detection of toluene gas.

2. Materials and Methods

2.1. Synthesis of Sensor Materials

Synthesis of TiO2 Nanosheets: TiO2 precursor was synthesized using a graphene oxide (GO) sacrificial template method. Specifically, 100 mg of GO was dispersed in 200 mL of ethanol and ultrasonicated for 1 h. Subsequently, 2.5 mmol of titanium butoxide (TBOT) was added dropwise under continuous stirring, and the mixture was stirred overnight (18 h) to obtain a gray suspension. The product was thoroughly washed three times with deionized water and ethanol, followed by drying at 80 °C overnight. Finally, the dried powder was calcined in a muffle furnace at 500 °C for 2 h with a heating rate of 2 °C/min, yielding white TiO2 nanosheets.
Synthesis of TiO2@MoS2 Nanocomposites: 0.1 mmol of ammonium heptamolybdate tetrahydrate (NH4)2MoO4·4H2O) was dissolved in 10 mL of deionized water under vigorous stirring. A predetermined amount of the as-prepared TiO2 nanosheets was added to the solution, and the mixture was incubated at room temperature for 12 h and dried at 80 °C for 12 h. Separately, 0.1 mmol of (NH4)2MoO4·4H2O and 3.05 mmol of thiourea (CH4N2S) were dissolved in 40 mL of deionized water. The (NH4)2MoO4·4H2O-loaded TiO2 nanosheets were then added to the solution, and the mixture was stirred at room temperature for 1 h. The homogeneous suspension was transferred to a 50 mL Teflon-lined stainless-steel autoclave and heated at 200 °C for 8 h. After cooling to room temperature, the black precipitate was collected, washed three times with deionized water and ethanol, and dried at 60 °C overnight to obtain TiO2@MoS2-8 nanosheets. By adjusting the reaction time to 4 h and 12 h under identical conditions, samples labeled TiO2@MoS2-4 and TiO2@MoS2-12 were prepared. For comparison, pure MoS2 nanosheets were synthesized using the same procedure without the addition of TiO2 nanosheets.

2.2. Gas-Sensing Performance Measurement

The as-prepared TiO2 NSs, TiO2@MoS2 NSs, and MoS2 NSs were mixed with deionized water to form a homogeneous paste. Each paste was uniformly coated onto a ceramic substrate equipped with a pair of gold electrodes using a fine brush (Figure S1). The operating temperature of the sensor was controlled by adjusting the heating current applied to the embedded heating electrode. The gas-sensing performance was evaluated using a static test system (Elite Technology Co., Ltd. Suzhou, China). Target gas concentrations were generated via a static gas distribution method.
The gas-sensing performance was evaluated using a custom-built static testing system, where sensor resistance was monitored continuously with a source meter under controlled conditions. Target analytes were introduced into the 20 L test chamber by injecting liquid toluene with a microsyringe and vaporizing it on a hotplate. The volume Q can be determined by:
Q = V C M 22.4 D ρ × 10 9 × 273 + T R 273 + T B
Here, V, C, M, d, ρ, TR, and TB are the test chamber volume (L), vapor concentration (ppm), molecular mass (g/mol), liquid density (kg/m3), liquid purity (%), environmental temperature (°C), and temperature (°C) in the testing chamber, respectively.
Sensitivity (S), a critical performance metric of gas sensors, is quantified as the resistance ratio under distinct gaseous environments. For n-type semiconductors, sensitivity is defined as:
S = R a / R g   ( N t y p e )   ( o x i d i z i n g   g a s e s )
S = R g / R a   ( N t y p e )   ( r e d u c i n g   g a s e s )
where Ra denotes the baseline resistance in dry air and Rg represents the steady-state resistance upon exposure to the target analyte. Conversely, p-type semiconductors exhibit inverse trends.

