Next Article in Journal
The Functional Fe3O4@SiO2@AuNPs SERS Nanomaterials for Rapid Enrichment and Detection of Mercury Ions in Licorice
Next Article in Special Issue
Recent Advances in Gas Sensing Technology Using Non-Oxide II-VI Semiconductors CdS, CdSe, and CdTe
Previous Article in Journal
High-Performance Sulfur Dioxide Gas Sensor Based on Graphite-Phase Carbon-Nitride-Functionalized Tin Diselenide Nanorods Composite
Previous Article in Special Issue
Materials for Chemical Sensing: A Comprehensive Review on the Recent Advances and Outlook Using Ionic Liquids, Metal–Organic Frameworks (MOFs), and MOF-Based Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Indium Oxide Decorated WS2 Microflakes for Selective Ammonia Sensors at Room Temperature

1
Liaoning Key Lab of Integrated Circuit and Biomedical Electronic System, School of Microelectronics, Dalian University of Technology, Dalian 116024, China
2
School of Biomedical Engineering, Dalian University of Technology, Dalian 116024, China
*
Authors to whom correspondence should be addressed.
Chemosensors 2022, 10(10), 402; https://doi.org/10.3390/chemosensors10100402
Submission received: 8 September 2022 / Revised: 2 October 2022 / Accepted: 4 October 2022 / Published: 8 October 2022
(This article belongs to the Special Issue Gas Sensing beyond MOX Semiconductors)

Abstract

:
Tungsten sulfide decorated with indium oxide nanoparticles (In2O3/WS2) was studied for a chemiresistive-type NH3 sensor at room temperature. It was found that the responses of the developed In2O3/WS2 heterostructure nanocomposite-based sensors are significantly improved to 3.81 from 1.45 for WS2. The response and recovery time of the heterostructure-based sensor was found to significantly decrease to 88 s/116 s (10 ppm) from 112 s/192 s for the WS2-based one. The sensor also exhibits excellent selectivity and signal reproducibility. In comparison to WS2 decorated with both ZnO and SnO2 in similar ways, the In2O3-decorated WS2 has overall better sensing performance in terms of sensitivity, selectivity and response/recovery speeds for NH3 from 1 ppm to 10 ppm at room temperature. The improved sensing properties of WS2 incorporating In2O3 could be attributed to the joint enhancement mechanisms of the “electronic and catalytic” sensitizations.

1. Introduction

Ammonia is a natural, colorless and corrosive gas that is potentially toxic to human health and the environment [1]. It comes from the decomposition of feces and dead animals and plants and exists in the atmosphere at low concentrations [2]. Ammonia is mainly used as a fertilizer to increase crop yields. It is also widely used in refrigeration, fuel, textile production and other industrial fields [3]. As a carbon-free hydrogen carrier, ammonia contains 17.6 wt.% hydrogen. It can be used as a green fuel for internal combustion engines and gas turbines which play an indispensable role in addressing the chemical fuel depletion crisis [4,5]. However, a high level of ammonia exposure can cause damage to human health; therefore, the upper exposure concentration of ammonia within 8 h is limited to 25 ppm [6]. Excessive ammonia emissions can subsequently lead to environmental acidification which needs to be monitored in real time [7,8].
To fulfill the demand for monitoring ammonia, sensors working at room temperature with low cost, low power consumption, high sensitivity, fast response, and good selectivity need to be developed [9]. Metal oxides such as SnO2, ZnO, TiO2, WO3 and In2O3 are widely explored as the sensing materials for detecting ammonia. However, they are usually operated at high temperatures between 100 °C and 550 °C, which requires high power consumption [10,11,12,13,14,15,16]. To fabricate low-temperature chemiresistive-type gas sensors, graphene with a two-dimensional (2D) network of carbon has become a promising candidate material due to its advantages of high specific surface area and excellent electrical conductivity [17]. However, its application is limited by the zero band gap feature, and graphene is not very responsive to ammonia at room temperature. Therefore, graphene was usually functionalized with metal nanoparticles, metal oxides, organic molecules and conductive polymers to enhance its functionality [18].
As analogs to graphene, metal dichalcogenides such as MoS2, SnS2 and WSe2 have also become potential candidate materials for fabricating nanoelectronic devices [19]. Singh et al. synthesized a p-MnS2/n-WO3 heterostructure for a chemical ammonia sensor and found it displayed an improved response to ammonia at 200 °C compared to the pure p-MnS2-based one [20]. Xia et al. reported a WS2/SnO2-based sensor with 2D/0D heterostructure and noticed the response to NO2 could be further improved under UV at room temperature [21]. Wang et al. found that the (2D/2D) rGO/WS2 heterostructure-based ammonia sensor showed improved response/recovery at room temperature. The reasons are attributed to the extra active sites introduced by the secondary 2D phase [22]. Kim et al. fabricated a Au-decorated WS2-SnO2 core–shell nanosheet-based gas sensor and demonstrated that it has enhanced sensing properties for CO at 3.4 V [23].
In previous studies, WS2 has been reported to show promise in detecting harmful gases at room temperature [24]. Metal oxides such as SnO2 and ZnO and noble metals such as Pt and Au have been strategically used to improve the sensing properties in terms of sensitivity, selectivity and response/recovery speeds. In2O3 is also a good semiconducting material for gas sensing because it has a lower resistance and a good response to HCHO [25], NO2 [26] and CO [27]. Therefore, in this work, the 2D WS2 heterostructured by indium oxide nanoparticles was examined. Indium oxide was selected as the secondary phase to improve the sensing properties of a WS2-based chemiresistive-type sensor at room temperature for the following two reasons: We found that the synthesized sensors all displayed higher responses and faster recovery to ammonia than the pure WS2 microflake-based one at room temperature. Comparing the response rates of the developed sensors to the pure sensor, In2O3/WS2 heterostructure nanocomposite-based sensors exhibited a much faster response, The improved performance of the In2O3/WS2-based sensors could be mainly attributed to the electronic sensitization between In2O3 and WS2. Sensors based on WS2 heterostructured with ZnO/SnO2 were also characterized and compared.

