Next Article in Journal
Improving Photovoltaic Performance of Hybrid Organic-Inorganic MAGeI3 Perovskite Solar Cells via Numerical Optimization of Carrier Transport Materials (HTLs/ETLs)
Previous Article in Journal
Photoelectrocatalytic Detection and Degradation Studies of a Hazardous Textile Dye Safranin T
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Conductometric Gas Sensor Based on MoO3 Nanoribbon Modified with rGO Nanosheets for Ethylenediamine Detection at Room Temperature

1
Key Laboratory of Functional Inorganic Material Chemistry, School of Chemical Engineering and Material, Heilongjiang University, Ministry of Education, 74 Xuefu Road, Harbin 150080, China
2
School of Chemical Engineering and Material, Heilongjiang University, 74 Xuefu Road, Harbin 150080, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(15), 2220; https://doi.org/10.3390/nano13152220
Submission received: 5 July 2023 / Revised: 26 July 2023 / Accepted: 27 July 2023 / Published: 31 July 2023
(This article belongs to the Topic Advanced Nanomaterials for Sensing Applications)

Abstract

:
An ethylenediamine (EDA) gas sensor based on a composite of MoO3 nanoribbon and reduced graphene oxide (rGO) was fabricated in this work. MoO3 nanoribbon/rGO composites were synthesized using a hydrothermal process. The crystal structure, morphology, and elemental composition of MoO3/rGO were analyzed via XRD, FT-IR, Raman, TEM, SEM, XPS, and EPR characterization. The response value of MoO3/rGO to 100 ppm ethylenediamine was 843.7 at room temperature, 1.9 times higher than that of MoO3 nanoribbons. The MoO3/rGO sensor has a low detection limit (LOD) of 0.235 ppm, short response time (8 s), good selectivity, and long-term stability. The improved gas-sensitive performance of MoO3/rGO composites is mainly due to the excellent electron transport properties of graphene, the generation of heterojunctions, the higher content of oxygen vacancies, and the large specific surface area in the composites. This study presents a new approach to efficiently and selectively detect ethylenediamine vapor with low power.

1. Introduction

Ethylenediamine (C2H8N2, EDA) is an important chemical feedstock widely used in petrochemical and pharmaceutical applications, printing, dyeing, electroplating, and fine chemical intermediates [1,2]. However, EDA is a hazardous chemical that is volatile, corrosive, and flammable, causing environmental pollution and representing a serious threat to human health [3]. EDA vapors can invade the body through the respiratory system and skin, thus causing serious health problems such as conjunctivitis, pneumonia, contact dermatitis, asthma, liver and kidney dysfunction, and even tumors [4,5]. The permissible concentration-short term exposure limit (PC-STEL) for EDA is 10 ppm [6]. Therefore, achieving real-time, effective detection and monitoring of EDA is essential to production life safety. Currently, most methods for detecting EDA rely on expensive and complex instruments such as mass spectrometry, gas chromatography, liquid chromatography, and fluorescence probes [7,8]. Therefore, there is a critical challenge in developing economical, portable, real-time, and selective EDA inspection devices that offer room-temperature operation.
Metal-oxide-based gas sensors (MOS) have attracted increasing attention due to their low cost, small size, high sensitivity, and convenience [9,10]. Orthorhombic MoO3 is an n-type semiconductor extensively applied in lithium-ion batteries, catalysts, supercapacitors, photodetectors, and gas sensors due to its unique layer structure, stable physicochemical properties, high dielectric constant, and good catalytic properties [11,12]. In particular, the layered structure of orthorhombic MoO3 is suitable for gas sensors due to its easy gas diffusion [13]. Morphology modulation is an effective means of improving the gas-sensitive properties of materials for practical applications [14]. One-dimensional (1D) metal oxides are preferred by researchers due to their high surface activity, large specific surface area, and axial electron transport [15]. Mo et al. [16] synthesized α-MoO3 nanobelts via the hydrothermal method and recorded a response value of 174 to 800 ppm ethanol at 300 °C. Bai et al. [15] reported rod-shaped α-MoO3 sensing materials, and the response value of the α-MoO3 materials was 542 to 500 ppm NO2 at 290 °C. Self-assembly α-MoO3 nanobelts synthesized by Zhang et al. [17] showed a higher response value to 100 ppm H2S, which was 223 at 176 °C. Despite the high response sensitivity of the 1D MoO3 sensing materials described above, their high operating temperatures and resulting temperature drift affect measurement accuracy, limiting their application for combustible and explosive gases [18].
Reduced graphene oxide (rGO) with a large specific surface area, extremely high electron mobility, and low electrical noise is considered a promising candidate for gas detection at room temperature [19,20]. Zhang et al. [21] studied the gas sensitivity performance of γ-Fe2O3/rGO composites with H2S at room temperature, and the response value of γ-Fe2O3/rGO to 97 ppm H2S was 520.7. He et al. [22] explored the use of γ-Bi2MoO6/rGO composites for NO2 detection, and the response value of the γ-Bi2MoO6/rGO composite was 6.4 times higher than that of γ-Bi2MoO6. The heterojunction between MOS and rGO in the composites can provide a large number of charge carrier transport channels and create more active sites [23]. Moreover, the incorporation of rGO can prevent the agglomeration of MOS nanomaterials and thus increase the total specific surface area [24].
In this paper, MoO3/rGO composites were synthesized using a hydrothermal process. The gas-sensitive performance of pure MoO3 and MoO3/rGO composites toward EDA at room temperature was investigated. Compared to pure MoO3 nanoribbons, the gas-sensitive properties of MoO3/rGO composites significantly improved. The enhanced gas-sensitive performance mechanism of MoO3/rGO composites is also discussed in detail.

2. Materials and Methods

2.1. Materials

Ammonium heptamolybdate tetrahydrate(AHM, (NH4)6Mo7O24.4H2O, ≥99.0%) and graphite powder were purchased from Shanghai Maclean Biochemical Co. Nitric acid (HNO3, 65–68%), concentrated sulfuric acid (H2SO4, 95–98%), hydrogen peroxide (H2O2, 30%), sodium nitrate (NaNO3, ≥99.0%), sodium sulfate (Na2SO4, ≥99.0%), and potassium permanganate (KMnO4, 99.5%) were purchased from Liaoning Quanrui Reagent Co. Graphene oxide (GO) was synthesized using a modified Hummers method [25].

2.2. Preparation of MoO3/rGO

In the preparation of MoO3/rGO nanocomposites, GO (7 mg) was ultrasonically dispersed in 10 mL of deionized water (DIW) for 2 h to obtain the GO solution. A quantity of 0.618 g of AHM was dissolved in 15 mL of DIW and then added to the above GO solution with continuous stirring for 0.5 h. Subsequently, 2.5 mL of HNO3 was added into the mixed solution under stirring conditions and maintained for 0.5 h. The solution was transferred to a 50 mL stainless steel reactor and heated at 180 °C for 20 h. The precipitate was washed several times with DIW and ethanol and then dried overnight at 70 °C in a vacuum oven. Pure MoO3 nanoribbons were synthesized without the addition of GO using the procedure described above.