3. Results and Discussion

3.1. Morphology and Structure Characterization

As demonstrated in Figure 1, TiO2 nanosheets were synthesized via a graphene oxide sacrificial template method followed by annealing at 500 °C. Subsequently, fish scale-like MoS2 nanosheet arrays were uniformly grown on the TiO2 surfaces through a controlled hydrothermal reaction (Figure 1). Optimizing precursor concentration and hydrothermal reaction time was crucial for MoS2 growth. The thickness of MoS2 nanostructures could be effectively tuned by adjusting the hydrothermal duration. The facile and reproducible synthesis protocol offers significant advantages for scalable production.
To characterize the morphology and structure of TiO2@MoS2 composites, SEM and low-magnification TEM analyses were performed. As shown in Figure 2a, ultrathin TiO2 nanosheets (~2.1 ± 0.3 nm thickness) were successfully synthesized via the graphene oxide sacrificial template method. In contrast, pristine MoS2 formed spherical aggregates with an average diameter of 530 ± 12 nm (Figure 2b). Figure 2c–e reveal the hierarchical heterostructure, where vertically aligned MoS2 nanosheets (193 ± 21 nm lateral size) uniformly coat TiO2 surfaces, forming an n–n heterojunction. The MoS2 coverage density exhibited a positive correlation with hydrothermal duration, enabling thickness-controlled growth. TEM characterization (Figure 2f) further confirmed the 2D–2D interfacial architecture. HR-TEM analysis (Figure 2g) resolved distinct lattice spacings of 0.250 nm, corresponding to the (102) plane of MoS2 [22,23,24,25]. The polycrystalline nature was verified by selective electron diffraction (SAED) patterns showing concentric diffraction rings (Figure 2h). Energy dispersion spectrum (EDS) elemental mapping (Figure 2i) demonstrated homogeneous spatial distribution of Ti, Mo, S, and O, conclusively confirming the formation of well-integrated TiO2@MoS2 heterostructures.
To determine the phase composition and interfacial interactions in the TiO2@MoS2 heterostructure, X-ray diffraction (XRD) and Raman analyses were performed. The XRD pattern of TiO2@MoS2 (Figure 3a) revealed distinct diffraction peaks at 25.3° (101), 37.8° (004), and 48.0° (200), corresponding to anatase TiO2 (JCPDS #97-002-4276), alongside peaks at 31.6° (102) and 45.5° (110), attributed to MoS2 (JCPDS #97-064-4259). A noticeable shift in the (004) peak of TiO2 and the (110) peak of MoS2 was observed, indicative of lattice strain arising from the intimate interfacial coupling in the heterojunction.
To analyze the structural advantages of the TiO2@MoS2 nanocomposite, nitrogen adsorption–desorption measurements were conducted to determine its specific surface area and pore distribution using BET theory and BJH methods. As shown in Figure 3b, all three materials (TiO2, MoS2, TiO2@MoS2-8) exhibit type IV isotherms with H3-type hysteresis loops in the relative pressure (P/P0) range of 0.6–1.0, characteristic of mesoporous structures. The average pore diameters determined by BJH analysis were 2.4 nm (TiO2), 3.1 nm (MoS2), and 8.3 nm (TiO2@MoS2-8). Notably, the TiO2@MoS2-8 composite demonstrated the highest BET surface area of 87.48 m2/g, significantly surpassing that of pristine TiO2 nanosheets (9.04 m2/g) and MoS2 nanosheets (19.68 m2/g), confirming the heterojunction’s role in enhancing porosity (Figure 3c).
X-ray photoelectron spectroscopy (XPS) was employed to investigate the surface composition and chemical states of TiO2, MoS2, and TiO2@MoS2-8 nanocomposites. As shown in Figure 4a, the survey spectra confirm the coexistence of Ti, Mo, S, and O in the TiO2@MoS2-8 heterostructure, consistent with EDS, XRD, and HRTEM results, with no impurity peaks except for residual carbon (C 1s). High-resolution Ti 2p spectra (Figure 4b) reveal characteristic peaks at 459.08 eV (Ti 2p3/2) and 464.87 eV (Ti 2p1/2) for TiO2@MoS2-8, with a spin–orbit splitting of 5.79 eV, typical of Ti4+ in TiO2. Notably, a 0.40 eV positive binding energy shift compared to pure TiO2 (458.68 eV and 464.47 eV) indicated electron transfer from TiO2 to MoS2. The Mo 3d spectrum (Figure 4c) exhibited peaks at 228.65 eV (Mo 3d5/2) and 231.87 eV (Mo 3d3/2), with a corresponding S 2s peak at 225.90 eV, while a 0.60 eV negative shift relative to pure MoS2 further confirmed interfacial charge redistribution. Similarly, the S 2p peaks (Figure 4d) at 162.06 eV (S 2p3/2) and 163.16 eV (S 2p1/2) show a 0.15 eV shift, corroborating strong electronic interactions. Deconvolution of O 1s spectra (Figure 4e) identifies three components: lattice oxygen (Olat, 529.5 ± 0.6 eV), oxygen vacancies (Odef, 531.6 ± 0.4 eV), and adsorbed oxygen species (Oads, 532.5 ± 0.6 eV). TiO2@MoS2-8 demonstrates elevated Odef (22.72%) and Oads (4.58%) concentrations compared to TiO2, which aligns with its enhanced ESR signal intensity (Figure 4f) reflecting higher oxygen vacancy density. These synergistic effects of enhanced charge transfer and oxygen-mediated surface activity collectively improved toluene-sensing performance by promoting gas adsorption and redox reactions at the heterojunction interface.

