Next Article in Journal
Synthesis and Investigation of Antibacterial Activity of Thin Films Based on TiO2-Ag and SiO2-Ag with Potential Applications in Medical Environment
Next Article in Special Issue
High Gas Sensitivity to Nitrogen Dioxide of Nanocomposite ZnO-SnO2 Films Activated by a Surface Electric Field
Previous Article in Journal
Polyethylene Oxide Assisted Fish Collagen-Poly-ε-Caprolactone Nanofiber Membranes by Electrospinning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Three-Dimensional MoS2/Reduced Graphene Oxide Nanosheets/Graphene Quantum Dots Hybrids for High-Performance Room-Temperature NO2 Gas Sensors

1
School of Optoelectronic Science and Engineering & Collaborative Innovation Center of Suzhou Nano Science and Technology, Soochow University, Suzhou 215006, China
2
Key Lab of Advanced Optical Manufacturing Technologies of Jiangsu Province & Key Lab of Modern Optical Technologies of Education Ministry of China, Soochow University, Suzhou 215006, China
3
Key Laboratory of Thin Film and Microfabrication (Ministry of Education), Department of Micro/Nano Electronics, School of Electronic Information and Electrical Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(6), 901; https://doi.org/10.3390/nano12060901
Submission received: 24 January 2022 / Revised: 25 February 2022 / Accepted: 3 March 2022 / Published: 9 March 2022
(This article belongs to the Special Issue Nanomaterials in Gas Sensors)

Abstract

:
This study presents three-dimensional (3D) MoS2/reduced graphene oxide (rGO)/graphene quantum dots (GQDs) hybrids with improved gas sensing performance for NO2 sensors. GQDs were introduced to prevent the agglomeration of nanosheets during mixing of rGO and MoS2. The resultant MoS2/rGO/GQDs hybrids exhibit a well-defined 3D nanostructure, with a firm connection among components. The prepared MoS2/rGO/GQDs-based sensor exhibits a response of 23.2% toward 50 ppm NO2 at room temperature. Furthermore, when exposed to NO2 gas with a concentration as low as 5 ppm, the prepared sensor retains a response of 15.2%. Compared with the MoS2/rGO nanocomposites, the addition of GQDs improves the sensitivity to 21.1% and 23.2% when the sensor is exposed to 30 and 50 ppm NO2 gas, respectively. Additionally, the MoS2/rGO/GQDs-based sensor exhibits outstanding repeatability and gas selectivity. When exposed to certain typical interference gases, the MoS2/rGO/GQDs-based sensor has over 10 times higher sensitivity toward NO2 than the other gases. This study indicates that MoS2/rGO/GQDs hybrids are potential candidates for the development of NO2 sensors with excellent gas sensitivity.

1. Introduction

Nowadays, with industrialization and the continuous development of technology and science, the detection of nitrogen oxides (NOX) has attracted increasing attention [1,2,3]. Due to environmental and health concerns, the development of a high-sensitivity gas sensor that can accurately, reliably, and quickly detect low-concentration NO2 gas is essential for air quality monitoring and protection of human health [4,5].
Up to now, a variety of materials have been used to synthesize NO2 sensors, including metal oxides [6], conducting polymers [7], nanocarbon materials [8], and transitional metal dichalcogenides [9]. Among these, graphene has received widespread attention as a potential gas-sensing material. Being a typical p-type semiconductor material, reduced graphene oxide (rGO) exhibits more structural defects and dangling bonds than pure GO, which offers advantageous conditions for gas adsorption [10,11,12]. However, pure rGO-based sensors generally show poor gas sensitivity toward NO2 gas at room temperature [13]. Therefore, several researchers have tried to combine rGO with other nanomaterials to improve its gas-sensing performance [14,15,16]. It has been shown that the three-dimensional (3D) nanostructure of rGO composite can accelerate electron transport and improve the conductivity of composite materials.
In recent years, few-layer or single-layer two-dimensional transition metal sulfides, including TiS2 [17], WS2 [18], MoSe2 [19], MoS2 [20], and WSe2 [21], have attracted increasing attention from the academic community. Among these, MoS2 has been widely used as a gas-sensitive material in various gas monitoring applications because of its low cost, unique electronic structure, and suitable bandgap. However, the pure MoS2 material exhibits few accessible active sites, poor conductivity, and restacking of aggregations, which hinder electron transport and gas adsorption [22,23]. Introduction of a second component to form binary hybrids is considered an effective route to tackle these issues. Many low-dimensional nanomaterials, including graphene [24], carbon nanotubes [25], carbon dots [26], graphene quantum dots [27], etc., have been used to improve the gas sensing performance of MoS2 nanosheets through the formation of hybrid structures [28,29,30].
GQDs, with a size smaller than 20 nm, possess numerous characteristics in common with graphene, including their boundary effects and unique quantum confinement effects. Hence, they are widely used in biology, materials, chemistry, and other fields [31,32,33]. Several studies have shown that small-size graphene has high conductivity and superior electron transport ability, which contribute to its gas sensitivity [34,35]. Binary hybrids of MoS2 with rGO and GQDs have been proposed to improve the conductivity, increase the number of active sites, and accelerate electron transport. The addition of rGO nanosheets and GQDs can effectively avoid the agglomeration of MoS2, thereby supplying many binding sites for the adsorption of gas molecules.
In this study, alternately stacked 3D structures based on MoS2/rGO/GQDs ternary hybrids are prepared for NO2 gas sensing. The introduction of GQDs can prevent the agglomeration of MoS2 and rGO nanosheets. The rGO nanosheets serve as a channel for carrier transmission and a substrate for the growth of MoS2 nanoflowers. Additionally, the 3D nanostructure of the composite material provides numerous good adsorption sites for NO2 gas. These sites are beneficial for electron transmission and further enhance the gas-sensing properties of the MoS2/rGO/GQDs hybrids. The results reveal that the MoS2/rGO/GQDs-based sensor has high-magnitude response, good selectivity, excellent stability, and quick response toward NO2 at room temperature.

2. Materials and Methods

2.1. Chemical Reagents

Sodium molybdate dihydrate (H4MoNa2O6), thiourea (CH4N2S), hydrochloric acid (HCl), ethanol (C2H6O), sodium hydroxide (NaOH), and polyvinylpyrrolidone (PVP) were obtained from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Benzopyrene was purchased from Tokyo Chemical Industry (Tokyo, Japan). Nitric acid (HNO3, 65–68%) was obtained from Chinasun Specialty Products Co., Ltd. (Changshu, China). None of the above chemical reagents required further purification.

2.2. Fabrication of the GQDs

The GQDs were synthesized using a modified version of the method reported by Wang [36]. Typically, 1 g of benzopyrene was added to 80 mL of HNO3 and stirred for 12 h at 80 °C to be nitrated. The mixed solution was washed by deionized water several times, repeatedly filtered using 0.22 µm microfiltration membranes until the filtrate became colorless, and freeze-dried at −52 °C for 12 h. Then, 1.5 g of the resulting yellowish powder was dispersed into 300 mL of NaOH solution (2 mol/L) via ultrasonication for 3 h and stirred for 5 h. The resulting liquid was poured into a reactor lined with polytetrafluoroethylene and kept in an oven for 8 h at 200 °C. After cooling down to room temperature, the obtained solution was filtered through a 0.22 μm microporous membrane with deionized water. In order to remove the unfused small molecules and sodium salts, the filtrate was subjected to a dialysis treatment for two days in a dialysis bag, and it was freeze-dried to obtain a dark brown GQDs powder. Finally, the obtained GQDs powder was dispersed into deionized water to form a GQDs solution of 1 mg/mL for later use.

2.3. Fabrication of the MoS2 Nanoflowers

The MoS2 nanoflowers were prepared via the hydrothermal process. As in a typical procedure, 1.21 g of sodium molybdate dihydrate and 2.36 g of thiourea were weighed into a beaker and prepared into a 40 mL solution. After 3 h stirring, the mixture was poured into an autoclave lined with polytetrafluoroethylene and placed in an oven for 12 h at 200 °C. When the solution cooled down to room temperature, the yellow supernatant was removed with a dropper. The bottom sediment was washed several times with ethanol and deionized water and collected by centrifugation. The resultant powder was transferred to a vacuum drying oven and dried for 8 h at 60 °C to obtain black MoS2 powder.