2.3. Characterization

The phase structure of the products was determined with an X-ray diffractometer (XRD, Bruker D8A25, CuK, λ = 1.5406 Å). The Fourier-transform infrared spectra (FT-IR) of the samples were obtained via FT-IR spectroscopy (PerkinElmer, Waltham, MA, USA). Raman spectra were recorded using a Raman spectrometer (Horiba, Labram HR800 Evolution, Kyoto, Japan) The morphologies and microstructures of the products were characterized with a cold-field emission scanning electron microscope (FESEM, Hitachi S4800, Hitachi, Tokyo, Japan) and transmission electron microscopy (TEM, Tecnai G2 20, FEI, Billerica, MA, USA). The N2 isotherms for adsorption and desorption were determined using the Brunauer–Emmett Teller method (BET, TriStar II 3020, TriStar, Culver City, CA, USA), and the pore size was determined with the Barrett–Joyner–Halenda technique (BJH). The surface chemical element information of the products was investigated via X-ray photoelectron spectroscopy (XPS, Shimadzu Corporation, Kyoto, Japan) with monochromatic radiation of Al Kα (1486.4 eV), and the corrected standard carbon peak was C1s (284.6 eV). The oxygen vacancy content was investigated using electron paramagnetic resonance (EPR, Bruker EMXPLUS, GER, Bruker, Billerica, MA, USA). Experiments with O2 temperature-programmed desorption (O2-TPD) were performed using an automated chemisorption analyzer (Autochem II 2920, Autochem, Repentigny, QC, Canada).

2.4. Sensor Fabrication and Measurement

The as-synthesized MoO3 and MoO3/rGO powders were ultrasonically dispersed in ethanol, which was applied directly onto Au-interdigitated electrodes (spacing: about 50 μm) through the drop-cast method. After drop casting, the prepared devices were heated at 70 °C for 12 h in a vacuum atmosphere to evaporate the water molecules completely, thus generating a dry thin film to bridge the interdigitated electrodes.
The static gas-sensing performance of the obtained products was tested using a WS-30B measurement system (China), and the bias voltage of 5 V, relative humidity (RH) of 20 ± 2%, and room temperature (23 ± 2 °C) were controlled during measurements. The operational process and test principles for the sensors are similar to those in previous studies [26]. S = Ra/Rg or S = Rg/Ra represents the response (S) of the sensor, where Ra and Rg are the resistance value of the sensor in the air and the target gas, respectively. Response/recovery time is defined here as the time required to achieve 90% of the total change in resistance value.

2.5. Electrochemical Measurements

All electrochemical tests were performed in a three-electrode system using a CHI660E electrochemical workstation (Shanghai, China) with 0.1 M sodium sulfate (Na2SO4) as the electrolyte. The saturated Ag/AgCl electrode, Pt electrode, and Glassy carbon electrode (GCE) were used as the reference electrode, counter electrode, and working electrode, respectively. An amount of 5 mg of the sample was dispersed in 2 mL of ethanol and 10 µL of Nafion solution, and then the solution was drop coated on the GCE and dried at room temperature. The prepared electrodes were immersed in the electrolyte for 1 h to ensure that the open-circuit voltage (OCP) was stabilized before starting electrochemical measurements. The electrochemical impedance spectroscopy (EIS) measurement assay settings were as follows: OCP bias 0.34 V, frequency range 0.1–100 kHz, and ac amplitude 5 mV. For the Mott–Schottky (MS) measurements, the increments were 20 mV, the frequency was 1000 Hz, and amplitude was 5 mV.