3.2. Gas-Sensitive Property

TiO2@MoS2 nanocomposites show great potential as chemiresistive gas sensors due to their high surface area and hierarchical 2D structures. The operating temperature affects surface resistance by altering adsorbed oxygen species and reaction kinetics. This is due to faster surface reactions at higher temperatures competing with reduced gas molecule adsorption because of fewer available active sites [26]. Temperature-dependent sensing tests were conducted from 200 to 280 °C for MoS2, TiO2@MoS2-4, TiO2@MoS2-8, and TiO2@MoS2-12. The temperature-dependent response exhibited a characteristic volcano profile (Figure 5a), peaking at 240 °C. Below this temperature, inadequate thermal activation hindered oxygen ionosorption and toluene oxidation. Beyond 240 °C, accelerated desorption prevailed over surface reactions, reducing sensor response value [27]. The TiO2@MoS2-8 composite achieved a maximum response of 9.8@10 ppm toluene, representing a 3.5-fold enhancement over pristine MoS2 (2.8), primarily due to oxygen vacancy-mediated toluene adsorption. Notably, TiO2@MoS2-8 exhibited rapid response/recovery kinetics (9 s/16 s, Figure 5b), outperforming other samples (MoS2: 4.1 s/18 s; TiO2@MoS2-4: 10 s/17 s; TiO2@MoS2-12: 12 s/20 s), benefiting from its optimized mesoporous structure facilitating gas diffusion. Selectivity tests against xylene, benzene, triethylamine, acetone, methanol, ethanol, and ammonia (Figure 5c) revealed superior toluene selectivity. Even under simulated exhaled breath conditions containing mixed interferents (Figure 5d), the sensor maintained stable toluene detection, confirming its robustness in complex environments. These critical attributes of high sensitivity, rapid reaction kinetics, and selective discrimination collectively establish TiO2@MoS2 nanocomposites as promising materials for toluene detection applications.
Figure 5e illustrates the dynamic response of TiO2@MoS2-based sensors to varying concentrations of toluene gas, ranging from 50 ppb to 10 ppm. The sensor response demonstrates stepwise enhancement with increasing concentration, with TiO2@MoS2 composites exhibiting superior sensitivity compared to pristine MoS2, particularly at higher concentrations. A limit of detection (LOD) of 50 ppb was achieved for the optimal TiO2@MoS2-8 sensor. As shown in Figure 5f, all samples maintain linear responses across the 0.05–10 ppm range, validating quantitative detection capability. Repeatability tests (Figure 5g) revealed stable performance over five cycles at 10 ppm, with response variations below ±5.2%. Under humidity challenges (25%–75% RH, Figure 5h), the response attenuation remained <20%, underscoring robust moisture resistance. Long-term stability assessments (30 days exposure to 10 ppm toluene, Figure 5i) showed <8% signal drift, confirming durability for practical deployment. Table 1 shows the comparison of sensor-based toluene gas detection characteristics with different detection materials, further confirming the great potential of TiO2@MoS2-based material as a toluene sensor due to its lower detection limit and rapid response/recovery time.
As shown in Figure 6a, time-resolved in situ Fourier transform infrared (in situ FTIR) spectroscopy was employed to analyze intermediate products during toluene sensing. Peaks observed in the 1450–1600 cm−1 range originate from C=C aromatic ring vibrations, with increasing intensity over time indicating progressive aromatic compound accumulation. Absorption bands between 1370–1450 cm−1 correspond to C-H bending modes of methyl groups directly bonded to the benzene ring, confirming toluene decomposition. Concurrently, diminishing toluene-specific peaks and emerging signals at 3000–3100 cm−1 (aromatic C-H stretching) suggest catalytic conversion pathways where toluene transforms into benzene derivatives. Resistance changes in TiO2@MoS2 nanocomposites arise from oxygen-mediated surface interactions: at 240 °C, chemisorbed oxygen species (O2, O) populate oxygen vacancies, forming electron-depletion layers that reduce baseline resistance in air (Equation (4)) [31]. Upon toluene exposure, adsorbed oxygen species react with methyl groups via stepwise dehydrogenation (Equations (5)–(8)), modulating conductivity through electron transfer between the semiconductor and reaction intermediates. This mechanistic framework aligns with the observed spectral evolution and sensor response dynamics.
O 2 g O 2 a d s
O 2 a d s + e O 2 a d s
O 2 a d s + e 2 O a d s
C 7 H 8 g + 3 O a d s C 6 H 6 g + C O 2 + H 2 O g + 3 e
C 6 H 6 g + 15 O a d s 6 C O 2 + 3 H 2 O g + 15 e
The enhanced toluene-sensing performance of the 2D stacked MoS2@TiO2 heterostructure compared to pristine MoS2 and TiO2 nanosheets arises from three synergistic mechanisms, as follows.
(I) Hierarchical architecture and surface engineering. Gas–solid interactions in chemiresistive sensors critically depend on surface-mediated oxygen adsorption and target gas diffusion. The MoS2@TiO2 heterostructure exhibits a superior specific surface area, providing abundant active sites for oxygen chemisorption. The vertically aligned MoS2 nanosheets on TiO2 substrates prevent nanosheet restacking while creating interconnected nanochannels that enhance toluene diffusion kinetics.
(II) Interface charge redistribution in n–n heterojunctions. The work function disparity between MoS2 (Φ = 4.05 eV) and TiO2 (Φ = 4.77 eV), determined by ultraviolet photoelectron spectroscopy (UPS, Figure S3), drives interfacial electron transfer from TiO2 to MoS2 [32,33]. This creates a built-in electric field at the heterointerface, forming an electron-depletion layer, which lowers the oxygen adsorption activation energy and facilitates hole–electron separation efficiency [34]. Band alignment analysis (Figure S2) reveals heterojunction characteristics with MoS2 and TiO2 bandgaps of 3.20 eV and 0.85 eV, respectively [35,36,37]. The conduction band offset promotes electron trapping by adsorbed oxygen species (O2/O), amplifying resistance modulation upon toluene exposure.
(III) Influence by oxygen species. Compared with TiO2 nanosheets, TiO2@MoS2 has the highest ESR intensity and the highest relative concentration of oxygen vacancy. Therefore, the TiO2@MoS2 hierarchical structure can provide additional Oads and Odef, which is conducive to the reaction of the target gas with oxygen, while increasing the surface active site, which is conducive to the sensing performance of toluene [38,39].