2.4. Fabrication of the MoS2/rGO Nanocomposites

GO was synthesized using an improved version of Hummers’ method [37]. To synthesize MoS2/rGO nanocomposites, 75 mg of MoS2 powder was dispersed into 25 mL of deionized water and continuously stirred for 1 h. Then, 15 mL of about 1 mg/mL GO solution was added. After 8 h of stirring, the mixture was placed in a reaction kettle and kept in an oven for 12 h at 200 °C. Subsequently, after cooling down to room temperature, the supernatant liquid was removed by a dropper. The product was then centrifuged at 10,000 rpm for 5 min and washed three times with deionized water. After repeated centrifugation and washing cycles, the black precipitate was placed in a vacuum drying oven and then dried for 8 h at 80 °C to obtain MoS2/rGO nanocomposites.

2.5. Fabrication of the MoS2/rGO/GQDs Hybrids

The MoS2 powder, GO, and GQDs dispersion obtained through the above steps were used to prepare the MoS2/rGO/GQDs hybrids. Briefly, 15 mL of the GQDs dispersion (1 mg/mL) was dispersed into the obtained MoS2/rGO nanocomposites. The suspension was then sonicated for 3 h and stirred for 5 h to achieve a homogeneous solution. The obtained solution was placed into a polytetrafluoroethylene-lined reactor and heated for 12 h at 200 °C. Subsequently, the reactor was cooled down to room temperature. The supernatant liquid was removed using a straw, and the remaining solution was centrifuged for 10 min at 10,000 rpm. The precipitate was washed several times with ethanol and deionized water and separated by centrifugation. Then, it was dispersed into an appropriate amount of deionized water with a concentration of 1 mg/mL. The resulting thick solution was placed in a refrigerator to be frozen and then freeze-dried for 16 h to prepare the foamy MoS2/rGO/GQDs hybrids. The obtained MoS2, rGO, and GQDs were admixed at a mass ratio of 50:10:10 at 200 °C for 12 h in the hydrothermal process. The obtained MoS2/rGO/GQDs hybrids are here referred to as MoS2/rGO/GQDs-1. For comparison, hybrids with mass ratios of 50:5:5, 50:3:3, and 50:2:2 were also synthesized and labeled as MoS2/rGO/GQDs-2, MoS2/rGO/GQDs-3, and MoS2/rGO/GQDs-4, respectively.

2.6. Characterization

The morphology and structure were examined via atomic force microscopy (AFM, Dimension Icon, Bruker, Billerica, MA, USA) and scanning electron microscopy (SEM, Sigma300, Carl Zeiss, Oberkochen, Germany). A laser microscope confocal Raman spectrometer (λ = 514 nm, HR800, HORIBA Jobin Yvon, France) was used to acquire the Raman spectra. An X-ray diffractometer (XRD, XPert-Pro MPD, Panalytical, Holland) was used to analyze the crystal structures. The atomic valence and molecular structure of the samples were determined via X-ray photoelectron spectroscopy (XPS, ESCALAB250XI, Thermo Fisher Scientific, Waltham, MA, USA).

2.7. Fabrication and Measurement of the Synthesized Sensors

The MoS2/rGO and MoS2/rGO/GQDs hybrids were used as the gas-sensitive materials for the synthesis of NO2 sensors. Traditional microfabrication procedures were adopted. First, the silicon wafer was hydrophilized, dried, and spin-coated with photoresist. It was then exposed and developed with a mask. After Au sputtering and degumming, interdigitated electrode fingers were obtained. Figure 1 shows that the resultant electrode is about 720 μm long and 600 μm wide. An appropriate amount of the previously synthesized MoS2/rGO/GQDs solution was dropped onto the interdigitated electrode using a microsyringe and dried for the subsequent NO2 gas-sensitivity test.
The gas-sensitivity measurement was carried out through a high-precision semiconductor tester (Agilent 4156C, Santa Clara, CA 95051, USA). During the measurement, the voltage was set to 5 mV, and the current change was recorded by the sensing device in real time. NO2 gas was used as the test gas, and different concentrations of NO2 gas were obtained by regulating the flow ratio between the background gas and NO2 gas. At the beginning of the test, the background gas was introduced for 100 s to maintain the output current in a stable range, and then, the background gas together with NO2 gas were introduced at a certain proportion. R0 refers to the initial resistance value of the measured sensor under the background gas, while R denotes the real-time resistance value of the measured sensor exposed to NO2 gas. The sensitivity can then be defined as S = (RR0)/R0 × 100%.