3. Results and Discussion

3.1. Structural and Morphological Characterization

The XRD spectra (Figure 1a) of MoO3 showed diffraction peaks of 12.8°, 23.2°, 25.8°, 27.4°, 33.7°, 38.0°, 46.3°, 49.2°, and 58.8° corresponding to the orthorhombic phases of MoO3 (020), (110), (040), (021), (111), (060), (210), (002), and (081) crystal planes (JCPDS No. 35-0609), respectively. The intensity of the diffraction peaks in the (020), (040), and (060) crystal planes of MoO3/rGO were found to be clearly higher than those of MoO3, indicating the presence of a layered crystal structure for the anisotropic growth of MoO3 in composites [27,28]. The characteristic diffraction peak of rGO was not found in the XRD pattern of MoO3/rGO composites. This is mainly attributed to the relatively low content and peak intensity of rGO [29]. Figure 1b indicates the FT-IR spectra of pure MoO3 and MoO3/rGO composites. Here, the three absorption peaks in the range of 500–1000 cm−1 for MoO3 at approximately 559, 875, and 998 cm−1 correspond to the stretching vibrations of (Mo3-O), (Mo2-O), and (Mo=O), respectively [11,30]. For MoO3/rGO, two new absorption peaks are visible at about 1231 and 1613 cm−1 and assigned to the C=C and C-O-C stretching vibrations of rGO [31]. The FT-IR test results confirmed the presence of rGO in the composites. Raman spectroscopy is an effective means of characterizing carbon materials with the usual features of G-band and D-band, and that of MoO3 and MoO3/rGO is shown In Figure 1c. The characteristic peaks at 994 and 818 cm−1 are ascribed to the stretching vibrations of the Mo=O bond, and at 665 cm−1, they correspond to the asymmetrical stretching vibrations of the Mo2-O bond [32]. The peaks in the range of 100–400 cm−1 are related to the various bending modes of α-MoO3 crystals [33]. Besides the peaks of MoO3, the D (1350 cm−1) and G (1598 cm−1) characteristic peaks of rGO also exist in MoO3/rGO. The D band is assigned to the breathing mode of sp3-hybridized carbon, structural defects, and amorphous carbon, whereas the G band corresponds to the scattering mode of sp2 carbon [34]. Raman spectroscopy further confirmed the successful preparation of MoO3/rGO composites.
SEM images of MoO3 and MoO3/rGO composites are illustrated in Figure 2a,b. It can be found that the MoO3 in both materials consists of nanoribbons. Figure 2c presents the TEM images of MoO3/rGO, which clearly show that the folded rGO nanosheets are closely connected to the MoO3 nanoribbons. This result further demonstrates the successful fabrication of MoO3/rGO composites. Figure 2d provides the HRTEM image of MoO3/rGO. Spacing in lattice stripes of 0.198 nm and 0.364 nm was discovered in the HRTEM image, attributed to the (200) and (040) crystal planes of MoO3 (JCPDS No. 35-0609), respectively. The selected area electron diffraction (SAED) pattern of the MoO3/rGO nanocomposite is displayed in the inset of Figure 2d. Here, the diffraction spots along the orthogonal MoO3 [010] zone axes correspond to the diffraction at the (200) and (002) crystal planes [35,36]. Together, the TEM and SAED suggest that orthorhombic MoO3 nanoribbons grow mainly along the [001] direction [37].
The full XPS spectra (Figure 3a) of MoO3 and MoO3/rGO indicate the presence of Mo, O, and C elements in both materials. The C 1s peak at 284.8 eV in the XPS full spectrum of pure MoO3 was caused by the C contamination of the analyzer. Figure 3b presents the C 1s high-resolution spectra of MoO3/rGO. The fitted peaks at 284.6, 286.1, and 288.7 eV were ascribed to the C-C, C-O, and C=O groups [18,31]. Figure 3c shows high-resolution Mo 3d XPS spectra of pure MoO3 and MoO3/rGO. The peaks at 233.1 and 236.2 eV are assigned to the binding energies of Mo 3d3/2 and Mo 3d5/2 orbital electrons of Mo6+ [35]. Moreover, the binding energies at about 232.1 eV (Mo 3d5/2) and 235.2 eV (Mo 3d3/2) correspond to Mo5+ [35]. Table 1 presents the relative content of Mo5+ and Mo6+ in the two materials. The relative content of Mo5+ in the composites increased from 4.6% to 9.7% compared to pure MoO3. This result demonstrates the higher content of oxygen vacancies in the composites [38]. Figure 3d illustrates the high-resolution spectra of O 1s. The peaks of binding energy around 530.4, 531.4, and 532.5 eV are ascribed to the lattice oxygen (OL), oxygen vacancies (OV), and chemisorbed oxygen (OC) of the samples, respectively [39]. The content of oxygen species in MoO3 and MoO3/rGO is listed in Table 1. Evidently, the OC and OV content in the composites is higher than that of MoO3.
XPS analysis indicated that the complexes contained more oxygen vacancies and chemisorbed oxygen. Both materials were examined via EPR and O2-TPD to further examine the content of oxygen vacancies and chemisorbed oxygen. Figure 4a provides the EPR spectra of MoO3 and MoO3/rGO. Both samples display a Lorentz line (g = 2.001). Here, the signal intensity of the MoO3/rGO composite is significantly higher than that of pure MoO3. This result means that MoO3/rGO has more oxygen vacancies [40]. Figure 4b presents the O2-TPD curves of MoO3 and MoO3/rGO composites. The resolved peaks at lower temperatures (<100 °C) are attributed to physisorbed oxygen, and the peaks at higher temperatures (250–550 °C) are associated with chemisorbed oxygen [41]. Compared to pure MoO3, the chemisorbed oxygen peak of the MoO3/rGO composite presented a lower resolution temperature and larger peak area. This result illustrates that the chemisorbed oxygen is more active in the composite [42]. The analysis of the EPR and O2-TPD spectra is in agreement with the XPS results. It is well known that the presence of more adsorbed oxygen species corresponds to more gas-sensing properties in sensors [43].
The N2 adsorption–desorption isotherms of MoO3 and MoO3/rGO are depicted in Figure 5. The specific surface areas of MoO3 and MoO3/rGO were 7.7 and 39.3 m2/g, respectively. It is evident that MoO3/rGO has a larger specific surface area. According to the IUPAC classification, the isotherms of both samples can be classified as type IV isotherms with H3-type hysteresis loops, indicating the presence of mesoporous structures [44]. The average pore sizes of MoO3 and MoO3/rGO were 3.46 and 3.39 nm, respectively, when calculated by the BJH method. This result demonstrates that MoO3 and MoO3/rGO are mesoporous materials. In addition, MoO3/rGO (0.047 m3/g) has a larger pore volume than MoO3 (0.037 m3/g). The larger specific surface area and rich pore channels of MoO3/rGO allow for the exposure of many active sites to interact with the target gas [9,45].
Figure 6a,b show the chemical impedance spectra (EIS) and Mott–Schottky plots (MS) of MoO3 and MoO3/rGO. The diameter of the semicircle represents the charge transfer resistance (Rct) at the semiconductor–electrolyte interface. The comparison shows that the Rct of the MoO3/rGO composites is obviously smaller than that of pure MoO3, which suggests that rGO helps to increase the charge migration rate in the composite [46]. The MS test results show that the slope of the curve is positive for both materials. This result illustrates that both MoO3 and MoO3/rGO have the conductive characteristics of n-type semiconductors.

3.2. Gas-Sensing Properties

Figure 7a,d show the response/recovery curves of pure MoO3 and MoO3/rGO for various EDA concentrations at room temperature, respectively. These curves suggest that the response values of both materials are positively correlated with EDA vapor concentration. The adsorption of reductive EDA on the material surface leads to a decrease in resistance, indicating that both materials have properties typical of n-type semiconductors. Figure 7b,e display histograms of the response values at different EDA concentrations with these two gas sensors. The response values of MoO3/rGO were 834.7–1.1 to 100–0.5 ppm EDA, and the response values of MoO3 were 435.1–1.03 to 100–1 ppm EDA. Apparently, the MoO3/rGO gas sensor exhibited better gas sensitivity at the same concentration. Figure 7c,f depict the response values of these two sensors as a function of EDA vapor concentration. The two fitted equations for response values y and EDA vapor concentration x for the MoO3 and MoO3/rGO sensors are expressed as (1) and (2), respectively. The fitted correlation coefficients yielded R2 values of 0.9966 and 0.9991. These results show that both MoO3 and MoO3/rGO have a good functional fit with EDA. The response values of the sensors at low EDA concentrations were linearly fitted to calculate the detection limit (inset of Figure 7c,f). The fitted equations for MoO3 and MoO3/rGO are expressed as (3) and (4), with R2 values of 0.9428 and 0.9989, respectively. The LODs of MoO3 and MoO3/rGO were 0.531 and 0.235 ppm based on the LOD calculation in Equations (5) and (6), respectively [26]. The test results indicate that MoO3/rGO has a wider detection range and higher sensitivity than the pure MoO3 sensor.
y = 0.0492x2 − 0.5241x + 1.0074  R² = 0.9966
y = 0.0821x2 + 0.0569x + 1.2799  R2 = 0.9991
y = 0.0955x + 0.8805  R² = 0.9428
y = 1.2917x + 0.5283  R² = 0.9989
LOD = 3RMS/K
RMS = (Z2/N)1/2
K represents the slope of the fitted curve at low concentrations, and Z represents the standard deviation of the response values.
Response/recovery time and repeatability of gas-sensitive materials to target gas are crucial indexes to assess the sensitivity of gas-sensitive sensors. Therefore, the response/recovery curves of the MoO3 and MoO3/rGO sensors were compared for 100 ppm EDA gas at room temperature (see Figure 8a,b). Response/recovery time for the MoO3 and MoO3/rGO sensors to 100 ppm EDA gas were 18/901 s and 8/357 s, respectively. The MoO3/rGO sensor yielded a shorter response/recovery time than the pure MoO3 sensor. Figure 8c,d show the cyclic response curves for MoO3 and MoO3/rGO to 100 ppm EDA gas at room temperature. The sensor response values do not vary significantly over the five cycles, indicating that both sensors have good cycling stability.
Selectivity and long-term stability are key parameters for measuring sensor performance in practical applications. The results of the selectivity tests for the MoO3 and MoO3/rGO gas sensors are plotted in Figure 9a. Obviously, the gas response values of MoO3 and MoO3/rGO sensors are much higher for 100 ppm EDA gas than for the other interfering gases (100 ppm triethylamine, NH3, ethanol, formaldehyde, and acetone gas). This result demonstrates the excellent selectivity of MoO3 and MoO3/rGO sensors. The long-term stability of MoO3 and MoO3/rGO tests was evaluated once a week for eight weeks, and the results are summarized in Figure 9b. The response values of MoO3 and MoO3/rGO varied less than 5% over time, indicating that the sensors have excellent long-term stability. The response and recovery times of these two sensors during long-term stability tests are shown in the inset of Figure 9b. The response and recovery time of the MoO3 sensor in general increased with the increase in the test period, which is caused by the agglomeration of the MoO3 nanoribbons. The response and recovery times of the MoO3/rGO sensor fluctuated in magnitude, though not significantly, compared to the first week. MoO3/rGO is more stable compared to MoO3 nanoribbons, which is mainly due to the incorporation of rGO, which can effectively prevent the agglomeration of nanoribbons in the composites. Overall, the MoO3/rGO sensor showed high response values, good function matching, fast responses, excellent selectivity, and long-term stability for EDA gas detection with potential for practical applications.
We compared the MoO3/rGO sensor with other sensors used for EDA detection (Table 2). The MoO3/rGO sensor has good gas-sensitive performance for EDA with high response values and low detection limits. The MoO3/rGO sensor fabricated in this work is potentially valuable for industrial applications.