2. Experimental Method

2.1. Synthesis of In2O3/WS2 Nanocomposites

The WS2 powders used in this study were purchased from Aladdin Co., Shanghai, China. In2O3 powders were purchased from Macklin Biochemical Co., Ltd., Shanghai, China. ZnO and SnO2 powders were obtained from Aladdin Co., Shanghai, China. All chemicals were of analytical grade and used as received. First, 50.0 mg WS2 powders and 55.5 mg In2O3 powders were added to 15 mL and 20 mL deionized (DI) water, respectively, followed by the sonication in an ultrasound water bath for 15 min. Different amounts of In2O3 solutions (0.1 mL, 0.2 mL, 0.6 mL and 1 mL) were added drop-wise to the WS2 dispersion solution with vigorous stirring for half an hour to obtain the WS2 suspensions with different concentrations of In2O3 (0.5%, 1%, 3% and 5%). WS2 suspensions with different ZnO contents and SnO2 contents (0.5%, 1%, 3% and 5%) for comparisons were obtained by the same method. Then the obtained mixtures were washed with absolute alcohol and DI water three times in sequence and dried at 60 °C overnight in a drying oven.

2.2. Characterization

X-ray diffraction (XRD, D8 Advance, Bruker Corporation, Billerica, MA, USA) was used to investigate the chemical information and crystal structure of indium oxide decorated tungsten sulfide (IO/WS2) heterostructure nanocomposite-based samples. A field emission scanning electron microscope (FESEM, Gemini 300, ZEISS Corporation, Jena, Germany) and a transmission electron microscope (TEM, Tecnai G2 F20, FEI Corporation, Hillsboro, OR, USA) were used to study the surface morphologies of the prepared IO/WS2 samples. X-ray photoelectron spectroscopy (XPS, ESCALAB 250Xi, Thermo Fisher Scientific Corporation, Waltham, MA, USA) was performed to explore the elemental compositions and chemical oxidation states of the IO/WS2 samples.

2.3. Fabrication and Measurement of the Gas Sensors

A printable paste was formed by grinding a small amount of the obtained IO/WS2 materials and DW in an agate mortar. Then the paste was spread on the gold interdigital electrodes pre-printed on the ceramic substrate, which were dried at 60 °C for 12 h in a drying oven to develop the IO/WS2-based chemiresistive-type sensor. To measure the sensing properties of the fabricated sensor, the sensor was placed in a testing chamber with a volume of 50 L, and the resistance of the sensor was recorded using a digital multimeter (Keysight 34465A). A corresponding amount of liquid solutions including ammonia, formaldehyde, toluene, methanol, ethanol and acetone, which were purchased from Kermel Co., Tianjin, China, was vaporized to achieve different gas vapors by heating after being injected into the small crucible at the corner of the testing chamber. The gas vapor concentrations were calculated by the following formula [28]:
C = (22.4 × ρ × d × V1/(M × V2)
where C (ppm) represents the targeted gas concentration, ρ (g/mL) represents the liquid density, d represents the liquid purity, M (g/mol) represents the liquid molecular weight, V1 (L) represents the liquid volume and V2 (L) represents the volume of the testing chamber (50 L). The sensor response (S) was defined by the resistance ratio of Rg/Ra, where Rg is the resistance of the sensor in target ammonia and Ra is the resistance of the sensor in air. The response time and recovery time were measured by the time course of 90% of the maximum response amplitude.