4. Conclusions

In summary, TiO2@MoS2 core–shell heterostructures were fabricated by assembling MoS2 nanosheet arrays onto TiO2 nanosheets prepared via a sacrificial template method. The unique 2D hierarchical heterojunction structure resulted in a larger specific surface area and increased generation of oxygen vacancies. Compared with individual TiO2 nanosheets and MoS2 nanosheets, the TiO2@MoS2 composite demonstrated the highest response (Ra/Rg = 9.8@10 ppm toluene) at 240 °C. Furthermore, the hierarchical TiO2@MoS2 core–shell structure exhibited fast response/recovery times (9 s/16 s), a low detection limit (50 ppb), excellent selectivity, and robust stability in humid environments. The enhanced gas-sensing performance is attributed to the synergistic effects between the two components and the distinctive core–shell hierarchical architecture with self-assembled nanostructures. This work provides a rational material structure design strategy, offering a promising candidate for detecting toluene gas, thereby demonstrating significant potential value for advancing environmental or disease diagnosis detection.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/chemosensors13050154/s1. Figure S1: Ceramic plate and test base; Figure S2: (a–d) UV-vis diffuse reflectance spectra (DRS) of TiO2 and MoS2 nanosheets with corresponding Tauc plots for bandgap determination; Figure S3: Ultraviolet photoelectron spectroscopy (UPS) spectra of pristine TiO2 and MoS2 recorded under He Iα excitation (21.22 eV), showing secondary electron cutoff (left) and valence band regions (right).

Author Contributions

D.W. (Dehui Wang): Investigation; writing original draft; conceptualization; visualization. J.H.: Investigation; conceptualization; writing original draft. H.X.: Writing, review & editing; project administration. D.W. (Ding Wang): Funding acquisition; project administration; writing, review & editing. G.L.: Funding acquisition; project administration. All authors have read and agreed to the published version of the manuscript.

Funding

The authors greatly appreciate the financial support from National Natural Science Foundation of China (62471303, 62071300, 22176127, 22301181, 22406130, 22476131), Shanghai Sailing Program (23YF1429000, 22YF1430400). We also acknowledge the USST Center for Instrumental Analysis for their support in materials characterization and High Performance Computing Center of USST for their support in the calculations.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Acknowledgments

