Next Article in Journal
Study of Biocrudes Obtained via Hydrothermal Liquefaction (HTL) of Wild Alga Consortium under Different Conditions
Next Article in Special Issue
Enhanced Propanol Response Behavior of ZnFe2O4 NP-Based Active Sensing Layer Induced by Film Thickness Optimization
Previous Article in Journal
Combustion Efficiency in a Fluidized-Bed Combustor with a Modified Perforated Plate for Air Distribution
Previous Article in Special Issue
Effect of  Nb5+ and In3+  Ions on Moisture Sensitivity of Electrospun Titanium/Tungsten Oxide Nanostructures: Microstructural Characterization and Electrical Response
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hierarchical Nanoflowers of Colloidal WS2 and Their Potential Gas Sensing Properties for Room Temperature Detection of Ammonia

1
Molecular Sciences Institute, School of Chemistry, University of The Witwatersrand, Johannesburg 2050, South Africa
2
Departamento de Física, Universidade Federal do Paraná, Curitiba 81531-980, Brazil
3
Microscopy and Microanalysis Unit, University of The Witwatersrand, Johannesburg 2050, South Africa
4
Institute for Water Technology and Sustainability, College of Science, Engineering and Technology, University of South Africa, Florida Campus, Johannesburg 1710, South Africa
*
Authors to whom correspondence should be addressed.
Processes 2021, 9(9), 1491; https://doi.org/10.3390/pr9091491
Submission received: 30 June 2021 / Revised: 28 July 2021 / Accepted: 2 August 2021 / Published: 25 August 2021

Abstract

:
A one-step colloidal synthesis of hierarchical nanoflowers of WS2 is reported. The nanoflowers were used to fabricate a chemical sensor for the detection of ammonia vapors at room temperature. The gas sensing performance of the WS2 nanoflowers was measured using an in-house custom-made gas chamber. SEM analysis revealed that the nanoflowers were made up of petals and that the nanoflowers self-assembled to form hierarchical structures. Meanwhile, TEM showed the exposed edges of the petals that make up the nanoflower. A band gap of 1.98 eV confirmed a transition from indirect-to-direct band gap as well as a reduction in the number of layers of the WS2 nanoflowers. The formation of WS2 was confirmed by XPS and XRD with traces of the oxide phase, WO3. XPS analysis also confirmed the successful capping of the nanoflowers. The WS2 nanoflowers exhibited a good response and selectivity for ammonia.

1. Introduction

Effective air quality management requires regular monitoring of both indoor and outdoor environments for the detection of harmful and toxic pollutants such as NH3. Even though semiconducting metal oxides (SMOs) are still the front runners in the category of chemiresistive gas sensors, their high energy input requirement (300–500 °C) is a major drawback [1,2]. Therefore, an alternative that can be incorporated into low-power operating air monitoring systems and still provide a high response even at room temperature (RT) is sought [3]. As a result, the gas sensing research community is putting a lot of effort into the research on other materials such as the transition metal dichalcogenides (TMDCs). TMDCs are layered semiconducting materials with outstanding electronic, chemical and mechanical stabilities [3,4] and a large surface-to-volume ratio [3,5]. Another interesting feature about TMDCs is their layer-dependent properties such as the band gap [4], indirect-to-direct band gap transition [6,7], electronic transport [5,8] and gas sensing [5]. WS2 is steadily catching up with its TMDC family member, MoS2, with respect to the research dedicated to its potential to detect NH3. Huo et al. reported on the dynamic response of exfoliated multilayer WS2 nanoflakes upon exposure to the reducing gas NH3 at room temperature [9]. A transfer of charge between the adsorbed NH3 and the exfoliated multilayer WS2 nanoflakes resulted in increased conductivity. NH3 donated electrons to the WS2 nanoflakes, thereby inducing an n-type behavior. High NH3 room temperature sensitivity was also observed with thin films of WS2 obtained via a plasma-assisted synthesis [6]. However, the thin films displayed incomplete recovery. This is a common feature with nanomaterial-based sensors operating at low temperatures; the thin films can be subjected to annealing or UV light illumination to speed up the recovery. Remarkable selectivity to interferents, good selectivity, significant response/recovery rates and a p-type response was reported by Li and co-workers when WS2 nanoflakes produced by ball milling commercial powders were exposed to NH3 at 20–200 °C [10]. The sensor response increased with an increase in humidity level (up to 73% relative humidity, RH).
The influence of layer numbers on the recovery rate after removal of NH3 gas also came under scrutiny in Qin et al.’s study [11]. The WS2 thin films in this study were fabricated from nanosheets which were obtained via lithium ion intercalation. The monolayer-based thin film had the shortest recovery time compared to its few-layered and bulk counterparts. It was proposed that the NH3 that is in the interlayers of WS2 is not easily desorbed, hence the slow or incomplete recovery by few-layered thin films. Results from a study by Perozzi and co-workers on the thermal stability of thin films prepared from commercially bought WS2 revealed progressive oxidation to WO3 within the 25–450 °C range [12]. Changes in morphology were noted at 150–250 °C and regarded as the indication of the formed WS2/WO3 composite. The gas sensing performance of the WS2/WO3 composite towards NH3 decreased at annealing temperatures above 150 °C. Meanwhile, the 150 °C annealed composite recovered completely and displayed no cross sensitivity to water vapor at 60% RH. High sensitivity and fast recovery to room temperature NH3 sensing and p-type character were shown by the hydrothermally prepared nanocomposite of Pt QDs/WS2 nanosheets [13]. Other composites, such as nanosheets of WS2|O [3] and WS2/TiO2 [14], also exhibited high sensitivity, excellent selectivity and fast recovery at room temperature.
The nanoflower morphology presents a larger surface area, a larger number of active sites at the edges of the WS2 layers and many interlayer spaces, which are all desirable in gas sensing. To exploit the properties of WS2 nanomaterials, methods that enable precise control of the morphology are required. The reported synthetic routes for WS2 nanoflowers are hydrothermal [15,16,17,18,19,20,21,22,23] and CVD [24,25]. Colloidal synthesis also offers precise control of the reaction parameters in order to obtain desirable morphology. It is catalyst-free, template-free, one-pot, easily scalable and short one-step synthesis at relatively low temperatures. It also accommodates the use of a capping agent which offers protection of the nanoparticles against agglomeration [26], modifies the surface of the nanoparticles, may introduce new functionalities [27] and can influence the type of morphology formed [28]. Oleylamine (OLA), an N-terminated ligand, is suitable for transition metal semiconductors. OLA is low cost, and can act as a solvent and reducing agent, thereby eliminating the use of many chemicals; it can also lower the decomposition temperature of a metal precursor. Nanoparticles capped by OLA are easily dispersed in various organic solvents with improved properties and potential applications in various fields [26,29]. Colloidally synthesized WS2 nanoflowers have not been extensively reported and their gas sensing potential has not been explored.
In this study, a colloidal method with OLA as both solvent and capping agent was followed to obtain hierarchical WS2 nanoflowers. The nanoflowers were obtained after 45 min and were tested at room temperature for potential ammonia gas sensing properties. A good response and higher selectivity were observed towards ammonia vapors than with acetone, chloroform, ethanol and toluene. The dependence of sensing properties of a particular sensor on its method of synthesis is well known. To our knowledge, colloidal hierarchical nanoflowers of WS2 have not been synthesized for application in gas sensing.