3. Results and Discussion

3.1. Microstructure Characterizations

Figure 1 shows the XRD patterns of the pure WS2 and IO/WS2 with 1% In2O3 (IO/WS2-1) heterostructure nanocomposite-based samples. The main peaks in the XRD pattern of WS2 (red) correspond to 2H-WS2 (PDF # 87-2417). The peaks in the XRD pattern of the IO/WS2-1 heterostructure nanocomposite-based sample contain the diffraction peaks of both WS2 and In2O3 (PDF # 71-2194) without other impurities. The size of In2O3 particles can be calculated by the Scherrer equation as follows [29]:
D = Kλ/(βcosθ)
where D is the size of the particle, λ is the X-ray wavelength (1.54178 Å), β is full width at half maximum (FWHM) of the diffraction peak, θ is the angle of diffraction and K is Scherrer’s constant (K = 0.943). The size of the In2O3 particles estimated from XRD by the Scherrer equation is 61 nm. However, the size of WS2 microflakes cannot be calculated by the Scherrer equation because it does not meet the required condition (D < 100 nm).
As shown in Figure 2a,b, the morphology of WS2 microflakes did not exhibit much change after being incorporated with In2O3. The WS2 microflakes are 0.4–4.3 μm in diameter. The In2O3 nanoparticles show a cube shape with particle size around 30 nm distributed on WS2 microflakes as shown in Figure 2c. Compared to the size calculated by the Scherrer equation, the size of In2O3 particles observed in the TEM was smaller. This may be caused by the strained lattice in the crystals of In2O3. The In2O3 particles and WS2 microflakes were physically mixed without further heating; as a result, the size and morphology of In2O3 particles remained unchanged. Figure 2d shows the high-resolution TEM (HRTEM) image of the IO/WS2-1 heterostructure nanocomposite-based sensor. The lattice spacing of 0.273 nm corresponds to the (100) plane of WS2, and the lattice spacing of 0.292 nm corresponds to the (222) plane of In2O3 nanoparticles, which confirmed the heterostructure formed between In2O3 and WS2. The SEM elemental mapping images of IO/WS2-1 heterostructure nanocomposites are shown in Figure S1 in the Supplementary Materials. The synthesized sample contains four elements: W, S, In and O. The relative concentrations of each element are summarized in Table S1.
Figure 3a shows the survey of the XPS spectra of WS2 microflakes, In2O3 nanoparticles and IO/WS2-1 heterostructure nanocomposites. As shown in Figure 3b, the peaks centered at 531.3 eV and 532.4 eV correspond to the adsorbed oxygen and SO42− of WS2, respectively [30]. The peaks at 531.3 eV, 532.4 eV, 530.1 eV and 533.32 eV shown in Figure 3c correspond to the adsorbed oxygen, SO42−, lattice oxygen and hydroxyl in the IO/WS2-1 heterostructure nanocomposites, respectively [28,31]. The concentrations of the adsorbed oxygen contents estimated from the integrated areas under each peak for the adsorbed O2- on both WS2 and IO/WS2-1 samples were found to increase from 25.57 at.% in the WS2 sample to 54.65 at.% in the IO/WS2-1 sample. However, the amount of SO42− in the IO/WS2-1 sample was reduced, which may be due to the reduction reaction with the hydroxyl groups. The refined core spectra of W 4f and S 2p of both WS2 and IO/WS2-1 samples are shown in Figure S2. The binding energies of both W and S shift to the lower binding energy direction by 0.05 eV after the incorporation of In2O3 nanoparticles. This could be attributed to the electronic equilibrium effect between WS2 and In2O3, in which WS2 obtained the electrons transferred from In2O3. In a comparison of Figure S3 (shown in the Supplementary Materials) and Figure 3d, the refined core spectra of In 3d of In2O3 and IO/WS2-1 samples demonstrate that the binding energy of In 3d shifts to a higher energy by 1.1 eV after the incorporation with WS2 microflakes, indicating the loss of electrons.