We acknowledge the USST Center for Instrumental Analysis for their support in material characterization.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ueda, T.; Hayashi, H.; Tsukahara, R.; Shimizu, Y.; Hyodo, T. Effects of Structure and Thickness of Ce0.9Pr0.1O2 Electrodes of YSZ-Based Gas Sensors on VOC-Sensing Properties. Sens. Actuators B Chem. 2025, 422, 136580. [Google Scholar] [CrossRef]
  2. Zhang, Y.-L.; Jia, C.-W.; Tian, R.-N.; Guan, H.-T.; Chen, G.; Dong, C.-J. Hierarchical Flower-like NiFe2O4 with Core–Shell Structure for Excellent Toluene Detection. Rare Met. 2021, 40, 1578–1587. [Google Scholar] [CrossRef]
  3. Wang, T.; Liu, S.; Sun, P.; Wang, Y.; Shimanoe, K.; Lu, G. Unexpected and Enhanced Electrostatic Adsorption Capacity of Oxygen Vacancy-Rich Cobalt-Doped In2O3 for High-Sensitive MEMS Toluene Sensor. Sens. Actuators B Chem. 2021, 342, 129949. [Google Scholar] [CrossRef]
  4. Liao, Z.; Chu, N.; Yuan, Z.; Shen, Y.; Meng, F. Highly Sensitive Toluene Sensor Based on Silver Ear-Like F-Doped Co3O4 Templated from Co(OH)F Guided by SiO2 Nanospheres. Chem. Eng. J. 2024, 498, 155688. [Google Scholar] [CrossRef]
  5. Zhang, H.; Hu, J.; Li, M.; Li, Z.; Yuan, Y.; Yang, X.; Guo, L. Highly Efficient Toluene Gas Sensor Based on Spinel Structured Hollow Urchin-like Core-Shell ZnFe2O4 Spheres. Sens. Actuators B Chem. 2021, 349, 130734. [Google Scholar] [CrossRef]
  6. Siraj, S.; Bansal, G.; Hasita, B.; Srungaram, S.; S, S.K.; Rybicki, F.J.; Sonkusale, S.; Sahatiya, P. MXene/MoS2 Piezotronic Acetone Gas Sensor for Management of Diabetes. ACS Appl. Nano Mater. 2024, 7, 11350–11361. [Google Scholar] [CrossRef]
  7. Thayil, R.; Krishna, K.G.; Cherukulappurath, S.; Kathirvelu, V.; Parne, S.R. MoS2 and MoS2-Based Nanocomposites for Enhanced Toluene Sensing Response at Room Temperature. Surf. Interfaces 2024, 46, 104134. [Google Scholar] [CrossRef]
  8. Tang, T.; Li, Z.; Liu, Y.Y.; Chen, Y.L.; Cheng, Y.F.; Liang, Y.; Zhuang, J.H.; Hu, X.Y.; Jannat, A.; Ou, R.; et al. Ultrathin two-dimensional titanium oxysulfide for enhanced sensitivity and stability of room temperature NO2 sensing. Ceram. Int. 2025, 51, 3216–3223. [Google Scholar] [CrossRef]
  9. Guan, X.; Zhao, L.; Zhang, P.; Liu, J.; Song, X.; Gao, L. Electrode Material of Core-Shell Hybrid MoS2@C/CNTs with Carbon Intercalated Few-Layer MoS2 Nanosheets. Mater. Today Energy 2020, 16, 100379. [Google Scholar] [CrossRef]
  10. Gough, J.J.; McEvoy, N.; O’ Brien, M.; Bell, A.P.; McCloskey, D.; Boland, J.B.; Coleman, J.N.; Duesberg, G.S.; Bradley, A.L. Dependence of Photocurrent Enhancements in Quantum Dot (QD)-Sensitized MoS2 Devices on MoS2 Film Properties. Adv. Funct. Mater. 2018, 28, 1706149. [Google Scholar] [CrossRef]
  11. Zhang, W.; Liu, Y.; Pei, X.; Yuan, Z.; Zhang, Y.; Zhao, Z.; Hao, H.; Long, R.; Liu, N. Stretchable MoS2 Artificial Photoreceptors for E-Skin. Adv. Funct. Mater. 2022, 32, 2107524. [Google Scholar] [CrossRef]
  12. Chen, Y.L.; Li, Z.; Tang, T.; Cheng, Y.F.; Cheng, L.; Wang, X.X.; Haidry, A.A.; Jannat, A.; Ou, J.Z. Room-Temperature Optoelectronic NO2 Sensing Using Two-Dimensional Gallium Oxyselenides. ACS Appl. Nano Mater. 2024, 7, 3229–3238. [Google Scholar] [CrossRef]
  13. Wang, H.; Shao, Z.; Shi, X.; Tang, Z.; Sun, B. Rapidly Detecting the Carcinogen Acetaldehyde: Preparation and Application of a Flower-like MoS2 Cataluminescence Sensor at Low Working Temperature. Anal. Methods 2023, 15, 5620–5629. [Google Scholar] [CrossRef] [PubMed]
  14. Niu, Y.; Zeng, J.; Liu, X.; Li, J.; Wang, Q.; Li, H.; de Rooij, N.F.; Wang, Y.; Zhou, G. A Photovoltaic Self-Powered Gas Sensor Based on All-Dry Transferred MoS2/GaSe Heterojunction for ppb-Level NO2 Sensing at Room Temperature. Adv. Sci. 2021, 8, 2100472. [Google Scholar] [CrossRef] [PubMed]