3. Results and Discussion

3.1. Nanocomposite Material Characterization

AFM measurements were conducted to characterize the morphology of the prepared GQDs. The GQDs were more similar to discs rather than spherical objects. Figure 2 shows that the diameters of the GQDs were in a range from 2 to 7 nm, and the average diameter was about 4 nm. Furthermore, the aggregation of the GQDs can also be observed in Figure 2. This aggregation may be due to the weak hydrogen bonds or the noncovalent bond interactions among the oxygenated functional groups that are present in the GQDs [38,39]. The AFM morphology measurement of MoS2/rGO/GQDs hybrids was also performed and an image is provided in the supplementary information (Figure S1) to confirm the existence of GQDs in the MoS2/rGO/GQDs hybrids.
The microstructure of the prepared composites was characterized through SEM. GO exhibits nearly transparent nanosheets with many wrinkles, as shown in Figure 3a [40,41]. It can be seen from Figure 3b that the pure MoS2 nanoflowers formed by the layered nanosheets have noticeable ripples, and their diameter is about 500 nm. It can be clearly observed from Figure 3c that the MoS2 nanoflowers anchored on the surface of rGO nanosheets are not uniformly distributed. Instead, they are stacked together and aggregated into nanospheres with the rGO nanosheets. The morphology of the MoS2/rGO/GQDs hybrids is shown in Figure 3d. It can be observed that the introduction of GQDs greatly improves the homogeneous distribution of rGO and MoS2 nanosheets. Additionally, numerous 3D interconnected foldable nanostructures are present in the MoS2/rGO/GQDs hybrids. The MoS2 nanoflowers and the small GQDs particles in the MoS2/rGO/GQDs hybrids are distributed on the exposed active sites of the rGO nanosheets. The introduction of GQDs provides nucleation sites for the formation of MoS2/rGO nanocomposites and prevents their agglomeration [42]. The morphologies of MoS2/rGO/GQDs hybrids with different ratios of GQDs are given in the supporting information (Figure S2).
As shown in Figure 4, Raman spectroscopy was used to detect the nonpolar vibrations between the same type of atom in the samples. The Raman spectra of GQDs, rGO, MoS2/rGO, and MoS2/rGO/GQDs show two characteristic peaks at ~1350 and ~1580 cm−1 corresponding to the D and G bands of graphene, respectively. The D-band can be attributed to the disorder degree or edge folding degree of graphene, whereas the G-band is due to the first-order scattering of the E2g mode. Usually, the intensity ratio of the D-band and G-band (ID/IG) reveals the extent of graphene reduction [43,44]. As shown in Figure 4, the ID/IG values of GQDs, rGO, MoS2/rGO, and MoS2/rGO/GQDs are 0.975, 1.226, 1.244, and 1.035, respectively. In comparison with MoS2/rGO, it can be clearly observed that the ID/IG value of MoS2/rGO/GQDs decreased from 1.244 to 1.035, implying that some of the defects of rGO were removed during the deposition of GQDs. The decrease in ID/IG ratio proves that GQDs are successfully fabricated onto the MoS2/rGO/GQDs hybrids [45,46]. The peaks of pure MoS2 are located at 377 and 403 cm−1, which correspond to the E 2 g 1 and A1g vibrational modes, respectively [47]. The E 2 g 1 peak corresponds to the Mo–S in-plane vibration of the MoS2 lattice, while the A1g peak is attributed to the Mo–S out-of-plane vibration [48]. In contrast to pure MoS2, the values of the E 2 g 1 and A1g peaks in the MoS2/rGO and MoS2/rGO/GQDs samples are significantly reduced, which confirms that GQDs effectively inhibit the aggregation of MoS2.
The crystallinity and crystal phase of the hybrids were revealed via XRD. Figure 5a shows that the GQDs have two broad diffraction peaks at 15.8° and 22.6°, which correspond to the (001) and (002) crystal planes, respectively [49]. As displayed in Figure 5b, the four diffraction peaks at 13.9°, 33.3°, 39.6° and 58.8° correspond to the (002), (100), (103) and (110) planes, which are in agreement with the standard JCPDS card of 2H–MoS2 (JCPDS No.37-1492) [50]. All the peaks of pure MoS2 are also observed in the XRD patterns of the MoS2/rGO and MoS2/rGO/GQDs samples, which indicates the successful formation of the MoS2 nanoflowers in these samples. In comparison with the sharp diffraction peak of pure bulk MoS2 located at 2θ = 13.9°, the diffraction peaks of MoS2/rGO and MoS2/rGO/GQDs are wider, which may be due to the poorer crystallinity of the obtained samples and the decrease in particle size [51]. Furthermore, a diffraction peak was observed around 22.5° in rGO, MoS2/rGO and MoS2/rGO/GQDs samples corresponding to the (002) plane of rGO, which confirms the presence of rGO and the successful reduction of GO in the composite material [52]. It can also be seen that the diffraction peak of MoS2/rGO nanocomposites located at 2θ = 22.5° shifted slightly to 2θ = 22.9° when GQDs were added to the MoS2/rGO nanocomposites. This shift was due to the higher functionality of GQDs, as the surface groups cause an increase in the lattice parameter of the rGO nanosheets [53]. This result indicates the successful incorporation of GQDs into the MoS2/rGO nanocomposites. It is also important to note that compared to MoS2/rGO nanocomposites, the peak intensity decreases for MoS2/rGO/GQDs hybrids, which verifies that the addition of GQDs can effectively avoid the agglomeration of MoS2 nanoflowers and rGO nanosheets. The XRD results demonstrate that the MoS2/rGO/GQDs hybrids were successfully fabricated.
The MoS2/rGO/GQDs hybrids were analyzed via XPS. Figure 6a shows the spectrum of the MoS2/rGO/GQDs hybrids. It can be clearly seen that the MoS2/rGO/GQDs hybrids contain oxygen, carbon, sulfur, and molybdenum elements. Figure 6b illustrates that the C 1s energy spectrum of the MoS2/rGO/GQDs hybrids can also be deconvolved into five peaks: C=C (284.6 eV), C–C (285.2 eV), C–O (286.8 eV), C=O (288.4 eV), and O–C=O (289.3 eV). In Figure 6c, two characteristic orbital peaks can be observed at 232.0 eV (3d3/2) and 228.9 eV (3d5/2), which are ascribed to the Mo4+ ions of MoS2 [54]. Additionally, the two small Mo6+ peaks at 235.1 eV (3d3/2) and 232.7 eV (3d5/2) confirm the presence of Mo–O bonds, which may be caused by residual MoO42− in the precursor [55]. The presence of S 2s in MoS2 is confirmed by the small peak located at 226 eV [56]. In Figure 6d, the two dominant S 2p1/2 and S 2p3/2 peaks at 162.8 eV and 161.6 eV are attributed to the divalent sulfide ions (S2−) in MoS2 [57]. These XPS results prove the successful synthesis of the MoS2/rGO/GQDs hybrids using the hydrothermal process.
Figure 7 shows the FTIR spectra of GQDs, MoS2, MoS2/rGO, and MoS2/rGO/GQDs. The sharp peak of the GQDs at 1579 cm−1 is related to the C=C stretching vibration [58], suggesting that the GQDs are mainly composed of C=C bonds. The peaks located at 1070, 2987, and 3363 cm−1 are related to the C–O, C–H, and –OH stretching vibrations [59,60], respectively. The peaks corresponding to the C–S and Mo–S bonds of MoS2 are located at 1401 and 662 cm−1. The C=C stretching vibration of the GQDs can also be observed at 1579 cm−1 in the MoS2/rGO/GQD hybrids, which indicates the existence of GQDs in the composite materials. The common peaks at 1401 and 662 cm−1 in the MoS2/rGO and MoS2/rGO/GQD hybrids are related to the C–S and Mo–S bonds, which confirms the presence of MoS2 [61]. It is noteworthy that the characteristic peaks associated with oxygenated groups, such as O–H, C=O, and C–O groups, were not clearly observed in the MoS2/rGO nanocomposites, which verifies that GO was successfully reduced via the hydrothermal approach [62].

3.2. Gas-Sensing Properties

First, the gas-sensing performance was tested by detecting NO2 gas at room temperature. The experiment results show that pure GQD materials have basically no response to NO2 gas. This may be due to the smaller size of the GQDs compared with that of the interdigitated electrode, which makes it difficult to form a stable conductive loop. The response values of the MoS2/rGO and MoS2/rGO/GQDs hybrids exposed to 30 and 50 ppm NO2 gas at room temperature are shown in Figure 8. The MoS2/rGO nanocomposites showed 16.8% and 16.9% response values to 30 and 50 ppm NO2. Compared with the MoS2/rGO nanocomposites, the addition of GQDs improved sensitivity to 21.1% and 23.2% when the sensor was exposed to 30 and 50 ppm NO2 gas, respectively. The GQDs also act as active sites, which can prevent the agglomeration of nanosheets during mixing of rGO and MoS2 and provide numerous reaction sites for NO2 gas adsorption. Consequently, the gas-sensing performance of the hybrids is enhanced [63].
Generally, materials with a large effective surface area can provide more active parts for gas adsorption and interaction, which contributes to the enhanced gas sensing performance. A CV test was carried out to confirm the effective surface area of hybrids with and without GQDs. Values of 2.253 and 1.165 can be calculated from the Randles–Sevcik equation (see details in the supporting information, Figures S3 and S4), indicating a great enhancement of effective surface area was achieved after the addition of GQDs [64,65,66].
As is well known, the nanostructure of gas-sensitive materials is crucial to improving the gas-sensing properties of gas sensors [67,68]. The MoS2/rGO/GQDs hybrids with 3D nanostructures were expected to exhibit higher gas sensitivity. In order to explore the effect of the GQDs content on the gas sensitivity, four types of MoS2/rGO/GQDs hybrids were synthesized by changing the mass ratio of MoS2, rGO, and GQDs from 50:10:10 to 50:2:2. These prepared composites are sequentially labeled MoS2/rGO/GQDs-1, MoS2/rGO/GQDs-2, MoS2/rGO/GQDs-3, and MoS2/rGO/GQDs-4. Figure 9 shows the comparison between the responses of these different gas sensors. The response values toward 5 ppm NO2 within 150 s were 12.1%, 15.2%, 11.3%, and 6.2% for MoS2/rGO/GQDs-1, MoS2/rGO/GQDs-2, MoS2/rGO/GQDs-3, and MoS2/rGO/GQDs-4, respectively. The resistance of all samples decreased remarkably when the sensor was exposed to NO2 gas and rapidly returned to the initial value after stopping the exposure. The MoS2/rGO/GQDs-2-based sensor exhibited better gas sensitivity than the MoS2/rGO/GQDs-1-based sensor, which indicates that the uniform distribution of MoS2 nanoflowers in the hybrids can effectively improve the gas-sensing performance of the composites. The enhanced gas sensitivity can be attributed to the interaction of the MoS2 nanoflowers with the rGO nanosheets, which results in the construction of a 3D network and improves the interconnectivity among MoS2, rGO, and GQDs [69]. However, with a further decrease in the content of GQDs in the hybrids, the gas sensitivity of the MoS2/rGO/GQDs-3 and MoS2/rGO/GQDs-4 hybrids decreased to 11.3% and 6.2%, respectively. Thus, the aggregation and restacking of MoS2 nanoflowers are not conducive to improving gas adsorption. A possible reason for this behavior may be that the aggregation of MoS2 nanoflowers weakens the supporting effect of the rGO nanosheets and reduces the probability of NO2 gas adsorption on the heterogeneous interface between the MoS2 nanoflowers and the rGO nanosheets, thereby causing a decrease in gas sensitivity [70]. Therefore, due to its highest response value, MoS2/rGO/GQDs-2 was selected to further study the gas-sensing performance.
For the purpose of comparison, the response curves towards 5 ppm NO2 of the pristine MoS2/rGO with various mass ratios of MoS2 and rGO are shown in the supplementary information (Figure S5). The response values to 5 ppm NO2 gas at room temperature were 7.2%, 10.4%, 9.5%, and 5.2% for the MoS2/rGO-based sensors with the mass ratio of MoS2 and rGO varying from 50:10 to 50:2, respectively. The response values of pristine MoS2/rGO samples were lower than those of MoS2/rGO/GQDs hybrids, indicating that the gas-sensing properties of the MoS2/rGO/GQDs-based sensors can be effectively facilitated by the addition of GQDs.
Figure 10 displays the dynamic gas-sensitive response curve of the sensor based on the MoS2/rGO/GQDs-2 hybrids to different NO2 gas concentrations at room temperature. As mentioned in previous reports, the gas sensitivity at high NO2 gas concentrations is stronger than that at low concentrations because of adsorption and desorption during the gas sensitivity tests [71,72]. At NO2 concentrations of 50, 30, 10, and 5 ppm, the response values were 23.2%, 21.1%, 19.9%, and 15.2%, respectively. The response values dropped with the decrease in NO2 gas concentration. The response and recovery times remained stable at 150 s. It can be clearly observed that the gas-sensitive response of the MoS2/rGO/GQDs-based sensor decreased sharply when exposed to NO2 and recovered immediately after the exposure was stopped. Compared to the previous results, the MoS2/rGO/GQDs-based sensor exhibited lower operating temperature, higher response, and lower detection limit (Table 1).
To test the stability and repeatability of the MoS2/rGO/GQDs-2-based sensor response, its gas-sensing properties were measured in four consecutive dynamic response processes. Figure 11 displays the stability and repeatability of the response of the MoS2/rGO/GQDs-based sensor exposed to 50 ppm NO2 gas at room temperature. The gas sensitivity, response time, and recovery time of the sensor did not change significantly after four cycles. After three cycles, the gas-sensitive response retained a value of 23.2%, which indicates the excellent repeatability and stability of the MoS2/rGO/GQDs hybrids.
In practical applications, NO2 is not found alone but is often accompanied by many other toxic and harmful gases. Thus, selectivity is another important indicator typically used to evaluate the performance of gas sensors in practical applications [73]. To explore the gas selectivity of the obtained MoS2/rGO/GQDs hybrids, the sensor was used to measure several conventional industrial organic gases, including isopropanol, acetone, formaldehyde, ethyl acetate, trichloromethane, and n-hexane. The saturated vapor in the solvent bottle was diluted with N2 to a concentration of 1%. Even though the concentration of these vapors was greater than that of NO2, the test results revealed that the response value to 50 ppm NO2 gas was more than 10 times the response to other vapors, as displayed in Figure 12. It can be concluded that the MoS2/rGO/GQDs hybrids have an outstanding reactivity to NO2 gas, while the reactivity to other vapors is negligible. Therefore, the experimental results demonstrate that the MoS2/rGO/GQDs-based sensor exhibits high selectivity toward NO2 and can be used in practical applications.