2. Materials and Methods

2.1. Chemicals

Analytical grade tungstic acid (H2WO4), thiourea (CS(NH2)2), oleylamine (OLA), ammonium hydroxide (NH4OH), ethanol (CH3CH2OH), hexane (C6H14), chloroform (CHCl3), toluene (C6H5CH3), acetone ((CH3)2CO) and isopropyl alcohol (C3H8O) were purchased from Sigma-Aldrich, Johannesburg, South Africa and DieLab, Curitiba, Brazil. Interdigitated electrodes (IDEs) of electroless nickel immersion gold (ENIG) (18 pairs, 7.9 mm long and 0.1 mm apart) were purchased from Micropress SA, Curitiba, Brazil.

2.2. Synthesis of WS2 Nanoflowers

Hierarchical nanoflowers of WS2 were synthesized by mixing H2WO4 (0.05 mol) and CS(NH2)2 (0.2 mol) in 20 mL of degassed OLA. The mixture was heated at 25 °C with continuous stirring under N2 gas flow for 15 min in a three-neck round-bottom flask. The temperature was increased rapidly to 320 °C to allow for the decomposition of both precursors and held at this temperature for 45 min. After cooling the black reaction mixture for 5 min, ethanol was added in order to separate the colloids. WS2 nanoflowers were collected by centrifugation and washed several times with a mixture of ethanol and hexane (1:1). The WS2 black powders were dried at room temperature.

2.3. Material Characterization

2.3.1. Optical Characterization

The WS2 powder was dispersed in CHCl3 and placed in quartz cuvettes (1 cm path length) for UV–Vis absorption and PL spectral measurements on the Specord 50 Analytik Jena UV–Vis spectrophotometer and Agilent Cary Eclipse fluorescence spectrometer, respectively. Raman spectroscopy (Bruker Senterra Infinity 1 software, 50× optical objective, 532 nm laser wavelength, 0.2 mV laser power and integration power of 15 s) measurements of the dry powder of WS2 nanoflowers were used to estimate the number of layers.

2.3.2. Structural Characterization

Measurements were performed on the WS2 powder with the Bruker MeasSrv (D2-205530)/D2-205530 diffractometer to shed light on the structure and phase of the powdered nanoflowers. X-ray photoelectron spectroscopy measurements were carried out with a PHI 5000 Versaprobe—Scanning ESCA Microprobe (100 µm 25 W 15 kV Al monochromatic X-ray beam) to determine the surface properties of the powders. Sizes and morphologies of the nanomaterials were studied using an FEI Nova NanoLab FIB/SEM and a JEOL JEM-2100 field emission gun transmission electron microscope (200 kV) Chemical composition of the crystal structure was analyzed with energy-dispersive X-ray spectroscopy (EDS) integrated into the TEM instrument.