  15. Cheng, X.; Yao, Y.T.; Zheng, S.L.; Wan, Y.; Wei, C.D.; Yang, G.C.; Yuan, Y.; Tsai, H.-S.; Wang, Y.; Hao, J.Y. Te@Se Core–Shell Heterostructures with Tunable Shell Thickness for Ultra-Stable NO2 Detection. ACS Sens. 2025, 10, 283–291. [Google Scholar] [CrossRef]
  16. Shi, J.; Xiong, J.; Qiao, L.; Liu, C.; Zeng, Y. Facile MOF-on-MOF Isomeric Strategy for ZnO@Co3O4 Single-Shelled Hollow Cubes with High Toluene Detection Capability. Appl. Surf. Sci. 2023, 609, 155271. [Google Scholar] [CrossRef]
  17. Liu, H.; Wang, Z.; Cao, G.; Pan, G.; Yang, X.; Qiu, M.; Sun, C.; Shao, J.; Li, Z.; Zhang, H. Construction of Hollow NiO/ZnO p-n Heterostructure for Ultrahigh Performance Toluene Gas Sensor. Mater. Sci. Semicond. Process. 2022, 141, 106435. [Google Scholar] [CrossRef]
  18. Ma, S.; Xu, J. Nanostructured Metal Oxide Heterojunctions for Chemiresistive Gas Sensors. J. Mater. Chem. A 2023, 11, 23742–23771. [Google Scholar] [CrossRef]
  19. Wang, Y.; Tan, L.; Tan, M.H.; Zhang, P.P.; Yoshiharu, Y.; Yang, G.H.; Tsubaki, N. Rationally Designing Bifunctional Catalysts as an Efficient Strategy to Boost CO2 Hydrogenation Producing Value-Added Aromatics. ACS Catal. 2018, 9, 895–901. [Google Scholar] [CrossRef]
  20. Hermawan, A.; Asakura, Y.; Inada, M.; Yin, S. A Facile Method for Preparation of Uniformly Decorated-Spherical SnO2 by CuO Nanoparticles for Highly Responsive Toluene Detection at High Temperature. J. Mater. Sci. Technol. 2020, 51, 119–129. [Google Scholar] [CrossRef]
  21. Liu, H.; Wang, Z.; Sun, C.; Shao, J.; Li, Z.; Zhang, H.; Qiu, M.; Pan, G.; Yang, X. Construction of Co3O4/Fe3O4 Heterojunctions from Metal Organic Framework Derivatives for High Performance Toluene Sensor. Sens. Actuators B Chem. 2023, 375, 132863. [Google Scholar] [CrossRef]
  22. Raghu, A.V.; Karuppanan, K.K.; Nampoothiri, J.; Pullithadathil, B. Wearable, Flexible Ethanol Gas Sensor Based on TiO2 Nanoparticles-Grafted 2D-Titanium Carbide Nanosheets. ACS Appl. Nano Mater. 2019, 2, 1152–1163. [Google Scholar] [CrossRef]
  23. Cao, S.; Sui, N.; Zhang, P.; Zhou, T.; Tu, J.; Zhang, T. TiO2 Nanostructures with Different Crystal Phases for Sensitive Acetone Gas Sensors. J. Colloid Interf. Sci. 2022, 607, 357–366. [Google Scholar] [CrossRef]
  24. Zhao, Y.; Song, J.-G.; Ryu, G.H.; Ko, K.Y.; Woo, W.J.; Kim, Y.; Kim, D.; Lim, J.H.; Lee, S.; Lee, Z.; et al. Low-Temperature Synthesis of 2D MoS2 on a Plastic Substrate for a Flexible Gas Sensor. Nanoscale 2018, 10, 9338–9345. [Google Scholar] [CrossRef]
  25. Kumar, R.; Zheng, W.; Liu, X.; Zhang, J.; Kumar, M. MoS2-Based Nanomaterials for Room-Temperature Gas Sensors. Adv. Mater. Technol. 2020, 5, 1901062. [Google Scholar] [CrossRef]
  26. Verma, G.; Gokarna, A.; Kadiri, H.; Nomenyo, K.; Lerondel, G.; Gupta, A. Multiplexed Gas Sensor: Fabrication Strategies, Recent Progress, and Challenges. ACS Sens. 2023, 8, 3320–3337. [Google Scholar] [CrossRef]
  27. Korotcenkov, G.; Cho, B.K. Metal Oxide Composites in Conductometric Gas Sensors: Achievements and Challenges. Sens. Actuators B Chem. 2017, 244, 182–210. [Google Scholar] [CrossRef]
  28. Gong, C.; Chen, M.; Song, F.; Yin, P.; Zhao, X.; You, X.; Fu, H.; Yu, S.; Liu, X.; Zhang, K.; et al. A Highly Sensitive Toluene Gas Sensor Based on Pd/PdO Decorated SnO2 Prepared by Electrospinning. ACS Appl. Electron. Mater. 2024, 6, 6036–6048. [Google Scholar] [CrossRef]
  29. Karmakar, S.; Deepak, M.S.; Nanda, O.P.; Sett, A.; Maity, P.C.; Karmakar, G.; Sha, R.; Badhulika, S.; Bhattacharyya, T.K. 2D V2 C MXene Based Flexible Gas Sensor for Highly Selective and Sensitive Toluene Detection at Room Temperature. ACS Appl. Electron. Mater. 2024, 6, 3717–3725. [Google Scholar] [CrossRef]
  30. Gao, H.; Zhao, L.; Wang, L.; Sun, P.; Lu, H.; Liu, F.; Chuai, X.; Lu, G. Ultrasensitive and Low Detection Limit of Toluene Gas Sensor Based on SnO2-Decorated NiO Nanostructure. Sens. Actuators B Chem. 2018, 255, 3505–3515. [Google Scholar] [CrossRef]