3.3. Gas-Sensing Mechanism

The resistance changes in the gas-sensitive characteristics of metal oxide semiconductors arise from the chemisorption and desorption of gases on the surfaces of materials [22]. The MoO3/rGO composite exhibited an n-type nature in performance tests measuring sensitivity to EDA vapor. Therefore, the resistance changes in MoO3/rGO are caused by variation in the electron concentration in the material [54]. Figure 10 illustrates the sensing mode of MoO3/rGO composites in the air and the EDA vapor. When the MoO3/rGO sensor was exposed to the air atmosphere, O2 molecules in the air were adsorbed onto the composite surface and captured electrons from the material to form adsorbed oxygen species ( O 2 ) (Equations (7) and (8)) [15]. Simultaneously, an electron depletion layer formed on the MoO3 surface, causing a decrease in the charge carrier density of the composite and increasing sensor resistance [29]. When the sensor was exposed to reduced EDA vapor, the EDA molecules adsorbed onto the material’s surface and interacted with the adsorbed oxygen species (Equation (9)) [55]. Meanwhile, the electrons were released from the reaction and back into the material, thereby inducing a decrease in the thickness of the depletion layer and reducing the resistance of the composites [43]. When the sensor returned to the air, O2 molecules were adsorbed back onto the surface of the nanobelts. This caused the electron depletion layer to rebuild and the resistance to return to its initial value.
O2(gas) → O2(ads)
O 2 ( ads ) + e O 2 ( ads )
C 2 H 8 N 2 + 4 O 2 ( ads ) 2 C O 2 + 4 H 2 O + N 2 + 4 e
The excellent gas-sensitive performance of MoO3/rGO with EDA vapor at room temperature can be attributed to three main factors. First, 1D MoO3 features a layered structure formed by the alternating stacking of octahedral MoO6 bilayer planes in the [010] direction, and the [010] crystal planes of MoO3 nanoribbons have higher catalytic activity [27]. This facilitates the adsorption and diffusion of gas molecules, exposes more active sites, and provides a fast transport path for electrons along the axial direction, thus improving gas-sensitive performance [56]. Secondly, rGO nanosheets can prevent the stacking and agglomeration of MoO3 nanobelts, and rGO itself has high electron mobility, which gives the composites a larger specific surface area, more abundant pore channels, and higher electron transport capacity [24]. This not only provides more adsorption sites and effective diffusion pathways for EDA gases but also shortens the response/recovery time, further improving the sensing performance [20]. Third, heterogeneous structures are formed between the two materials when MoO3 is combined with rGO. The work function of the n-type material MoO3 (5.3 eV) [57] is different from that of the p-type material rGO (4.8 eV) [19]. The electrons in rGO are transferred to MoO3 to balance the Fermi energy level (Ef), which causes the energy band to bend and increases the electron concentration of MoO3 in the composite [58]. This process allows for more electrons to be trapped by O2 molecules adsorbed on the MoO3 surface, thereby forming more adsorbed oxygen species and providing more active sites for the material [21,59]. Hence, the gas-sensitive performance is considerably enhanced.

4. Conclusions

In summary, MoO3/rGO composites were fabricated using a hydrothermal method to develop the first MOS-based resistive gas sensor for the detection of EDA gases. At room temperature, the MoO3/rGO composites exhibited higher response values (834.7), shorter response/recovery times (8/357 s), and lower detection limits (0.235 ppm) for EDA compared to pure MoO3 nanobelts. In addition, the MoO3/rGO composites exhibited good selectivity and long-term stability. The outstanding gas-sensitive performance of this sensor mainly contributed to the formation of heterojunctions between MoO3 nanoribbons and rGO alongside the large specific surface area, abundant oxygen vacancies, and good electron transport properties of rGO. This study provides a new direction for the design and application of highly selective and responsive ethylenediamine sensors at room temperature.

Author Contributions

Conceptualization, H.L., C.H. and G.Z.; methodology, C.H. and G.Z.; formal analysis, H.L. and J.L.; funding acquisition, C.H.; investigation, H.L. and J.L.; resources, C.H. and G.Z.; supervision, C.H. and G.Z.; visualization, H.L. and Y.L.; writing—original draft, H.L.; writing—review and editing, G.Z., J.L. and Q.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Opening Foundation of Key Laboratory of Functional Inorganic Material Chemistry (Heilongjiang University), Ministry of Education.