3.3. Gas-Sensing Mechanism

It can be seen from Figure 13 that the generation of the heterojunction and the modification of the GQDs are responsible for the improvement in gas-sensing properties. The ternary combination of MoS2, GQDs, and rGO results in the formation of many nanostructures similar to pn junctions at the interfaces. These special nanostructures are essential to improve the electron transmission efficiency. When the MoS2/rGO/GQDs hybrids were exposed to background gas, oxygen was adsorbed onto the surface of the hybrids because of the strong adsorption properties of rGO and MoS2. Since rGO has a greater work function value (W = 4.7 eV) than MoS2 (W = 4.3 eV), electrons are transferred from the conduction band of MoS2 to the conduction band of rGO until the Fermi level equilibrium is reached [74,75]. In the meantime, the oxygen molecules attached to the surface of the MoS2/rGO/GQDs hybrids are converted into oxygen anions ( O 2 ) after capturing electrons, owing to their high affinity toward electrons. The reaction is the following:
O 2 + e O 2
The generation of oxygen anions ( O 2 ) reduces the concentration of electrons; this is the reason why the resistance of the fabricated MoS2/rGO/GQDs hybrids was relatively high in the background gas. When the MoS2/rGO/GQDs hybrids were exposed to NO2, the NO2 molecules underwent a reaction with the minority carriers (electrons) and the oxygen ions ( O 2 ) [76] adsorbed onto the MoS2/rGO/GQDs hybrids to generate NO 3 ions as follows:
2 NO 2 + O 2 + e 2 NO 3
Due to this reaction, the captured electrons travel back to the conduction band of MoS2 to increase the electron concentration in the hybrids. The above reaction leads to decreased thickness of the charge layer between MoS2 and rGO, thereby decreasing the resistance of the obtained sensor. In addition, the GQDs modification on the surface of the MoS2/rGO heterojunction is also responsible for improving the sensing performance. The GQDs serve as an electron mediator at the interface heterojunction, and supply numerous active sites for the hybrids. The active sites allow chemisorbed oxygen to react with NO2, which further improves the gas sensitivity of the hybrids [77,78].

4. Conclusions

A novel 3D structured sensor based on MoS2/rGO/GQDs hybrids was prepared for detecting NO2 at room temperature. The MoS2/rGO/GQDs hybrids were obtained by anchoring MoS2 nanoflowers and GQDs nanoparticles onto rGO nanosheets. The introduction of the GQDs inhibited the agglomeration of the MoS2/rGO nanocomposites, considerably improved the homogeneous distribution of rGO and MoS2 nanosheets, and provided numerous reaction sites for NO2 gas adsorption. The prepared MoS2/rGO/GQDs-based sensor had a response of 23.2% toward 50 ppm NO2 gas, while it retained a response of 15.2% when exposed to NO2 concentration as low as 5 ppm. Furthermore, it was also found that the MoS2/rGO/GQDs-based sensor exhibited a very low detection limit, high response, excellent stability, outstanding repeatability, excellent selectivity, and quick response/recovery characteristics toward NO2 gas at room temperature. The superior gas-sensing ability was due to the synergistic effects of the 3D nanostructures, heterojunctions, and GQDs in the MoS2/rGO/GQDs hybrids. The proposed MoS2/rGO/GQDs-based sensor exhibits outstanding gas-sensing properties and, thus, has great potential in the detection of NO2 gas.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12060901/s1, Figure S1: AFM image of MoS2/rGO/GQDs; Figure S2: SEM images of (a) MoS2/rGO/GQDs-1, (b) MoS2/rGO/GQDs-2 (c) MoS2/rGO/GQDs-3 and (d) MoS2/rGO/GQDs-4; Figure S3: Cyclic voltammograms of (a) MoS2/rGO nanocomposites and (b) MoS2/rGO/GQDs hybrids in 10 mM [Fe(CN)6]3−/4− and 0.1 M KCl solution at different scan rates from 25 to 300 mV·s−1; Figure S4: Peak currents as a function of scan rate for the determination of the effective surface area; Figure S5: Response and recovery curves of MoS2/rGO-based sensors to 5 ppm NO2.