3.2. Electrical and Sensing Performance

Figure 4 shows the I–V curves with applied voltages between −5 V and +5 V of the WS2 microflake-based sensor and the IO/WS2 heterostructure nanocomposite-based sensors. It is found that the resistances of the fabricated IO/WS2 heterostructure nanocomposite-based sensors increased with the increase in the In2O3 nanoparticle content. The resistance values of WS2, IO/WS2-0.5, IO/WS2-1, IO/WS2-3 and IO/WS2-5 sensors estimated from Figure 4 are 2.7 M, 3.6 M, 6.6 M, 10.4 M and 31.3 M, respectively. The I–V curves are linear in an ohmic connection between the sensing materials and gold electrodes.
The responses of sensors based on WS2 and IO/WS2 heterostructure nanocomposites with different In2O3 concentrations to 10 ppm ammonia at room temperature are shown in Figure 5a. The IO/WS2 heterostructure nanocomposite-based sensor exhibited higher responses than the pure WS2 microflake-based one. With an increase in the In2O3 content, the response of IO/WS2 heterostructure nanocomposite-based sensors firstly increases as the In2O3 percentage increases from 0.5% to 1% and then decreases above 3% In2O3. It achieves the maximum response of 3.81 with 1% In2O3 content. However, as shown in Figure 5b, the response and recovery time of IO/WS2 heterostructure nanocomposite-based sensors continue to decrease as the concentration of In2O3 increases from 0.5% to 5% with In2O3 introduction.
The gas-sensing mechanism of the IO/WS2 heterostructure nanocomposite-based sensor is illustrated in Figure 6. As shown in Figure 6a, as the WS2 is in contact with In2O3, the electrons will move from the conducting band of In2O3 to the conducting band of WS2 while the holes will move from the valence band of WS2 to the valence band of In2O3 due to the work function of In2O3 (4.28 ev) being lower than that of WS2 (4.9 ev). The transfer of the electrons and holes creates a hole depletion layer (HDL) on the WS2 side and a hole accumulation layer (HAL) on the In2O3 side, which leads to an increase in the resistance on the WS2 and a decrease in the resistance on the In2O3 side. As a p-type semiconductor, the total resistance of the fabricated IO/WS2 heterostructure nanocomposite-based sensor increased for the WS2 determination. When the Fermi levels reach equilibrium, the IO/WS2 heterostructure is formed, as shown in Figure 6b. Figure 6c explains the reason for the improved gas-sensing performance of the IO/WS2 heterostructure nanocomposite-based sensor. As the IO/WS2 heterostructure formed, WS2 obtained the electrons transferring from In2O3, leading to more oxygen molecules adsorbed on WS2 (reaction 4) with more oxygen ions produced. The result is consistent with the XPS result. Reaction 5 is thus promoted. When exposed to ammonia, the generated oxygen ions would react with the ammonia to produce more nitric oxide, water and electrons [32].
The released electrons would return to WS2, leading to the improved sensing response of the IO/WS2 heterostructure nanocomposite-based sensor. However, beyond the optimum concentrations of 1%, the concentration of IO/WS2 heterostructures decreased because the formed IO/WS2 heterostructures were fully covered by the introduced In2O3 particles, leading to a lower sensor response [33]. The involved chemical reaction equations are as follows [30]:
O2 (air) → O2 (ads),
O2 (ads) + e → O2 (ads) (below 100 °C)
4NH3 + 5O2 → 4NO + 6H2O + 5e
The sensing properties of the WS2 microflake-based sensor and the IO/WS2 heterostructure nanocomposite-based sensors in response to different concentrations of ammonia from 1 ppm to 10 ppm at room temperature are displayed in Figure 7a. It can be seen that all the IO/WS2 heterostructure nanocomposite-based sensors have much higher responses than the pure WS2 microflake-based sensor. The IO/WS2-1 sample with 1% In2O3 content exhibited the maximum response to different concentrations of ammonia. The calibration plots of the sensors are shown in Figure 7b. The response of the sensors to NH3 from 1 ppm to 10 ppm indicates a typical tendency as reported in most studies. Figure 7c demonstrates that the fabricated IO/WS2-1 sensor exhibits good repeatability at room temperature. The good long-term stability of the IO/WS2-1 sample is evidenced in Figure 7d. This sample has small baseline fluctuation under the same conditions.
Figure 8a exhibits the selectivity of the WS2 microflake-based sensor and the IO/WS2-1 heterostructure nanocomposite-based sensor for ammonia, formaldehyde, toluene, methanol, ethanol and acetone at room temperature. It indicates that the synthesized IO/WS2-1 heterostructure nanocomposite-based sensor has an obvious response to ammonia. However, this sensor has no obvious response to other test gases, showing good selectivity at room temperature. Based on a previous report for WS2, the WS2 matrix phase has good response to ammonia [24]. However, it does not have responses to other gases such as formaldehyde, toluene, methanol, ethanol and acetone at room temperature. With the introduction of In2O3, the formed IO/WS2 heterostructure improved the performance of WS2 in response to ammonia while not changing the non-responsive function of WS2 for other gases.
Figure 8b shows the influence of the relative humidity on the response/baseline of the IO/WS2-1 sensor for 10 ppm NH3 at room temperature. The IO/WS2-1 sensor attained its maximum response when the humidity was about 50%. When the humidity was lower or higher than 50%, the response of the sensor decreased. The corresponding baseline resistance also had the minimum value at 50% RH, which became larger when the humidity was above or below 50% RH.
In comparison, the ZnO/WS2 and SnO2/WS2 heterostructure nanocomposites were also examined in this work. The detailed sensing characterizations are provided in Figures S4–S8 in the Supplementary Materials. The preparation and testing methods for the two materials are basically the same as those for the In2O3/WS2 heterostructure nanocomposites. The major merits of each type of WS2-based sensor are summarized in Table 1 together with the comparisons to more results reported in the literature. The In2O3-incorporated WS2 nanocomposites show more advantages in detecting NH3 at room temperature in terms of the sensitivity and response/recovery speeds. This indicates that the special “chemical catalytic” properties of the In2O3 in response to ammonia may play a role in this difference in performance. It is reported that In2O3 nanoparticles could provide more hydroxide species, resulting in the promotion of ammonia electro-oxidation [34].