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jia, Y.; Hu, J.-P.; Dang, L.-R.; Yao, H.; Shi, B.; Zhang, Y.-M.; Wei, T.-B.; Lin, Q. Rational Tuning of Binding Properties of Pillar[5]arene-Based Crystalline Material by Synergistic Effect and Its Application for Fluorescent Detection and Adsorption of 1,2-Ethylenediamine. ACS Sustain. Chem. Eng. 2021, 9, 16203–16209. [Google Scholar] [CrossRef]
  2. Li, P.; Yang, D.; Li, H. Luminescence ethylenediamine sensor based on terbium complexes entrapment. Dye. Pigment. 2016, 132, 306–309. [Google Scholar] [CrossRef]
  3. Chuang, P.-M.; Tu, Y.-J.; Wu, J.-Y. A thiadiazole-functionalized Zn(II)-based luminescent coordination polymer with seven-fold interweaved herringbone nets showing solvent-responsive fluorescence properties and discriminative detection of ethylenediamine. Sens. Actuators B Chem. 2022, 366, 131967. [Google Scholar] [CrossRef]
  4. Zhang, X.-Z.; Zhu, W.-J.; Yang, Z.-X.; Feng, Y.; Fan, L.-L.; Gao, G.-G.; Liu, H. Ultrasensitive photochromic and Raman dual response to ethylenediamine gas through polyoxometalate–viologen crystalline hybrid. J. Mater. Chem. C 2022, 10, 15451–15457. [Google Scholar] [CrossRef]
  5. Ke, Y.; Liu, Y.; Zu, B.; Lei, D.; Wang, G.; Li, J.; Ren, W.; Dou, X. Electronic Tuning in Reaction-Based Fluorescent Sensing for Instantaneous and Ultrasensitive Visualization of Ethylenediamine. Angew. Chem. Int. Ed. Engl. 2022, 61, e202203358. [Google Scholar] [CrossRef]
  6. Nizamidin, P.; Yimit, A.; Yan, Y.; Kutilike, B.; Kari, N.; Mamtimin, G. Fast fabrication and gas-sensing characteristics of petal-like Co-MOF membrane optical waveguide. Sens. Actuators B Chem. 2021, 346, 130342. [Google Scholar] [CrossRef]
  7. Ni, J.; Li, M.Y.; Liu, Z.; Zhao, H.; Zhang, J.J.; Liu, S.Q.; Chen, J.; Duan, C.Y.; Chen, L.Y.; Song, X.D. Discrimination of Various Amine Vapors by a Triemissive Metal-Organic Framework Composite via the Combination of a Three-Dimensional Ratiometric Approach and a Confinement-Induced Enhancement Effect. ACS Appl. Mater. Interfaces 2020, 12, 12043–12053. [Google Scholar] [CrossRef]
  8. Huang, Y.; Liu, X.; Wang, Q.; Fu, J.; Zhao, L.; Liu, Z.; Jin, D. Highly responsive ethylenediamine vapor sensor based on a perylenediimide–camphorsulfonic acid complex via ionic self-assembly. J. Mater. Chem. C 2017, 5, 7644–7651. [Google Scholar] [CrossRef]
  9. Li, J.; Yang, M.; Cheng, X.; Zhang, X.; Guo, C.; Xu, Y.; Gao, S.; Major, Z.; Zhao, H.; Huo, L. Fast detection of NO2 by porous SnO2 nanotoast sensor at low temperature. J. Hazard. Mater. 2021, 419, 126414. [Google Scholar] [CrossRef]
  10. Yang, S.; Wang, Z.; Hu, Y.; Luo, X.; Lei, J.; Zhou, D.; Fei, L.; Wang, Y.; Gu, H. Highly Responsive Room-Temperature Hydrogen Sensing of alpha-MoO3 Nanoribbon Membranes. ACS Appl. Mater. Interfaces 2015, 7, 9247–9253. [Google Scholar] [CrossRef]
  11. Sui, L.-L.; Xu, Y.-M.; Zhang, X.-F.; Cheng, X.-L.; Gao, S.; Zhao, H.; Cai, Z.; Huo, L.-H. Construction of three-dimensional flower-like α-MoO3 with hierarchical structure for highly selective triethylamine sensor. Sens. Actuators B Chem. 2015, 208, 406–414. [Google Scholar] [CrossRef]
  12. Fu, H.; Yang, X.; Wu, Z.; He, P.; Xiong, S.; Han, D.; An, X. Gas-sensing performance of In2O3@MoO3 hollow core-shell nanospheres prepared by a two-step hydrothermal method. Sens. Actuators B Chem. 2022, 352, 131007. [Google Scholar] [CrossRef]
  13. Sui, L.; Zhang, X.; Cheng, X.; Wang, P.; Xu, Y.; Gao, S.; Zhao, H.; Huo, L. Au-Loaded Hierarchical MoO3 Hollow Spheres with Enhanced Gas-Sensing Performance for the Detection of BTX (Benzene, Toluene, And Xylene) And the Sensing Mechanism. ACS Appl. Mater. Interfaces 2017, 9, 1661–1670. [Google Scholar] [CrossRef]
  14. Hermawan, A.; Septiani, N.L.W.; Taufik, A.; Yuliarto, B.; Suyatman; Yin, S. Advanced Strategies to Improve Performances of Molybdenum-Based Gas Sensors. Nano-Micro Lett. 2021, 13, 207. [Google Scholar] [CrossRef] [PubMed]
  15. Bai, S.; Chen, S.; Chen, L.; Zhang, K.; Luo, R.; Li, D.; Liu, C.C. Ultrasonic synthesis of MoO3 nanorods and their gas sensing properties. Sens. Actuators B Chem. 2012, 174, 51–58. [Google Scholar] [CrossRef]
  16. Mo, Y.; Tan, Z.; Sun, L.; Lu, Y.; Liu, X. Ethanol-sensing properties of α-MoO3 nanobelts synthesized by hydrothermal method. J. Alloys Compd. 2020, 812, 152166. [Google Scholar] [CrossRef]
  17. Zhang, L.; Liu, Z.; Jin, L.; Zhang, B.; Zhang, H.; Zhu, M.; Yang, W. Self-assembly gridding α-MoO3 nanobelts for highly toxic H2S gas sensors. Sens. Actuators B Chem. 2016, 237, 350–357. [Google Scholar] [CrossRef]
  18. Cao, P.; Cai, Y.; Pawar, D.; Han, S.; Xu, W.; Fang, M.; Liu, X.; Zeng, Y.; Liu, W.; Lu, Y.; et al. Au@ZnO/rGO nanocomposite-based ultra-low detection limit highly sensitive and selective NO2 gas sensor. J. Mater. Chem. C 2022, 10, 4295–4305. [Google Scholar] [CrossRef]
  19. Moon, D.-B.; Bag, A.; Lee, H.-B.; Meeseepong, M.; Lee, D.-H.; Lee, N.-E. A stretchable, room-temperature operable, chemiresistive gas sensor using nanohybrids of reduced graphene oxide and zinc oxide nanorods. Sens. Actuators B Chem. 2021, 345, 130373. [Google Scholar] [CrossRef]