Author Contributions

Conceptualization, Y.W.; methodology, Y.W., C.Y. and Z.Z.; software, C.Y. and Z.W.; validation, Y.W.; formal analysis, Z.Z. and Z.W.; investigation, C.Y., Z.Z. and Z.W.; resources, Y.W. and N.H.; data curation, C.Y. and Z.W.; writing—original draft preparation, C.Y.; writing—review and editing, Y.W. and N.H.; visualization, C.Y. and Z.W.; supervision, Y.W. and C.P.; project administration, Y.W.; funding acquisition, Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors gratefully acknowledge financial supports from the National Natural Science Foundation of China (Grant No. 61871281 and 51302179), the International Cooperation Project by MOST of China (2018YFE0125800), and the Priority Academic Program Development (PAPD) of Jiangsu Higher Education Institutions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Duy, L.T.; Kim, D.J.; Trung, T.Q.; Dang, V.Q.; Kim, B.Y.; Moon, H.K.; Lee, N.E. High performance three-dimensional chemical sensor platform using reduced graphene oxide formed on high aspect-ratio micro-pillars. Adv. Funct. Mater. 2014, 25, 883–890. [Google Scholar] [CrossRef]
  2. Liu, B.L.; Chen, L.; Liu, G.; Abbas, A.N.; Fathi, M.; Zhou, C.W. High-performance chemical sensing using schottky-contacted chemical vapor deposition grown monolayer MoS2 transistors. ACS Nano 2014, 8, 5304–5314. [Google Scholar] [CrossRef] [PubMed]
  3. Casals, O.; Markiewicz, N.; Fabrega, C.; Gràcia, I.; Cané, C.; Wasisto, H.S.; Waag, A.; Prades, J.D. A parts per billion (ppb) sensor for NO2 with microwatt (μW) power requirements based on micro light plates. ACS Sens. 2019, 4, 822–826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Xian, F.; Zong, B.; Mao, S. Metal-organic framework-based sensors for environmental contaminant sensing. Nano-Micro Lett. 2018, 10, 64. [Google Scholar]
  5. Lustig, W.P.; Mukherjee, S.; Rudd, N.D.; Desai, A.V.; Li, J.; Ghosh, S.K. Metal—Organic frameworks: Functional luminescent and photonic materials for sensing applications. Chem. Soc. Rev. 2017, 46, 3242–3285. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, Z.H.; Men, G.L.; Zhang, R.X.; Gu, F.B.; Han, D.M. Pd loading induced excellent NO2 gas sensing of 3DOM In2O3 at room temperature. Sens. Actuators B Chem. 2018, 263, 218–228. [Google Scholar] [CrossRef]
  7. Liu, C.H.; Tai, H.L.; Zhang, P.; Ye, Z.B.; Su, Y.J.; Jiang, Y.D. Enhanced ammonia-sensing properties of PANI-TiO2-Au ternary self assembly nanocomposite thin film at room temperature. Sens. Actuators B Chem. 2017, 246, 85–95. [Google Scholar] [CrossRef]
  8. Lee, S.W.; Lee, W.; Hong, Y.; Lee, G.; Yoon, D.S. Recent advances in carbon material-based NO2 gas sensors. Sens. Actuators B Chem. 2018, 255, 1788–1804. [Google Scholar] [CrossRef]
  9. Zhang, D.Z.; Sun, Y.E.; Jiang, C.X.; Yao, Y.; Wang, D.Y.; Zhang, Y. Room-temperature highly sensitive CO gas sensor based on Ag-loaded zinc oxide/molybdenum disulfifide ternary nanocomposite and its sensing properties. Sens. Actuators B Chem. 2017, 253, 1120–1128. [Google Scholar] [CrossRef]
  10. Zhang, J.; Wu, J.J.; Wang, X.X.; Zeng, D.W.; Xie, C.S. Enhancing room-temperature NO2 sensing properties via forming heterojunction for NiO-rGO composited with SnO2 nanoplates. Sens. Actuators B Chem. 2017, 243, 1010–1019. [Google Scholar] [CrossRef]
  11. Yin, F.F.; Li, Y.; Yue, W.J.; Gao, S.; Zhang, C.W.; Chen, X.Z. Sn3O4/rGO heterostructure as a material for formaldehyde gas sensor with a wide detecting range and low operating temperature. Sens. Actuators B Chem. 2020, 312, 127954. [Google Scholar] [CrossRef]
  12. Han, J.T.; Kim, B.J.; Kim, B.G.; Kim, J.S.; Jeong, B.H.; Jeong, S.Y.; Jeong, H.J.; Cho, J.H.; Lee, G.W. Enhanced electrical properties of reduced graphene oxide multilayer films by in-situ insertion of a TiO2 layer. ACS Nano 2011, 5, 8884–8891. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, Z.; Wang, J.; Pan, D.; Wang, Y.; Noetzel, R.; Li, H.; Xie, P.; Pei, W.L.; Umar, A.; Jiang, L.; et al. Mimicking a dog’s nose: Scrolling graphene nanosheets. ACS Nano 2018, 12, 2521–2530. [Google Scholar] [CrossRef] [PubMed]
  14. Zhao, B.; Xu, Y.T.; Huang, S.Y.; Zhang, K.; Yuen, M.M.F.; Xu, J.B.; Fu, X.Z.; Sun, R.; Wong, C.P. 3D RGO frameworks wrapped hollow spherical SnO2-Fe2O3 mesoporous nano-shells: Fabrication, characterization and lithium storage properties. Electrochim. Acta 2016, 202, 186–196. [Google Scholar] [CrossRef]
  15. Wang, Z.Y.; Gao, S.; Fei, T.; Liu, S.; Zhang, T. Construction of ZnO/SnO2 heterostructure on reduced graphene oxide for enhanced nitrogen dioxide sensitive performances at room temperature. ACS Sens. 2019, 4, 2048–2057. [Google Scholar] [CrossRef]
  16. Mao, S.; Cui, S.M.; Lu, G.H.; Yu, K.H.; Wen, Z.H.; Chen, J.H. Tuning gas sensing properties of reduced graphene oxide using tin oxide nanocrystals. J. Mater. Chem. 2012, 22, 11009–11013. [Google Scholar] [CrossRef]
  17. Das, T.; Chakraborty, S.; Ahuja, R.; Kawazoe, Y.; Das, G.P. Charge transfer driven interaction of CH4, CO2 and NH3 with TiS2 monolayer: Influence of vacancy defect. Catal. Today 2020, 370, 189–195. [Google Scholar] [CrossRef]
  18. Cao, J.; Wang, W.; Zhou, J.; Chen, J.; Deng, H.; Zhang, Y.; Liu, X. Controllable gas sensitive performance of 1T’ WS2 monolayer instructed by strain: First-principles simulations. Chem. Phys. Lett. 2020, 758, 137921. [Google Scholar] [CrossRef]
  19. Hu, X.Y.; Gui, Y.G.; Liu, Y.J.; Ran, L.; Chen, X.P. Adsorption characteristics of H2S, SO2, SO2F2, SOF2, and N2 on NiO–MoSe2 monolayer for gas-sensing applications. Vacuum 2021, 193, 110506. [Google Scholar] [CrossRef]
  20. Zhao, P.C.; Ni, M.J.; Xu, Y.T.; Wang, C.X.; Chen, C.; Zhang, X.R.; Li, C.Y.; Xie, Y.X.; Fei, J.J. A novel ultrasensitive electrochemical quercetin sensor based on MoS2 carbon nanotube@graphene oxide nanoribbons/HS-cyclodextrin/graphene quantum dots composite fifilm. Sens. Actuators B Chem. 2019, 299, 126997. [Google Scholar] [CrossRef]
  21. Lu, Z.; Zhai, Y.; Liang, Q.Z.; Wu, W. Promoting sensitivity and selectivity of NO2 gas sensor based on metal (Pt, Re, Ta)-doped monolayer WSe2: A DFT study. Chem. Phys. Lett. 2020, 755, 137737. [Google Scholar] [CrossRef]
  22. Zhou, Y.; Liu, G.Q.; Zhu, X.Y.; Guo, Y.C. Ultrasensitive NO2 gas sensing based on rGO/MoS2 nanocomposite fifilm at low temperature. Sens. Actuators B Chem. 2017, 251, 280–290. [Google Scholar] [CrossRef]