4. Conclusions

Compared to the pure WS2 microflake-based sensor, the sensors based on In2O3-incorporated WS2 heterostructure nanocomposites exhibited a significantly improved response to ammonia at room temperature. In comparison to the ZnO/WS2 and SnO2/WS2 nanocomposites, In2O3-decorated WS2 has better sensing features overall, including a better response with a faster response speed. The WS2 with 1 wt.% In2O3 nanocomposites achieve the highest response and the fastest response/recovery speeds. The sensor exhibited excellent signal repeatability and good long-term stability. It also possesses a good selectivity for formaldehyde, toluene, methanol, ethanol and acetone and is thus a promising candidate for detecting low-concentration NH3 at room temperature.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemosensors10100402/s1, Figure S1. Elemental distribution of IO/WS2-1 heterostructure nanocomposites based sensor; Figure S2. (a,b) W 4f and S 2p refined core spectra of WS2, (c,d) W 4f and S 2p refined core spectra of IO/WS2-1 heterostructures nanocomposites based sensor; Figure S3. In 3d refined core spectra of In2O3; Figure S4. Elemental distribution of SnO2/WS2-1 heterostructure nanocomposites based sensor; Figure S5. (a,b) Response curves and response values of WS2 and ZnO/WS2 heterostructure nanocomposites based sensors to different concentrations of ammonia at room temperature, (c) response/recovery time of WS2 and ZnO/WS2 heterostructure nanocomposites based sensors to 10 ppm ammonia at room temperature, (d) relationship between ZnO/WS2-1 heterostructure nanocomposites based sensor’s response/baseline resistance and humidity; Figure S6. (a). The response curve of the prepared ZnO/WS2 heterostructure nanocomposites based sensors to 0.47 ppm ammonia at room temperature, (b) the selectivity of the WS2 and ZnO/WS2 heterostructure nanocomposites based sensors to different gases at room temperature; Figure S7. (a,b) Response curves and values of WS2 microflakes based sensor and SnO2/WS2 heterostructure nanocomposites based sensors to different concentrations of ammonia at room temperature, (c) response/recovery time of the SnO2/WS2 heterostructure nanocomposites based sensors to 10 ppm ammonia at room temperature, (d) repeatability of SnO2/WS2-1 heterostructure nanocomposites based sensor; Figure S8. (a) The relationship between the response/baseline resistance and humidity of SnO2/WS2-1 heterostructure nanocomposites based sensor, (b) selectivity of the WS2 microflakes based sensor and SnO2/WS2 heterostructure nanocomposites based sensors to different gases; Table S1. The relative concentrations of each element in IO/WS2-1 heterostructure nanocomposites based sensor.