  31. Jung, G.; Shin, W.; Hong, S.; Jeong, Y.; Park, J.; Kim, D.; Bae, J.-H.; Park, B.-G.; Lee, J.-H. Comparison of the Characteristics of Semiconductor Gas Sensors with Different Transducers Fabricated on the Same Substrate. Sens. Actuators B Chem. 2021, 335, 129661. [Google Scholar] [CrossRef]
  32. Hu, J.W.; Wang, F.; Yu, J.J.; Hong, Z.J.; Zhang, W.H.; Li, H.-J.; Lai, Z.C.; Wang, D.; Deng, Y.H.; Li, G.S. MOF-derived porous Co3O4 nanosheets array assembled on SnO2 nanofibers for humidity-resistant high efficiency acetone detection. Chin. Chem. Lett. 2025, 110863. [Google Scholar] [CrossRef]
  33. Lashkov, A.V.; Fedorov, F.S.; Vasilkov, M.Y.; Kochetkov, A.V.; Belyaev, I.V.; Plugin, I.A.; Varezhnikov, A.S.; Filipenko, A.N.; Romanov, S.A.; Nasibulin, A.G.; et al. The Ti Wire Functionalized with Inherent TiO2 Nanotubes by Anodization as One-Electrode Gas Sensor: A Proof-of-Concept Study. Sens. Actuators B Chem. 2020, 306, 127615. [Google Scholar] [CrossRef]
  34. Shokri, A.; Salami, N. Gas Sensor Based on MoS2 Monolayer. Sens. Actuators B Chem. 2016, 236, 378–385. [Google Scholar] [CrossRef]
  35. Hou, J.-L.; Lin, Y.-T.; Hsueh, T.-J. Room-Temperature TiO2 Gas Sensor with Conical-TSV Using a Thermal Oxidation Process. ACS Appl. Electron. Mater. 2023, 5, 6542–6548. [Google Scholar] [CrossRef]
  36. Liu, H.; Shen, W.; Chen, X. A Room Temperature Operated Ammonia Gas Sensor Based on Ag-Decorated TiO2 Quantum Dot Clusters. RSC Adv. 2019, 9, 24519–24526. [Google Scholar] [CrossRef]
  37. Hu, J.W.; Zou, Y.D.; Deng, Y.; Li, H.-J.; Xu, H.; Wang, D.; Wu, L.M.; Deng, Y.H.; Li, G.S. Recent Advances in Non-Ionic Surfactant Templated Synthesis of Porous Metal Oxide Semiconductors for Gas Sensing Applications. Prog. Mater. Sci. 2025, 150, 101409. [Google Scholar] [CrossRef]
  38. Hou, L.; Duan, J.; Xiong, F.; Carraro, C.; Shi, T.; Maboudian, R.; Long, H. Low Power Gas Sensors: From Structure to Application. ACS Sens. 2024, 9, 6327–6357. [Google Scholar] [CrossRef]
  39. Doan, T.H.P.; Ta, Q.T.H.; Sreedhar, A.; Hang, N.T.; Yang, W.; Noh, J.-S. Highly Deformable Fabric Gas Sensors Integrating Multidimensional Functional Nanostructures. ACS Sens. 2020, 5, 2255–2262. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the synthesis process for TiO2@MoS2 nanosheets via a graphene oxide sacrificial template combined with hydrothermal reaction.
Figure 1. Schematic illustration of the synthesis process for TiO2@MoS2 nanosheets via a graphene oxide sacrificial template combined with hydrothermal reaction.
Chemosensors 13 00154 g001
Figure 2. Morphological and structural characterization of TiO2@MoS2 heterostructures: SEM images of (a) TiO2, (b) MoS2, (c) TiO2@MoS2-4, (d) TiO2@MoS2-8, and (e) TiO2@MoS2-12; (fi) TEM, HR-TEM, SAED, and EDS elemental mapping images of TiO2@MoS2-8, respectively.
Figure 2. Morphological and structural characterization of TiO2@MoS2 heterostructures: SEM images of (a) TiO2, (b) MoS2, (c) TiO2@MoS2-4, (d) TiO2@MoS2-8, and (e) TiO2@MoS2-12; (fi) TEM, HR-TEM, SAED, and EDS elemental mapping images of TiO2@MoS2-8, respectively.
Chemosensors 13 00154 g002
Figure 3. (a) XRD patterns of TiO2, MoS2, TiO2@MoS2-4 (pink line), TiO2@MoS2-8 (orange line), and TiO2@MoS2-12 (green line) samples; (b,c) nitrogen adsorption–desorption isotherms and BJH pore size distribution of TiO2, MoS2, and TiO2@MoS2-8.
Figure 3. (a) XRD patterns of TiO2, MoS2, TiO2@MoS2-4 (pink line), TiO2@MoS2-8 (orange line), and TiO2@MoS2-12 (green line) samples; (b,c) nitrogen adsorption–desorption isotherms and BJH pore size distribution of TiO2, MoS2, and TiO2@MoS2-8.
Chemosensors 13 00154 g003
Figure 4. (a) X-ray photoelectron spectroscopy (XPS) survey spectra of TiO2, MoS2, and TiO2@MoS2-8; (b) high-resolution Ti 2p spectra of TiO2 and TiO2@MoS2-8; (c) Mo 3d spectra of MoS2 and TiO2@MoS2-8; (d) S 2p spectra of MoS2 and TiO2@MoS2-8; (e) O 1s spectra of MoS2 and TiO2@MoS2-8; (f) electron spin resonance (ESR) spectra of TiO2 and TiO2@MoS2-8.