  23. Chen, C.Z.; Shen, M.; Li, Y.Z. One pot synthesis of 1T@2H-MoS2/SnS2 heterojunction as a photocatalyst with excellent visible light response due to multiphase synergistic effect. Chem. Phys. 2021, 548, 111230. [Google Scholar] [CrossRef]
  24. Geng, X.; Lu, P.F.; Zhang, C.; Lahem, D.; Olivier, M.G.; Debliquy, M. Room-temperature NO2 gas sensors based on rGO@ZnO1−x composites: Experiments and molecular dynamics simulation. Sens. Actuators B Chem. 2019, 282, 690–702. [Google Scholar] [CrossRef]
  25. Drozdowska, K.; Rehman, A.; Krajewska, A.; Lioubtchenko, D.V.; Paviov, K.; Rumyantsev, S.; Smulko, J.; Cywainski, G. Effects of UV light irradiation on fluctuation enhanced gas sensing by carbon nanotube networks. Sens. Actuators B Chem. 2021, 131069. [Google Scholar] [CrossRef]
  26. Supchocksoonthorn, P.; Hanchaina, R.; Sinoy, M.C.A.; Luna, M.D.G.D.; Kangamasksin, T.; Paoprasert, P. Novel solution- and paper-based sensors based on label-free fluorescent carbon dots for the selective detection of pyrimethanil. Appl. Surf. Sci. 2021, 564, 150372. [Google Scholar] [CrossRef]
  27. Ibrahim, I.; Lim, H.N.; Huang, N.M. In-situ formation of electron acceptor to inhibit charge separation of photo-electrochemical sensor of dopamine-based CdS/Au/GQDs. Electrochim. Acta 2020, 360, 137013. [Google Scholar] [CrossRef]
  28. Lu, S.; Chen, M.Z.; Wang, Y.L.; Li, R.; Lin, J.; Zhang, X.T. Highly efficient MoS2/rGO electrocatalysts for triiodide reduction as Pt-free counter electrode for dye-sensitized solar cells. Sol. Energy 2021, 220, 788–795. [Google Scholar] [CrossRef]
  29. Lv, K.L.; Suo, W.Q.; Shao, M.D.; Zhu, Y.; Wang, X.P.; Feng, J.J.; Fang, M.W. Nitrogen doped MoS2 and nitrogen doped carbon dots composite catalyst forelectroreduction CO2 to CO with high Faradaic efficiency. Nano Energy 2019, 63, 103834. [Google Scholar] [CrossRef]
  30. Wu, H.; Guo, Z.S.; Li, M.; Hu, G.H.; Tang, T.; Wen, J.F.; Li, X.Y.; Huang, H.F. Enhanced pseudocapacitive performance of MoS2 by introduction of both N-GQDs and HCNT for flexible supercapacitors. Electrochim. Acta 2021, 370, 137758. [Google Scholar] [CrossRef]
  31. Wei, L.S.; Cai, J.H.; Li, X.Y.; Wang, X.Y. Fabrication of graphene quantum dots/chitosan composite film and its catalytic reduction for 4-nitrophenol. Ferroelectrics 2019, 548, 124–132. [Google Scholar] [CrossRef]
  32. Shen, J.H.; Zhu, Y.H.; Yang, X.L.; Li, C.Z. Graphene quantum dots: Emergent nanolights for bioimaging, sensors, catalysis and photovoltaic devices. Chem. Commun. 2012, 48, 3686–3699. [Google Scholar] [CrossRef] [PubMed]
  33. Ponomarenko, L.A.; Schedin, F.; Katsnelson, M.I.; Yang, R.; Hill, E.W.; Novoselov, K.S.; Geim, A.K. Chaotic dirac billiard in graphene quantum dots. Science 2008, 320, 356–358. [Google Scholar] [CrossRef] [Green Version]
  34. Wongrat, E.; Nuengnit, T.; Panyathip, R.; Chanlek, N.; Hongsith, N.; Choopun, S. Highly selective room temperature ammonia sensors based on ZnO nanostructures decorated with graphene quantum dots (GQDs). Sens. Actuators B Chem. 2020, 326, 128983. [Google Scholar] [CrossRef]
  35. Chen, Z.L.; Wang, D.; Wang, X.Y.; Yang, J.H. Preparation and formaldehyde sensitive properties of N-GQDs/SnO2 nanocomposite. Chin. Chem. Lett. 2020, 31, 2063–2066. [Google Scholar] [CrossRef]
  36. Wang, L.; Wang, Y.L.; Xu, T.; Liao, H.B.; Yao, C.J.; Liu, Y.; Li, Z.; Chen, Z.W.; Pan, D.Y.; Sun, L.T.; et al. Gram-scale synthesis of single-crystalline graphene quantum dots with superior optical properties. Nat. Commun. 2014, 5, 5357. [Google Scholar] [CrossRef] [Green Version]
  37. Bhangare, B.; Ramgir, N.S.; Pathak, A.; Sinju, K.R.; Debnath, A.K.; Jagtap, S.; Suzuki, N.; Muthe, K.P.; Terashima, C.; Aswal, D.K.; et al. Role of sensitizers in imparting the selective response of SnO2/RGO based nanohybrids towards H2S, NO2 and H2. Mater Sci. Semicond. Process. 2020, 105, 104726. [Google Scholar] [CrossRef]
  38. Li, L.L.; Ji, J.; Fei, R.; Wang, C.Z.; Lu, Q.; Zhang, J.R.; Jiang, L.P.; Zhu, J.J. A facile microwave avenue to electrochemiluminescent two-color graphene quantum dots. Adv. Funct. Mater. 2012, 22, 2971–2979. [Google Scholar] [CrossRef]
  39. Chua, C.K.; Sofer, Z.; Simek, P.; Jankovsky, O.; Klimova, K.; Bakardjieva, S.; Kuckova, S.H.; Pumera, M. Synthesis of strongly fluorescent graphene quantum dots by cage-opening buckminsterfullerene. ACS Nano 2015, 9, 2548–2555. [Google Scholar] [CrossRef]
  40. Zhang, D.Z.; Liu, A.M.; Chang, H.Y.; Xia, B.K. Room-temperature high-performance acetone gas sensor based on hydrothermal synthesized SnO2 reduced graphene oxide hybrid composite. RSC Adv. 2015, 5, 3016–3022. [Google Scholar] [CrossRef]
  41. Zhang, D.Z.; Yin, N.L.; Xia, B.K. Facile fabrication of ZnO nanocrystalline-modifified graphene hybrid nanocomposite toward methane gas sensing application. J. Mater. Sci. Mater. Electron. 2015, 26, 5937–5945. [Google Scholar] [CrossRef]
  42. Sun, Q.H.; Wu, Z.F.; Duan, H.M.; Jia, D.Z. Detection of triacetone triperoxide (TATP) precursors with an array of sensors based on MoS2/RGO composites. Sensors 2019, 19, 1281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Yu, X.L.; Tang, J.; Terabe, K.; Sasaki, T.; Gao, R.S.; Ito, Y.; Nakura, K.; Asano, K.; Suzuki, M. Fabrication of graphene/MoS2 alternately stacked structure for enhanced lithium storage. Mater. Chem. Phys. 2020, 239, 121987. [Google Scholar] [CrossRef]
  44. Ren, H.B.; Gu, C.P.; Joo, S.W.; Cui, J.Y.; Sun, Y.F.; Huang, J.R. Preparation of SnO2 nanorods on reduced graphene oxide and sensing properties of as-grown nanocomposites towards hydrogen at low working temperature. Mater. Express 2018, 8, 263–271. [Google Scholar] [CrossRef]
  45. Lecaros, R.L.G.; Bismonte, M.E.; Doma, B.T., Jr.; Hung, W.S.; Hu, C.C.; Tsai, H.A.; Huang, S.H.; Lee, K.R.; Lai, J.Y. Alcohol dehydration performance of pervaporation composite membranes with reduced graphene oxide and graphene quantum dots homostructured fifiller. Carbon 2020, 162, 318–327. [Google Scholar] [CrossRef]
  46. Chen, Q.; Hu, Y.; Hu, C.G.; Cheng, H.H.; Zhang, Z.P.; Shao, H.B.; Qu, L.T. Graphene quantum dots–three-dimensional graphene composites for high-performance supercapacitors. Phys. Chem. Chem. Phys. 2014, 16, 19307–19313. [Google Scholar] [CrossRef] [PubMed]