Author Contributions

Conceptualization, J.Y.; methodology, X.L.; validation, B.H.; formal analysis, Q.G.; investigation, Q.G.; resources, J.Z.; data curation, Q.G. and B.H.; writing—original draft preparation, Q.G.; writing—review and editing, X.L.; supervision, X.L.; project administration, X.L.; funding acquisition, B.H. and X.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors highly appreciate the financial support from the National Key R&D Program of China (No. 2021YFB3201302), National Natural Science Foundation of China (No. 61971085, 61874018, 62201117, 62111530055), the Fundamental Research Funds for the Central Universities (No. DUT19RC(3)054) and Dalian Science and Technology Innovation Fund Project (No. 2019J12GX048).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Negwamba, T.; Dinka, M.O. Assessing the performance of Trichardt Wastewater Treatment Plant (South Africa). Phys. Chem. Earth 2019, 114, 102784. [Google Scholar] [CrossRef]
  2. Cecilia, C.; Federica, B.; Marcella, G. Ammonia concentration and recommended threshold values in pig farming: A review. In Proceedings of the IEEE International Workshop on Metrology for Agriculture and Forestry, Trento-Bolzano, Italy, 3–5 November 2021; pp. 162–166. [Google Scholar]
  3. Macedonio, F.; Frappa, M.; Bamaga, O. Application of a membrane condenser system for ammonia recovery from humid waste gaseous streams at a minimum energy consumption. Appl. Water Sci. 2022, 12, 90. [Google Scholar] [CrossRef]
  4. Berwal, P.; Kumar, S.; Khandelwal, B. A comprehensive review on synthesis, chemical kinetics, and practical application of ammonia as future fuel for combustion. J. Energy Inst. 2021, 99, 273–298. [Google Scholar] [CrossRef]
  5. Mallouppas, G.; Ioannou, C.; Yfantis, E.A. A Review of the Latest Trends in the Use of Green Ammonia as an Energy Carrier in Maritime Industry. Energies 2022, 15, 1453. [Google Scholar] [CrossRef]
  6. Gardner, R. Use of the reciprocal calculation procedure for setting workplace emergency action levels for hydrocarbon mixtures and their relationship to lower explosive limits. Ann. Occup. Hyg. 2011, 55, 326–339. [Google Scholar]
  7. Pedersen, J.M.; Feilberg, A.; Kamp, J.N.; Hafner, S.; Nyord, T. Ammonia emission measurement with an online wind tunnel system for evaluation of manure application techniques. Atmos. Environ. 2020, 230, 117562. [Google Scholar] [CrossRef]
  8. Orozco, J.L.; Caneghem, J.V.; Hens, L.; González, L.; Lugo, R.; Díaz, S.; Pedroso, I. Assessment of an ammonia incident in the industrial area of Matanzas. J. Clean. Prod. 2019, 222, 934–941. [Google Scholar] [CrossRef]
  9. Kwak, D.; Lei, Y.; Maric, R. Ammonia gas sensors: A comprehensive review. Talanta 2019, 204, 713–730. [Google Scholar] [CrossRef]
  10. Aarya, S.; Kumar, Y.; Chahota, R.K. Recent Advances in Materials, Parameters, Performance and Technology in Ammonia Sensors: A Review. J. Inorg. Organomet. Polym. 2020, 30, 269–290. [Google Scholar] [CrossRef]
  11. Tsymbalov, A.V.; Kalygina, V.M.; Maksimova, N.K.; Chernikov, E.V. Ammonia sensors based on resistive structures M–SnO2: Sb–M. J. Phys. Conf. Ser. 2020, 1499, 012037. [Google Scholar] [CrossRef]
  12. Kathwate, L.H.; Umadevi, G.; Kulal, P.M.; Nagaraju, P.; Dubal, D.P.; Nanjundan, A.K.; Mote, V.D. Ammonia gas sensing properties of Al doped ZnO thin films. Sens. Actuators A Phys. 2020, 313, 112193. [Google Scholar] [CrossRef]
  13. Yamazoe, N.; Tamaki, J.; Miura, N. Role of hetero-junctions in oxide semiconductor gas sensors. Mater. Sci. Eng. B 1996, 41, 178–181. [Google Scholar] [CrossRef]
  14. Chen, Y.; Wu, J.; Xu, Z.; Shen, W.; Wu, Y.; Corriou, J.P. Computational assisted tuning of Co-doped TiO2 nanoparticles for ammonia detection at room temperatures. Appl. Surf. Sci. 2022, 601, 154214. [Google Scholar] [CrossRef]
  15. Büyükköse, S. Highly selective and sensitive WO3 nanoflakes based ammonia sensor. Mater. Sci. Semicond. Process. 2020, 110, 104969. [Google Scholar] [CrossRef]
  16. Bhardwaj, S.K.; Bhardwaj, N.; Kukkar, M.; Sharma, A.L.; Kim, K.H.; Deep, A. Formation of High-Purity Indium Oxide Nanoparticles and Their Application to Sensitive Detection of Ammonia. Sensors 2015, 15, 31930–31938. [Google Scholar] [CrossRef] [Green Version]
  17. Sharieh, J.K.; Rahim, G.; Maryam, S. Attached two folded graphene nanoribbons as sensitive gas sensor. Phys. B Condens. Matter 2022, 628, 413630. [Google Scholar]
  18. Tang, X.; Debliquy, M.; Lahem, D.; Yan, Y.; Raskin, J.P. A Review on Functionalized Graphene Sensors for Detection of Ammonia. Sensors 2021, 21, 1443. [Google Scholar] [CrossRef]