Figure 4. (a) X-ray photoelectron spectroscopy (XPS) survey spectra of TiO2, MoS2, and TiO2@MoS2-8; (b) high-resolution Ti 2p spectra of TiO2 and TiO2@MoS2-8; (c) Mo 3d spectra of MoS2 and TiO2@MoS2-8; (d) S 2p spectra of MoS2 and TiO2@MoS2-8; (e) O 1s spectra of MoS2 and TiO2@MoS2-8; (f) electron spin resonance (ESR) spectra of TiO2 and TiO2@MoS2-8.
Chemosensors 13 00154 g004
Figure 5. (a) Response (Ra/Rg) of MoS2, TiO2@MoS2-4, TiO2@MoS2-8, and TiO2@MoS2-12 to 10 ppm toluene across 200–280 °C; (b) response/recovery time; (c) selectivity toward 10 ppm toluene, xylene, benzene, triethylamine, acetone, methanol, ethanol, and ammonia; (d) gas responses to mixed vapors at 240 °C; (e) dynamic response curves for varied toluene concentrations (0.05–10 ppm) at 240 °C; (f) relationship between sensor response and toluene concentrations (0.05–10 ppm); (g) reproducibility over five cycles at 10 ppm toluene (240 °C); (h) humidity-dependent responses (25 °C, 25–75% RH); (i) long-term stability (30 days, 10 ppm).
Figure 5. (a) Response (Ra/Rg) of MoS2, TiO2@MoS2-4, TiO2@MoS2-8, and TiO2@MoS2-12 to 10 ppm toluene across 200–280 °C; (b) response/recovery time; (c) selectivity toward 10 ppm toluene, xylene, benzene, triethylamine, acetone, methanol, ethanol, and ammonia; (d) gas responses to mixed vapors at 240 °C; (e) dynamic response curves for varied toluene concentrations (0.05–10 ppm) at 240 °C; (f) relationship between sensor response and toluene concentrations (0.05–10 ppm); (g) reproducibility over five cycles at 10 ppm toluene (240 °C); (h) humidity-dependent responses (25 °C, 25–75% RH); (i) long-term stability (30 days, 10 ppm).
Chemosensors 13 00154 g005
Figure 6. (a) In situ DRIFTS analysis of the TiO2@MoS2-8 nanocomposite during gas exposure; (b) band alignment diagrams illustrating the interfacial energy structure of TiO2 and MoS2 pre- and post-heterojunction formation.
Figure 6. (a) In situ DRIFTS analysis of the TiO2@MoS2-8 nanocomposite during gas exposure; (b) band alignment diagrams illustrating the interfacial energy structure of TiO2 and MoS2 pre- and post-heterojunction formation.
Chemosensors 13 00154 g006
Table 1. Comparison of sensor-based toluene gas detection characteristics with different detection materials.
Table 1. Comparison of sensor-based toluene gas detection characteristics with different detection materials.
MaterialsT
(°C)
Concentration
(ppm)
Detection Limit (ppm)Tres/Trec
(s)
Ref.
Pd/PdO-decorated SnO22851000.0550/138[28]
V2C MXene252000.04714/34[29]
SnO2-decorated NiO1002500.01-[30]
MoS2-Fe3O42520558/23[7]
ZnO@Co3O4290100511.2/12.5[16]
NiO/ZnO1003000.12/33[17]
TiO2@MoS2240100.059/16This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, D.; Hu, J.; Xu, H.; Wang, D.; Li, G. Construction of 2D TiO2@MoS2 Heterojunction Nanosheets for Efficient Toluene Gas Detection. Chemosensors 2025, 13, 154. https://doi.org/10.3390/chemosensors13050154

AMA Style

Wang D, Hu J, Xu H, Wang D, Li G. Construction of 2D TiO2@MoS2 Heterojunction Nanosheets for Efficient Toluene Gas Detection. Chemosensors. 2025; 13(5):154. https://doi.org/10.3390/chemosensors13050154

Chicago/Turabian Style

Wang, Dehui, Jinwu Hu, Hui Xu, Ding Wang, and Guisheng Li. 2025. "Construction of 2D TiO2@MoS2 Heterojunction Nanosheets for Efficient Toluene Gas Detection" Chemosensors 13, no. 5: 154. https://doi.org/10.3390/chemosensors13050154

APA Style

Wang, D., Hu, J., Xu, H., Wang, D., & Li, G. (2025). Construction of 2D TiO2@MoS2 Heterojunction Nanosheets for Efficient Toluene Gas Detection. Chemosensors, 13(5), 154. https://doi.org/10.3390/chemosensors13050154

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