  47. Niu, Y.; Wang, R.G.; Jiao, W.C.; Ding, G.M.; Hao, L.F.; Yang, F.; He, X.D. MoS2 graphene fifiber based gas sensing devices. Carbon 2015, 95, 34–41. [Google Scholar] [CrossRef]
  48. Ma, L.; Ye, J.B.; Chen, W.X.; Chen, D.Y.; Lee, J.Y. Gemini surfactant assisted hydrothermal synthesis of nanotile-like MoS2/graphene hybrid with enhanced lithium storage performance. Nano Energy 2014, 10, 144–152. [Google Scholar] [CrossRef]
  49. Li, Y.; Hu, Y.; Zhao, Y.; Shi, G.Q.; Deng, L.; Hou, Y.B.; Qu, L.T. An electrochemical avenue to green-luminescent graphene quantum dots as potential electron-acceptors for photovoltaics. Adv. Mater. 2011, 23, 776–780. [Google Scholar] [CrossRef]
  50. Li, W.Z.; Li, F.; Wang, X.; Tang, Y.; Yang, Y.Y.; Gao, W.B.; Li, R. A facile lyophilization synthesis of MoS2 QDs@graphene as a highly active electrocatalyst for hydrogen evolution reaction. Appl. Surf. Sci. 2017, 401, 190–197. [Google Scholar] [CrossRef]
  51. Zhang, K.; Ye, M.Q.; Han, A.J.; Yang, J.L. Preparation, characterization and microwave absorbing properties of MoS2 and MoS2-reduced graphene oxide (RGO) composites. J. Solid State Chem. 2019, 277, 68–76. [Google Scholar] [CrossRef]
  52. Xu, X.B.; Sun, Y.; Qiao, W.; Zhang, X.; Chen, X.; Song, X.Y.; Wu, L.Q.; Zhong, W.; Du, Y.W. 3D MoS2-graphene hybrid aerogels as catalyst for enhanced efficient hydrogen evolution. Appl. Surf. Sci. 2017, 396, 1520–1527. [Google Scholar] [CrossRef]
  53. Riaz, R.; Ali, M.; Sahito, I.A.; Arbab, A.A.; Maiyalagan, T.; Anjum, A.S.; Ko, M.J.; Jeong, S.H. Self-assembled nitrogen-doped graphene quantum dots (N-GQDs) over graphene sheets for superb electro-photocatalytic activity. Appl. Surf. Sci. 2019, 480, 1035–1046. [Google Scholar] [CrossRef]
  54. Chen, T.D.; Yan, W.H.; Xu, J.G.; Li, J.H.; Zhang, G.P.; Ho, D. Highly sensitive and selective NO2 sensor based on 3D MoS2/rGO composites prepared by a low temperature self-assembly method. J. Alloys Compd. 2019, 793, 541–551. [Google Scholar] [CrossRef]
  55. Long, L.N.; Thi, P.T.; Kien, P.T.; Trung, P.T.; Ohta, M.; Kumabe, Y. Controllable synthesis of MoS2/graphene lowdimensional nanocomposites and their electrical properties. Appl. Surf. Sci. 2019, 504, 144193. [Google Scholar] [CrossRef]
  56. Lkram, M.; Liu, L.J.; Liu, Y.; Ma, L.F.; Lv, H.; Ullah, M.; He, L.; Wu, H.Y.; Wang, R.H.; Shi, K.Y. Fabrication and characterization of a high-surface area MoS2@WS2 heterojunction for the ultrasensitive NO2 detection at room temperature. J. Mater. Chem. A 2019, 7, 14602. [Google Scholar]
  57. Li, W.R.; Xu, H.Y.; Zhai, T.; Yu, H.Q.; Chen, Z.R.; Qiu, Z.W.; Song, X.P.; Wang, J.Q.; Cao, B.Q. Enhanced triethylamine sensing properties by designing Au@SnO2/MoS2 nanostructure directly on alumina tubes. Sens. Actuators B Chem. 2017, 253, 97–107. [Google Scholar] [CrossRef]
  58. Sun, Y.Q.; Wang, S.Q.; Li, C.; Luo, P.H.; Tao, L.; Wei, Y.; Shi, G.Q. Large scale preparation of graphene quantum dots from graphite with tunable fluorescence properties. Phys. Chem. Chem. Phys. 2013, 15, 9907–9913. [Google Scholar] [CrossRef]
  59. Shen, J.H.; Zhu, Y.H.; Yang, X.L.; Zong, J.; Zhang, J.M.; Li, C.Z. One-pot hydrothermal synthesis of graphene quantum dots surface-passivated by polyethylene glycol and their photoelectric conversion under near-infrared light. New J. Chem. 2012, 36, 97–101. [Google Scholar] [CrossRef]
  60. Xie, M.M.; Su, Y.J.; Lu, X.N.; Zhang, Y.Z.; Yang, Z.; Zhang, Y.F. Blue and green photoluminescence graphene quantum dots synthesized from carbon fifibers. Mater. Lett. 2013, 93, 161–164. [Google Scholar] [CrossRef]
  61. Hou, X.H.; Wang, Z.W.; Fan, G.J.; Ji, H.P.; Yi, S.S.; Li, T.; Wang, Y.; Zhang, Z.T.; Yuan, L.; Zhang, R.; et al. Hierarchical three-dimensional MoS2/GO hybrid nanostructures for triethylamine-sensing applications with high sensitivity and selectivity. Sens. Actuators B Chem. 2020, 317, 128236. [Google Scholar] [CrossRef]
  62. Wang, T.; Sun, Z.; Huang, D.; Yang, Z.; Ji, Q.; Hu, N.T.; Yin, G.L.; He, D.N.; Wei, H.; Zhang, Y.F. Studies on NH3 gas sensing by zinc oxide nanowire-reduced graphene oxide nanocomposites. Sens. Actuators B Chem. B 2017, 252, 284–294. [Google Scholar] [CrossRef]
  63. Wang, C.X.; Jin, J.L.; Sun, Y.Y.; Yao, J.R.; Zhao, G.Z.; Liu, Y.Q. In-situ synthesis and ultrasound enhanced adsorption properties of MoS2/graphene quantum dot nanocomposite. Chem. Eng. J. 2017, 327, 774–782. [Google Scholar] [CrossRef]
  64. Huang, D.; Yang, Z.; Li, X.L.; Zhang, L.L.; Hu, J.; Su, Y.J.; Hu, N.T.; Yin, G.L.; He, D.N.; Zhang, Y.F. Three-dimensional conductive networks based on stacked SiO2@graphene frameworks for enhanced gas sensing. Nanoscale 2017, 9, 109–118. [Google Scholar] [CrossRef] [PubMed]
  65. Silamabarasan, K.; Harish, S.; Hara, K.; Archana, J.; Navaneethan, M. Ultrathin layered MoS2 and N-doped graphene quantum dots (N-GQDs) anchored reduced graphene oxide (rGO) nanocomposite-based counter electrode for dye-sensitized solar cells. Carbon 2021, 181, 107–117. [Google Scholar] [CrossRef]
  66. Balamurugan, J.; Peera, S.G.; Guo, M.; Nguyen, T.T.; Kim, N.H.; Lee, J.H. A hierarchical 2D Ni-Mo-S nanosheet@nitrogen doped graphene hybrid as a Pt-free cathode for high-performance dye sensitized solar cells and fuel cells. J. Mater. Chem. A 2017, 5, 17896–17908. [Google Scholar] [CrossRef]
  67. Wang, Z.Y.; Zhang, T.; Han, T.Y.; Fei, T.; Liu, S.; Lu, G.Y. Oxygen vacancy engineering for enhanced sensing performances: A case of SnO2 nanoparticles-reduced graphene oxide hybrids for ultrasensitive ppb-level room-temperature NO2 sensing. Sens. Actuators B Chem. 2018, 266, 812–822. [Google Scholar] [CrossRef]
  68. Gui, Y.H.; Wang, H.Y.; Tian, K.; Yang, L.L.; Guo, H.S.; Zhang, H.Z.; Fang, S.M.; Wang, Y. Enhanced gas sensing properties to NO2 of SnO2/rGO nanocomposites synthesized by microwave-assisted gas-liquid interfacial method. Ceram. Int. 2018, 44, 4900–4907. [Google Scholar] [CrossRef] [Green Version]
  69. Wan, K.C.; Yang, J.L.; Wang, D.; Wang, X.Y. Graphene oxide@3D hierarchical SnO2 nanofifiber/nanosheets nanocomposites for highly sensitive and low-temperature formaldehyde detection. Molecules 2020, 25, 35. [Google Scholar] [CrossRef] [Green Version]
  70. Bhangare, B.; Ramgir, N.S.; Jagtap, S.; Debnath, A.K.; Muthe, K.P.; Terashima, C.; Aswal, D.K.; Gosavi, S.W.; Fujishima, A. XPS and Kelvin probe studies of SnO2/RGO nanohybrids based NO2 sensors. Appl. Surf. Sci. 2019, 487, 918–929. [Google Scholar]