  19. Rani, S.; Sharma, M.; Verma, D.; Ghanghass, A.; Bhatia, R.; Sameera, I. Two-dimensional transition metal dichalcogenides and their heterostructures: Role of process parameters in top-down and bottom-up synthesis approaches. Mater. Sci. Semicond. Process. 2022, 139, 106313. [Google Scholar] [CrossRef]
  20. Sukhwinder, S.; Jyotirmoy, D.; Utpal, S.; Sandeep, S. MoS2/WO3Nanosheets for Detection of Ammonia. ACS Appl. Nano Mater. 2021, 4, 2594–2605. [Google Scholar]
  21. Xia, Y.; Xu, L.; He, S.; Zhou, L.; Wang, M.; Wang, J.; Komarneni, S. UV-activated WS2/SnO2 2D/0D heterostructures for fast and reversible NO2 gas sensing at room temperature. Sens. Actuators B Chem. 2022, 364, 131903. [Google Scholar] [CrossRef]
  22. Wang, X.; Huang, B.; Wu, X.; Gu, D.; Li, X. Enhanced ammonia sensing properties of rGO/WS2 heterojunction based chemiresistive sensor by marginal sulfonate decoration. Sens. Actuators B Chem. 2021, 337, 129776. [Google Scholar] [CrossRef]
  23. Kim, J.H.; Mirzaei, A.; Kim, J.Y.; Yang, D.H.; Kim, S.S.; Kim, H.W. Selective CO gas sensing by Au-decorated WS2-SnO2 core-shell nanosheets on flexible substrates in self-heating mode. Sens. Actuators B Chem. 2022, 353, 131197. [Google Scholar] [CrossRef]
  24. Li, X.; Li, X.; Li, Z.; Wang, J.; Zhang, J. WS2 nanoflakes based selective ammonia sensors at room temperature. Sens. Actuators B Chem. 2017, 240, 273–277. [Google Scholar] [CrossRef]
  25. Tian, J.; Chen, X.; Wang, T.; Pei, W.; Li, F.; Li, D.; Yang, Y.; Dong, X. Modification of indium oxide nanofibers by polyoxometalate electron acceptor doping for enhancement of gas sensing at room temperature. Sens. Actuators B Chem. 2021, 344, 130227. [Google Scholar]
  26. Yang, Z.; Zhang, D.; Chen, H. MOF-derived indium oxide hollow microtubes/MoS2 nanoparticles for NO2 gas sensing. Sens. Actuators B Chem. 2019, 300, 127037. [Google Scholar] [CrossRef]
  27. Zhang, D.; Wu, J.; Cao, Y. Cobalt-doped indium oxide/molybdenum disulfide ternary nanocomposite toward carbon monoxide gas sensing. J. Alloy. Compd. 2019, 777, 443–453. [Google Scholar] [CrossRef]
  28. Zhang, S.; Song, P.; Zhang, J.; Yan, H.; Li, J.; Yang, Z.; Wang, Q. Highly sensitive detection of acetone using mesoporous In2O3 nanospheres decorated with Au nanoparticles. Sens. Actuators B 2017, 242, 983–993. [Google Scholar] [CrossRef]
  29. Zoleikha, H.; Reza, T.L.; Fereshteh, R.A. 3-Identification and analytical methods. In Heterogeneous Micro and Nanoscale Composites for the Catalysis of Organic Reactions; Elsevier: Amsterdam, The Netherlands, 2022; pp. 33–51. [Google Scholar]
  30. Guang, Q.; Huang, B.; Li, X. Au-Decorated WS2 Microflakes Based Sensors for Selective Ammonia Detection at Room Temperature. Chemosensors 2022, 10, 9. [Google Scholar] [CrossRef]
  31. Huang, B.; Wang, Y.; Hu, Q.; Mu, X.; Zhang, Y.; Bai, J.; Wang, Q.; Sheng, Y.; Zhang, Z.; Xie, E. A low temperature and highly sensitive ethanol sensor based on Au modified In2O3 nanofibers by coaxial electrospinning. J. Mater. Chem. C 2018, 6, 10935. [Google Scholar] [CrossRef]
  32. Liu, Z.; He, T.; Sun, H.; Huang, B.; Li, X. Layered MXene heterostructured with In2O3 nanoparticles for ammonia sensors at room temperature. Sens. Actuators B Chem. 2022, 365, 131918. [Google Scholar] [CrossRef]
  33. Liu, W.; Gu, D.; Li, X. Ultrasensitive NO2 Detection Utilizing Mesoporous ZnSe/ZnO Heterojunction-Based Chemiresistive-Type Sensors. ACS Appl. Mater. Interfaces 2019, 11, 29029–29040. [Google Scholar] [CrossRef] [PubMed]
  34. Silva, J.C.M.; Piasentin, R.M.; Spinace, E.V.; Neto, A.O.; Baranova, E.A. The effect of antimony-tin and indium-tin oxide supports on the catalytic activity of Pt nanoparticles for ammonia electro-oxidation. Mater. Chem. Phys. 2016, 180, 97–103. [Google Scholar] [CrossRef]
  35. Wang, X.Y.; Gu, D.; Li, X.; Lin, S.; Zhao, S.; Rumyantseva, M.N.; Gaskov, A.M. Reduced graphene oxide hybridized with WS2 nanoflakes based heterojunctions for selective ammonia sensors at room temperature. Sens. Actuators B Chem. 2019, 282, 290–299. [Google Scholar] [CrossRef]
  36. Zhao, C.; Wang, P.; Niu, G.; Luo, D.; Wang, Q.; Wang, F. Rapid and Efficient Detection of NH3 at Room Temperature Using CuO/WS2 Nanohybrids. IEEE Sens. J. 2022, 17, 3175827. [Google Scholar] [CrossRef]
  37. Perrozzi, F.; Emamjomeh, S.M.; Paolucci, V.; Taglieri, G.; Ottaviano, L.; Cantalini, C. Thermal stability of WS2 flakes and gas sensing properties of WS2/WO3 composite to H2, NH3 and NO2. Sens. Actuators B Chem. 2017, 243, 812–822. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of WS2 and IO/WS2-1 samples.
Figure 1. XRD patterns of WS2 and IO/WS2-1 samples.
Chemosensors 10 00402 g001
Figure 2. SEM images of (a) WS2 and (b) IO/WS2-1; (c) TEM image of IO/WS2-1; (d) HRTEM image of IO/WS2-1.