  20. Li, W.; Guo, J.; Cai, L.; Qi, W.; Sun, Y.; Xu, J.-L.; Sun, M.; Zhu, H.; Xiang, L.; Xie, D.; et al. UV light irradiation enhanced gas sensor selectivity of NO2 and SO2 using rGO functionalized with hollow SnO2 nanofibers. Sens. Actuators B Chem. 2019, 290, 443–452. [Google Scholar] [CrossRef]
  21. Zhang, C.; Zhang, S.; Yang, Y.; Yu, H.; Dong, X. Highly sensitive H2S sensors based on metal-organic framework driven γ-Fe2O3 on reduced graphene oxide composites at room temperature. Sens. Actuators B Chem. 2020, 325, 128804. [Google Scholar] [CrossRef]
  22. He, L.; Lv, H.; Ma, L.; Li, W.; Si, J.; Ikram, M.; Ullah, M.; Wu, H.; Wang, R.; Shi, K. Controllable synthesis of intercalated γ-Bi2MoO6/graphene nanosheet composites for high performance NO2 gas sensor at room temperature. Carbon 2020, 157, 22–32. [Google Scholar] [CrossRef]
  23. Singkammo, S.; Wisitsoraat, A.; Sriprachuabwong, C.; Tuantranont, A.; Phanichphant, S.; Liewhiran, C. Electrolytically exfoliated graphene-loaded flame-made Ni-doped SnO2 composite film for acetone sensing. ACS Appl. Mater. Interfaces 2015, 7, 3077–3092. [Google Scholar] [CrossRef]
  24. Bai, H.; Guo, H.; Wang, J.; Dong, Y.; Liu, B.; Xie, Z.; Guo, F.; Chen, D.; Zhang, R.; Zheng, Y. A room-temperature NO2 gas sensor based on CuO nanoflakes modified with rGO nanosheets. Sens. Actuators B Chem. 2021, 337, 129783. [Google Scholar] [CrossRef]
  25. Li, X.; Shen, J.; Li, N.; Ye, M. Fabrication of γ-MnS/rGO composite by facile one-pot solvothermal approach for supercapacitor applications. J. Power Sources 2015, 282, 194–201. [Google Scholar] [CrossRef]
  26. Liu, Y.; Liu, J.; Pan, Q.; Pan, K.; Zhang, G. Metal-organic framework (MOF) derived In2O3 and g-C3N4 composite for superior NOx gas-sensing performance at room temperature. Sens. Actuators B Chem. 2022, 352, 131001. [Google Scholar] [CrossRef]
  27. Mane, A.A.; Suryawanshi, M.P.; Kim, J.H.; Moholkar, A.V. Highly selective and sensitive response of 30.5 % of sprayed molybdenum trioxide (MoO3) nanobelts for nitrogen dioxide (NO2) gas detection. J. Colloid Interface Sci. 2016, 483, 220–231. [Google Scholar] [CrossRef]
  28. Chen, J.; Han, S.; Zhao, H.; Bai, J.; Wang, L.; Sun, G.; Zhang, Z.; Pan, X.; Zhou, J.; Xie, E. Robust wire-based supercapacitors based on hierarchical α-MoO3 nanosheet arrays with well-aligned laminated structure. Chem. Eng. J. 2017, 320, 34–42. [Google Scholar] [CrossRef]
  29. Bai, S.; Sun, L.; Sun, J.; Han, J.; Zhang, K.; Li, Q.; Luo, R.; Li, D.; Chen, A. Pine dendritic bismuth vanadate loaded on reduced graphene oxide for detection of low concentration triethylamine. J. Colloid Interface Sci. 2021, 587, 183–191. [Google Scholar] [CrossRef]
  30. Kawase, M.; Akaike, K.; Aoyama, K.; Ito, Y.; Tamura, M.; Kanai, K. Elucidation of the enhanced photoactivity of melon calcined with MoO3. Appl. Catal. B Environ. 2020, 273, 119068. [Google Scholar] [CrossRef]
  31. Wang, Z.; Han, T.; Fei, T.; Liu, S.; Zhang, T. Investigation of Microstructure Effect on NO2 Sensors Based on SnO2 Nanoparticles/Reduced Graphene Oxide Hybrids. ACS Appl. Mater. Interfaces 2018, 10, 41773–41783. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, L.; Gao, P.; Bao, D.; Wang, Y.; Chen, Y.; Chang, C.; Li, G.; Yang, P. Synthesis of Crystalline/Amorphous Core/Shell MoO3 Composites through a Controlled Dehydration Route and Their Enhanced Ethanol Sensing Properties. Cryst. Growth Des. 2014, 14, 569–575. [Google Scholar] [CrossRef]
  33. Hou, X.; Ma, C.; Ji, H.; Yi, S.; Zhang, L.; Zhang, Z.; Wang, Y.; Yuan, L.; Chen, D.; Zhou, Y. Loading regulation of gold nanoparticles on self-assembled 3D MoO3 hierarchical structure for high triethylamine sensing. Sens. Actuators B Chem. 2023, 393, 134241. [Google Scholar] [CrossRef]
  34. Wu, Z.-S.; Ren, W.; Gao, L.; Zhao, J.; Chen, Z.; Liu, B.; Tang, D.; Yu, B.; Jiang, C.; Cheng, H.-M. Synthesis of Graphene Sheets with High Electrical Conductivity and Good Thermal Stability by Hydrogen Arc Discharge Exfoliation. ACS Nano 2009, 3, 411–417. [Google Scholar] [CrossRef]
  35. Yang, J.; Xiao, X.; Chen, P.; Zhu, K.; Cheng, K.; Ye, K.; Wang, G.; Cao, D.; Yan, J. Creating oxygen-vacancies in MoO3-x nanobelts toward high volumetric energy-density asymmetric supercapacitors with long lifespan. Nano Energy 2019, 58, 455–465. [Google Scholar] [CrossRef]
  36. Zi, X.; Wan, J.; Yang, X.; Tian, W.; Zhang, H.; Wang, Y. Vacancy-rich 1T-MoS2 monolayer confined to MoO3 matrix: An interface-engineered hybrid for efficiently electrocatalytic conversion of nitrogen to ammonia. Appl. Catal. B Environ. 2021, 286, 119870. [Google Scholar] [CrossRef]
  37. Fang, L.; Shu, Y.; Wang, A.; Zhang, T. Green Synthesis and Characterization of Anisotropic Uniform Single-Crystal α-MoO3 Nanostructures. J. Phys. Chem. C 2007, 111, 2401–2408. [Google Scholar] [CrossRef]
  38. Jiang, Y.; Sun, M.; Ni, J.; Li, L. Ultrastable Sodium Storage in MoO3 Nanotube Arrays Enabled by Surface Phosphorylation. ACS Appl. Mater. Interfaces 2019, 11, 37761–37767. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Chen, P.; Wang, Q.; Wang, Q.; Zhu, K.; Ye, K.; Wang, G.; Cao, D.; Yan, J.; Zhang, Q. High-Capacity and Kinetically Accelerated Lithium Storage in MoO3 Enabled by Oxygen Vacancies and Heterostructure. Adv. Energy Mater. 2021, 11, 2101712. [Google Scholar] [CrossRef]