  71. Jiang, L.L.; Tu, S.H.; Yu, H.T.; Meng, Y.M.; Zhao, Y.S.; Hou, X.G. Gas sensitivity of heterojunction TiO2NT/GO materials prepared by a simple method with low-concentration acetone. Ceram. Int. 2020, 46, 5344–5350. [Google Scholar] [CrossRef]
  72. Wang, S.; Wang, P.; Li, Z.; Xiao, C.; Xiao, B.; Zhao, R.; Yang, T.; Zhang, M. Facile fabrication and enhanced gas sensing properties of In2O3 nanoparticles. New J. Chem. 2014, 38, 4879–4884. [Google Scholar] [CrossRef]
  73. Zhang, S.S.; Zhang, B.; Sun, G.; Li, Y.W.; Zhang, B.; Wang, Y.; Gao, J.L.; Zhang, Z.Y. One-step synthesis of Ag/SnO2/rGO nanocomposites and their trimethylamine sensing properties. Mater. Res. Bull. 2019, 114, 61–67. [Google Scholar] [CrossRef]
  74. Gong, Y.X.; Wang, Y.; Sun, G.; Jia, T.K.; Jia, L.; Zhang, F.M.; Lin, L.; Zhang, B.Q.; Cao, J.L.; Zhang, Z.Y. Carbon nitride decorated ball-flflower like Co3O4 hybrid composite: Hydrothermal synthesis and ethanol gas-sensing application. Nanomaterials 2018, 8, 132. [Google Scholar] [CrossRef] [Green Version]
  75. Ye, Z.; Tai, H.; Guo, R.; Yuan, Z.; Liu, C.; Su, Y.; Chen, Z.; Jiang, Y. Excellent ammonia sensing performance of gas sensor based on graphene/titanium dioxide hybrid with improved morphology. Appl. Surf. Sci. 2017, 419, 84–90. [Google Scholar] [CrossRef]
  76. Cho, B.; Yoon, J.; Lim, S.K.; Kim, A.R.; Kim, D.H.; Park, S.G.; Kwon, J.D.; Lee, Y.J.; Lee, K.H.; Lee, B.H.; et al. Chemical sensing of 2D graphene/MoS2 heterostructure device. ACS Appl. Mater. Interface 2015, 7, 16775–16780. [Google Scholar] [CrossRef]
  77. Guo, T.; Wang, L.N.; Sun, S.; Wang, Y.; Chen, X.L.; Zhang, K.N.; Zhang, D.X.; Xue, Z.H.; Zhou, X.B. Layered MoS2@graphene functionalized with nitrogen-doped graphene quantum dots as an enhanced electrochemical hydrogen evolution catalyst. Chin. Chem. Lett. 2019, 30, 1253–1260. [Google Scholar] [CrossRef]
  78. Sangeetha, M.; Madhan, D. Ultra sensitive molybdenum disulfide (MoS2)/graphene based hybrid sensor for the detection of NO2 and formaldehyde gases by fiber optic clad modified method. Opt. Laser Technol. 2020, 127, 106193. [Google Scholar] [CrossRef]
Figure 1. SEM image of the interdigitated electrode.
Figure 1. SEM image of the interdigitated electrode.
Nanomaterials 12 00901 g001
Figure 2. AFM image of GQDs.
Figure 2. AFM image of GQDs.
Nanomaterials 12 00901 g002
Figure 3. SEM images of (a) GO, (b) MoS2, (c) MoS2/rGO and (d) MoS2/rGO/GQDs.
Figure 3. SEM images of (a) GO, (b) MoS2, (c) MoS2/rGO and (d) MoS2/rGO/GQDs.
Nanomaterials 12 00901 g003
Figure 4. Raman spectra of (a) GQDs, (b) rGO, (c) MoS2, (d) MoS2/rGO, and (e) MoS2/rGO/GQDs.
Figure 4. Raman spectra of (a) GQDs, (b) rGO, (c) MoS2, (d) MoS2/rGO, and (e) MoS2/rGO/GQDs.
Nanomaterials 12 00901 g004
Figure 5. XRD patterns of (a) GQDs, (b) rGO, (c) MoS2, (d) MoS2/rGO, and (e) MoS2/rGO/GQDs.
Figure 5. XRD patterns of (a) GQDs, (b) rGO, (c) MoS2, (d) MoS2/rGO, and (e) MoS2/rGO/GQDs.
Nanomaterials 12 00901 g005
Figure 6. XPS spectrum of (a) MoS2/rGO/GQDs, (b) C 1s XPS spectrum of MoS2/rGO/GQDs, (c) Mo 3d XPS spectrum of MoS2/rGO/GQDs and (d) S 2p XPS spectrum of MoS2/rGO/GQDs.
Figure 6. XPS spectrum of (a) MoS2/rGO/GQDs, (b) C 1s XPS spectrum of MoS2/rGO/GQDs, (c) Mo 3d XPS spectrum of MoS2/rGO/GQDs and (d) S 2p XPS spectrum of MoS2/rGO/GQDs.
Nanomaterials 12 00901 g006
Figure 7. FT-IR spectra of GQDs MoS2, MoS2/GQDs and MoS2/rGO/GQDs.
Figure 7. FT-IR spectra of GQDs MoS2, MoS2/GQDs and MoS2/rGO/GQDs.
Nanomaterials 12 00901 g007
Figure 8. Response and recovery curves of MoS2/rGO and MoS2/rGO/GQD-based sensors exposed to 30 and 50 ppm NO2.
Figure 8. Response and recovery curves of MoS2/rGO and MoS2/rGO/GQD-based sensors exposed to 30 and 50 ppm NO2.
Nanomaterials 12 00901 g008
Figure 9. Response and recovery curves of MoS2/rGO/GQDs-based sensors exposed to 5 ppm NO2.
Figure 9. Response and recovery curves of MoS2/rGO/GQDs-based sensors exposed to 5 ppm NO2.
Nanomaterials 12 00901 g009
Figure 10. Response and recovery curves of the MoS2/rGO/GQDs-2 sensor towards different concentrations of NO2.
Figure 10. Response and recovery curves of the MoS2/rGO/GQDs-2 sensor towards different concentrations of NO2.
Nanomaterials 12 00901 g010
Figure 11. Reproducibility of response after exposure to 50 ppm NO2.
Figure 11. Reproducibility of response after exposure to 50 ppm NO2.
Nanomaterials 12 00901 g011
Figure 12. Selectivity toward different kinds of gases.
Figure 12. Selectivity toward different kinds of gases.
Nanomaterials 12 00901 g012
Figure 13. Schematic illustration of the gas sensing mechanism for MoS2/rGO/GQDs composites.
Figure 13. Schematic illustration of the gas sensing mechanism for MoS2/rGO/GQDs composites.
Nanomaterials 12 00901 g013
Table 1. Comparison of main properties and performance characteristics of nanomaterials used to detect NO2 gas.
Table 1. Comparison of main properties and performance characteristics of nanomaterials used to detect NO2 gas.
Material Operating Temperature (°C) Concentration Sensitivity Reference
NiO/SnO2/rGO RT 60 ppm 62.27 [10]
MoS2/rGO 60 °C 2 ppm 59.8% [22]
3D MoS2/rGO 80 °C 1 ppm 2483% [54]
MoS2/WS2 RT 50 ppm 26.12 [56]
SnO2/(0.3%)rGO RT 10 ppm 2.021 [68]
SnO2/rGO 200 °C 4 ppm 4.56 [70]
MoS2/rGO/GQDs RT 50 ppm 23.2% this work
MoS2/rGO/GQDs RT 5 ppm 15.2% this work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, C.; Wang, Y.; Wu, Z.; Zhang, Z.; Hu, N.; Peng, C. Three-Dimensional MoS2/Reduced Graphene Oxide Nanosheets/Graphene Quantum Dots Hybrids for High-Performance Room-Temperature NO2 Gas Sensors. Nanomaterials 2022, 12, 901. https://doi.org/10.3390/nano12060901

AMA Style

Yang C, Wang Y, Wu Z, Zhang Z, Hu N, Peng C. Three-Dimensional MoS2/Reduced Graphene Oxide Nanosheets/Graphene Quantum Dots Hybrids for High-Performance Room-Temperature NO2 Gas Sensors. Nanomaterials. 2022; 12(6):901. https://doi.org/10.3390/nano12060901

Chicago/Turabian Style

Yang, Cheng, Yanyan Wang, Zhekun Wu, Zhanbo Zhang, Nantao Hu, and Changsi Peng. 2022. "Three-Dimensional MoS2/Reduced Graphene Oxide Nanosheets/Graphene Quantum Dots Hybrids for High-Performance Room-Temperature NO2 Gas Sensors" Nanomaterials 12, no. 6: 901. https://doi.org/10.3390/nano12060901

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