Figure 2. SEM images of (a) WS2 and (b) IO/WS2-1; (c) TEM image of IO/WS2-1; (d) HRTEM image of IO/WS2-1.
Chemosensors 10 00402 g002
Figure 3. (a) XPS spectra of WS2, In2O3 nanoparticles and IO/WS2-1 heterostructure nanocomposites; (b) O 1s refined core spectra of WS2 microflakes; (c) O 1s refined core spectra of IO/WS2-1 heterostructure nanocomposites; (d) In 3d refined core spectra of IO/WS2-1 heterostructure nanocomposites.
Figure 3. (a) XPS spectra of WS2, In2O3 nanoparticles and IO/WS2-1 heterostructure nanocomposites; (b) O 1s refined core spectra of WS2 microflakes; (c) O 1s refined core spectra of IO/WS2-1 heterostructure nanocomposites; (d) In 3d refined core spectra of IO/WS2-1 heterostructure nanocomposites.
Chemosensors 10 00402 g003
Figure 4. I–V curves of samples based on WS2 microflakes and IO/WS2 heterostructure nanocomposites.
Figure 4. I–V curves of samples based on WS2 microflakes and IO/WS2 heterostructure nanocomposites.
Chemosensors 10 00402 g004
Figure 5. (a) Responses of WS2 and IO/WS2 heterostructure nanocomposite-based sensors to 10 ppm ammonia at room temperature; (b) corresponding response time and recovery time.
Figure 5. (a) Responses of WS2 and IO/WS2 heterostructure nanocomposite-based sensors to 10 ppm ammonia at room temperature; (b) corresponding response time and recovery time.
Chemosensors 10 00402 g005
Figure 6. (a) Band diagram of the IO/WS2 heterostructure before contact; (b) band diagram of the IO/WS2 heterostructure after contact; (c) illustration of redox reaction occurring at the interface of WS2 and In2O3.
Figure 6. (a) Band diagram of the IO/WS2 heterostructure before contact; (b) band diagram of the IO/WS2 heterostructure after contact; (c) illustration of redox reaction occurring at the interface of WS2 and In2O3.
Chemosensors 10 00402 g006
Figure 7. (a,b) Response curves and values of WS2 microflake-based sensor and IO/WS2 heterostructure nanocomposite-based sensors upon exposure to different concentrations of ammonia at room temperature; (c) repeatability of IO/WS2-1 heterostructure nanocomposite-based sensor at room temperature; (d) long-term stability and baseline of IO/WS2-1 heterostructure nanocomposite-based sensor.
Figure 7. (a,b) Response curves and values of WS2 microflake-based sensor and IO/WS2 heterostructure nanocomposite-based sensors upon exposure to different concentrations of ammonia at room temperature; (c) repeatability of IO/WS2-1 heterostructure nanocomposite-based sensor at room temperature; (d) long-term stability and baseline of IO/WS2-1 heterostructure nanocomposite-based sensor.
Chemosensors 10 00402 g007
Figure 8. (a) Gas-sensing selectivity of WS2 microflake-based sensor and IO/WS2-1 heterostructure nanocomposite-based sensor for different gases (10 ppm); (b) response and baseline resistance of IO/WS2-1 heterostructure nanocomposite-based sensor to 10 ppm ammonia under different humidities.
Figure 8. (a) Gas-sensing selectivity of WS2 microflake-based sensor and IO/WS2-1 heterostructure nanocomposite-based sensor for different gases (10 ppm); (b) response and baseline resistance of IO/WS2-1 heterostructure nanocomposite-based sensor to 10 ppm ammonia under different humidities.
Chemosensors 10 00402 g008
Table 1. A comparison of the performance of NH3 sensors based on WS2 materials.
Table 1. A comparison of the performance of NH3 sensors based on WS2 materials.
Sensitive
Material
NH3
Concentration
Temperature (°C)Response (%)Response/
Recovery (s/s)
Ref.
WS2 nanoflakes5 ppmRT 217 a120/150[24]
rGO/WS210 ppmRT71240/600[35]
2:1 CuO/WS2
nanohybrids
60 ppmRT (30 °C)59.535/213[36]
WS2/WO3
Au/WS2
IO/WS2
ZnO/WS2
SnO2/WS2
10 ppm
1 ppm
1 ppm
1 ppm
1 ppm
150 °C
RT
RT
RT
RT
400
191.54
129.84
133.63
117
~150/~100
968/788
160/44
288/44
224/24
[37]
[30]
This work
This work
This work
a: data obtained by conversion of data in the literature.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guang, Q.; Huang, B.; Yu, J.; Zhang, J.; Li, X. Indium Oxide Decorated WS2 Microflakes for Selective Ammonia Sensors at Room Temperature. Chemosensors 2022, 10, 402. https://doi.org/10.3390/chemosensors10100402

AMA Style

Guang Q, Huang B, Yu J, Zhang J, Li X. Indium Oxide Decorated WS2 Microflakes for Selective Ammonia Sensors at Room Temperature. Chemosensors. 2022; 10(10):402. https://doi.org/10.3390/chemosensors10100402

Chicago/Turabian Style

Guang, Qiyilan, Baoyu Huang, Jun Yu, Jianwei Zhang, and Xiaogan Li. 2022. "Indium Oxide Decorated WS2 Microflakes for Selective Ammonia Sensors at Room Temperature" Chemosensors 10, no. 10: 402. https://doi.org/10.3390/chemosensors10100402

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