  40. Ma, J.; Fan, H.; Zhang, W.; Sui, J.; Wang, C.; Zhang, M.; Zhao, N.; Kumar Yadav, A.; Wang, W.; Dong, W.; et al. High sensitivity and ultra-low detection limit of chlorine gas sensor based on In2O3 nanosheets by a simple template method. Sens. Actuators B Chem. 2020, 305, 127456. [Google Scholar] [CrossRef]
  41. Zhang, B.; Bao, N.; Wang, T.; Xu, Y.; Dong, Y.; Ni, Y.; Yu, P.; Wei, Q.; Wang, J.; Guo, L.; et al. High-performance room temperature NO2 gas sensor based on visible light irradiated In2O3 nanowires. J. Alloys Compd. 2021, 867, 159076. [Google Scholar] [CrossRef]
  42. Lei, Z.; Cheng, P.; Wang, Y.; Xu, L.; Lv, L.; Li, X.; Sun, S.; Hao, X.; Zhang, Y.; Zhang, Y.; et al. Pt-doped α-Fe2O3 mesoporous microspheres with low-temperature ultra-sensitive properties for gas sensors in diabetes detection. Appl. Surf. Sci. 2023, 607, 154558. [Google Scholar] [CrossRef]
  43. Wang, X.; Gao, Y.; Zhang, Q.; He, X.; Wang, X. Synthesis of MoO3 (1D) @SnO2 (2D) core-shell heterostructures for enhanced ethanol gas sensing performance. Sens. Actuators B Chem. 2023, 382, 133484. [Google Scholar] [CrossRef]
  44. Okai Amu-Darko, J.N.; Hussain, S.; Zhang, X.; Alothman, A.A.; Ouladsmane, M.; Nazir, M.T.; Qiao, G.; Liu, G. Metal-organic frameworks-derived In2O3/ZnO porous hollow nanocages for highly sensitive H2S gas sensor. Chemosphere 2023, 314, 137670. [Google Scholar] [CrossRef] [PubMed]
  45. He, H.; Liu, J.; Liu, H.; Pan, Q.; Zhang, G. The development of high-performance room temperature NOX one-dimensional Na0.23TiO2/TiO2 compound gas sensor. Colloids Surf. A Physicochem. Eng. Asp. 2022, 648, 129444. [Google Scholar] [CrossRef]
  46. Yin, H.; Chen, Z.; Peng, Y.; Xiong, S.; Li, Y.; Yamashita, H.; Li, J. Dual Active Centers Bridged by Oxygen Vacancies of Ruthenium Single-Atom Hybrids Supported on Molybdenum Oxide for Photocatalytic Ammonia Synthesis. Angew. Chem. Int. Ed. 2022, 61, e202114242. [Google Scholar] [CrossRef]
  47. Jin, Y.-J.; Kwak, G. Detection of biogenic amines using a nitrated conjugated polymer. Sens. Actuators B Chem. 2018, 271, 183–188. [Google Scholar] [CrossRef]
  48. Shang, C.; Wang, L.; An, Y.; Chen, P.; Chang, X.; Qi, Y.; Kang, R.; Fang, Y. Langmuir-Blodgett films of perylene bisimide derivatives and fluorescent recognition of diamines. Phys. Chem. Chem. Phys. 2017, 19, 23898–23904. [Google Scholar] [CrossRef]
  49. Ma, J.-X.; Zhou, T.; Ma, T.; Yang, Z.; Yang, J.-H.; Guo, Q.; Liu, W.; Yang, Q.; Liu, W.; Yang, T. Construction of Transition Metal Coordination Polymers with Free Carboxyl Groups and Turn-On Fluorescent Detection for α,β-Diamine. Cryst. Growth Des. 2020, 21, 383–395. [Google Scholar] [CrossRef]
  50. Caraballo, R.M.; Onna, D.; López Abdala, N.; Soler Illia, G.J.A.A.; Hamer, M. Metalloporphyrins into mesoporous photonic crystals: Towards molecularly-tuned photonic sensing devices. Sens. Actuators B Chem. 2020, 309, 127712. [Google Scholar] [CrossRef]
  51. Wang, S.; Liu, J.; Zhao, H.; Guo, Z.; Xing, H.; Gao, Y. Electrically Conductive Coordination Polymer for Highly Selective Chemiresistive Sensing of Volatile Amines. Inorg. Chem. 2017, 57, 541–544. [Google Scholar] [CrossRef] [PubMed]
  52. Chen, Y.Q.; Tian, Y.; Yao, S.L.; Zhang, J.; Feng, R.Y.; Bian, Y.J.; Liu, S.J. Cd(II) -Organic Frameworks Fabricated with a N-Rich Ligand and Flexible Dicarboxylates: Structural Diversity and Multi-Responsive Luminescent Sensing for Toxic Anions and Ethylenediamine. Chem. Asian J. 2019, 14, 4420–4428. [Google Scholar] [CrossRef] [PubMed]
  53. Xu, X.; Yan, B. Wearable Eu@HOF luminescent fabric as a highly selective and sensitive optical synapse sensor for identification of six laboratory volatile compounds by neuromorphic computing. J. Mater. Chem. A 2022, 10, 15427–15437. [Google Scholar] [CrossRef]
  54. Sun, L.; Sun, J.; Han, N.; Liao, D.; Bai, S.; Yang, X.; Luo, R.; Li, D.; Chen, A. rGO decorated W doped BiVO4 novel material for sensing detection of trimethylamine. Sens. Actuators B Chem. 2019, 298, 126749. [Google Scholar] [CrossRef]
  55. Singh, S.K.; Samanta, U.K.; Dhar, A.; Pal, M.; Paul, M.C. Preparation of Bi-doped ZnO thin film over optical fiber and their application as detection of ethylenediamine in an aqueous medium based on the evanescent field technique. Phys. Status Solidi (a) 2020, 217, 2000537. [Google Scholar] [CrossRef]
  56. Li, W.; Ou, Q.; Wang, X.; Xing, K.; Tesfamichael, T.; Motta, N.; Qi, D.-C. Large-sized α-MoO3 layered single crystals for superior NO2 gas sensing. Appl. Surf. Sci. 2022, 586, 152793. [Google Scholar] [CrossRef]
  57. Jin, W.; Wang, H.; Liu, Y.; Yang, S.; Zhou, J.; Chen, W. SnO2 Quantum Dots-Functionalized MoO3 Nanobelts for High-Selectivity Ethylene Sensing. ACS Appl. Nano Mater. 2022, 5, 10485–10494. [Google Scholar] [CrossRef]
  58. Wei, Q.; Sun, J.; Song, P.; Li, J.; Yang, Z.; Wang, Q. MOF-derived α-Fe2O3 porous spindle combined with reduced graphene oxide for improvement of TEA sensing performance. Sens. Actuators B Chem. 2020, 304, 127306. [Google Scholar] [CrossRef]
  59. Mo, R.; Han, D.; Yang, C.; Tang, J.; Wang, F.; Li, C. MOF-derived porous Fe2O3 nanocubes combined with reduced graphene oxide for n-butanol room temperature gas sensing. Sens. Actuators B Chem. 2021, 330, 129326. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns and (b) FT-IR spectra. (c) Raman spectra of MoO3 and MoO3/rGO.
Figure 1. (a) XRD patterns and (b) FT-IR spectra. (c) Raman spectra of MoO3 and MoO3/rGO.
Nanomaterials 13 02220 g001
Figure 2. SEM images of (a) pure (b) MoO3/rGO. TEM (c) and HRTEM (d) images of MoO3/rGO (SAED pattern inset (d)).
Figure 2. SEM images of (a) pure (b) MoO3/rGO. TEM (c) and HRTEM (d) images of MoO3/rGO (SAED pattern inset (d)).
Nanomaterials 13 02220 g002
Figure 3. (a) XPS spectra of MoO3 and MoO3/rGO. (b) C 1s spectra of MoO3/rGO. XPS spectra of (c) Mo 3d and (d) O 1s for MoO3 and MoO3/rGO.
Figure 3. (a) XPS spectra of MoO3 and MoO3/rGO. (b) C 1s spectra of MoO3/rGO. XPS spectra of (c) Mo 3d and (d) O 1s for MoO3 and MoO3/rGO.
Nanomaterials 13 02220 g003
Figure 4. EPR spectra (a) and O2-TPD spectra (b) of MoO3 and MoO3/rGO.
Figure 4. EPR spectra (a) and O2-TPD spectra (b) of MoO3 and MoO3/rGO.
Nanomaterials 13 02220 g004
Figure 5. N2 adsorption–desorption isotherms (inset: pore size distribution) of (a) MoO3 and (b) MoO3/rGO.
Figure 5. N2 adsorption–desorption isotherms (inset: pore size distribution) of (a) MoO3 and (b) MoO3/rGO.
Nanomaterials 13 02220 g005
Figure 6. (a) Nyquist EIS plots and (b) Mott–Schottky plots of MoO3 and MoO3/rGO.
Figure 6. (a) Nyquist EIS plots and (b) Mott–Schottky plots of MoO3 and MoO3/rGO.
Nanomaterials 13 02220 g006
Figure 7. Resistance change curves of (a) MoO3 and (d) MoO3/rGO sensors to various EDA concentrations. Response value histograms of (b) MoO3 and (e) MoO3/rGO sensors in various EDA concentrations. The relationship between the response values of (c) MoO3 and (f) MoO3/rGO sensors with EDA concentrations.
Figure 7. Resistance change curves of (a) MoO3 and (d) MoO3/rGO sensors to various EDA concentrations. Response value histograms of (b) MoO3 and (e) MoO3/rGO sensors in various EDA concentrations. The relationship between the response values of (c) MoO3 and (f) MoO3/rGO sensors with EDA concentrations.
Nanomaterials 13 02220 g007
Figure 8. Response/recovery time of (a) MoO3 and (b) MoO3/rGO gas sensors. Cyclic response curves of (c) MoO3 and (d) MoO3/rGO sensors.
Figure 8. Response/recovery time of (a) MoO3 and (b) MoO3/rGO gas sensors. Cyclic response curves of (c) MoO3 and (d) MoO3/rGO sensors.
Nanomaterials 13 02220 g008
Figure 9. (a) Selectivity of MoO3 and MoO3/rGO sensors to 100 ppm various gases. (b) Long-term stability of MoO3 and MoO3/rGO sensors to 100 ppm EDA gas (response time (1) and recovery time (2) of these two sensors during long-term stability tests inset (b)).
Figure 9. (a) Selectivity of MoO3 and MoO3/rGO sensors to 100 ppm various gases. (b) Long-term stability of MoO3 and MoO3/rGO sensors to 100 ppm EDA gas (response time (1) and recovery time (2) of these two sensors during long-term stability tests inset (b)).
Nanomaterials 13 02220 g009
Figure 10. (a) Schematic diagrams of the gas-sensing mechanism for MoO3/rGO sensors for EDA. (b) Band diagram of the rGO and MoO3 before contacting. Band diagram of the MoO3/rGO composites (c) in air and (d) in EDA gas.
Figure 10. (a) Schematic diagrams of the gas-sensing mechanism for MoO3/rGO sensors for EDA. (b) Band diagram of the rGO and MoO3 before contacting. Band diagram of the MoO3/rGO composites (c) in air and (d) in EDA gas.
Nanomaterials 13 02220 g010
Table 1. The relative content of Mo5+, Mo6+, and oxygen species in MoO3 and MoO3/rGO.
Table 1. The relative content of Mo5+, Mo6+, and oxygen species in MoO3 and MoO3/rGO.
SamplesMo5+ (%)Mo6+ (%)OL (%)OV (%)OC (%)
MoO34.665.490.96.22.9
MoO3/rGO9.790.380.311.97.8
Table 2. Comparison of the EDA-sensing properties of the MoO3/rGO in this work and previous reports.
Table 2. Comparison of the EDA-sensing properties of the MoO3/rGO in this work and previous reports.
MaterialsMethodEDA (ppm)ResponseLOD (ppm)Ref.
Nitrated polythiophenColorimetry50101.99 (A/A0)5.6[47]
Perylene bisimideFluorescence85.21.39 (I0/I)4.0[48]
[Zn4(HIDCPy)4(DMSO)(DMF)3]nFluorescence4501.77 (I/I0 − 1)3.9[49]
MP@MOPOptical800.17 (∆Abs)15[50]
OPTA-MSAFluorescence5080.3 (∆E)0.70[5]
Zn2(bcpBTD)2(bpBTD)(H2O)2]·DMF(1)Fluorescence59.83.8 (I/I0)0.052[3]
[Cd(H2L)2]·3H2O·2DMFResistance90045%/[51]
{[Cd(L)(glu)]·3H2O}∞Fluorescence25003.7 (I0/I)64.5[52]
Eu@IsoMe@Cu/Ni fabricOptical20051.14.74[53]
MoO3/rGOResistance100834.70.235this 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

Liu, H.; Liu, J.; Liu, Q.; Li, Y.; Zhang, G.; He, C. Conductometric Gas Sensor Based on MoO3 Nanoribbon Modified with rGO Nanosheets for Ethylenediamine Detection at Room Temperature. Nanomaterials 2023, 13, 2220. https://doi.org/10.3390/nano13152220

AMA Style

Liu H, Liu J, Liu Q, Li Y, Zhang G, He C. Conductometric Gas Sensor Based on MoO3 Nanoribbon Modified with rGO Nanosheets for Ethylenediamine Detection at Room Temperature. Nanomaterials. 2023; 13(15):2220. https://doi.org/10.3390/nano13152220

Chicago/Turabian Style

Liu, Hongda, Jiongjiang Liu, Qi Liu, Yinghui Li, Guo Zhang, and Chunying He. 2023. "Conductometric Gas Sensor Based on MoO3 Nanoribbon Modified with rGO Nanosheets for Ethylenediamine Detection at Room Temperature" Nanomaterials 13, no. 15: 2220. https://doi.org/10.3390/nano13152220

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