2.4. Device Fabrication and Gas Sensing Measurements

The WS2 nanoflower-based sensor was fabricated by drop casting 20 μL of the 5 mg/mL WS2 toluene dispersion on clean IDEs. The sensor was dried in an oven at 130 °C for 30 min. A surface profiler (Dektak XT; Bruker) was used to measure the thickness of the film (about 600 nm). A custom-made gas chamber, depicted in Figure 1, housed the sensor. The gas chamber was grounded and stabilized for 1 h before measurements were started.
An LCR meter (Agilent 4284A 20 Hz–1 MHz Precision LCR meter) was attached to a computer interfaced with a GPIB for data acquisition and set at an operational voltage and frequency of 1000 mV and 10 kHz, respectively. A large variation in conductance and relative signal-to-noise were obtained at these parameters. The measurements were performed in the dark and under dry nitrogen to provide a controlled atmosphere. The sensor was exposed to incremental concentrations of NH3 (NH4OH was used as the source of NH3) at RT (~23 °C) and 25% RH. A 50 s stabilizing period followed by introduction of 1.5 µL of NH4OH (analyte) at every 200 s was applied per measurement. The time interval was enough to evaporate the analyte and saturate the chamber. The measurements were taken with concentrations of NH4OH in the range of 240–958 ppm. The selective character of the sensor was established by exposing it to incremental concentrations of interferents such as acetone, chloroform, ethanol and toluene. Response and recovery measurements were carried out to determine how fast the sensor responds to NH3 and how fast it recovers to its initial chemical status after NH3 is removed. The effect of humidity on the sensor’s gas performance was determined by measuring its response to NH3 under various RH levels.

3. Results

3.1. Characterization of WS2 Nanoflowers

The crystal phase, crystallinity and composition of the WS2 nanoflowers were determined with PXRD. The diffraction patterns, as shown in Figure 2, confirmed the nanoflowers to be the 2H-WS2 polytype according to PDF No: 00-002-0131\JCP2.2CA. The diffraction peaks correspond to the (002), (004), (101), (103), (006), (008) and (200) planes. The prominence of the (002) plane suggested the existence of more than one layer of WS2 with good crystallinity while its broadness is an indication of a reduction in size. Closer analysis hinted at partial oxidation based on the small peak at 29.5° (denoted by an asterisk) which was assigned to WO3. This came as no surprise as WS2 is known for spontaneous oxidation.
XPS studies provided detailed composition of the nanoflowers. The W and S peaks are clearly displayed on the survey spectrum, shown in Figure 3. Partial oxidation of WS2 to WO3 was confirmed by the presence of the O 1s peak. The oxygen peak is also attributed to the oxidation of OLA. The successful capping of the nanoflowers by the organic ligand OLA was confirmed by the presence of the strong C 1s peak. Atewolegun et al. used XPS in their studies to confirm the success of ligand exchange on colloidal quantum dots [30]. The capping of the samples is further confirmed by the large carbon composition (77.4%) of the sample, as seen in Table 1.
Figure 4 shows the high-resolution core level spectra of WS2 with the deconvoluted C-C peak by OLA. The O 1s, C-O and O-C=O peaks were attributed to the oxidation of OLA. Four components of the W4f core level spectrum were identified; W 4f7/2 (30.9 eV), W 4f5/2 (33.2 eV) doublet, W 5p3/2 (34.4 eV) and WO3 (36.3 eV). Meanwhile, the W 4f7/2 and W 4f5/2 are ascribed to the 2H-WS2 polytype. W 5p3/2 is attributed to the partially coordinated W in WO3. The S2p spectrum has three peaks, S2Wdef (160.5 eV), S2W (161.6 eV) and S2W (162.8 eV) which were observed in the deconvoluted core level spectrum.
Raman spectroscopy was used to estimate the number of the layers in the nanoflowers. The WS2 nanoflowers showed Raman features which are characteristic of the second order longitudinal acoustic 2LA(M) and out-of-plane A1g(Γ) at approximately 351 cm1 and 417 cm1, respectively, as seen in Figure 5. The shoulder peak at 313 cm1 belongs to the in-plane E12g(Γ). The shift to lower frequencies from bulk for both A1g(Γ) and 2LA(M) (356 and 421 cm1, respectively) is associated with decreasing interlayer interactions by van der Waals forces. This suggested the formation of few layers, in agreement with Varghese et al. [31] and Tan et al. [32]. The calculated frequency difference between A1g(Γ) and 2LA(M) as well as the intensity of 2LA(M), suggesting a reduction in size from bulk to a few layers.
SEM analysis of as-synthesized WS2 revealed nanosheets arranged as petals to form the nanoflower morphology, as seen from Figure 6A,B. These nanoflowers self-assembled to form hierarchical structures. Figure 6B gives a clearer picture of the well-ordered nanosheets with interspaces. This feature of nanoflowers translates to a larger surface area, which is ideal for gas sensing.
Meanwhile, the TEM image from Figure 7A confirmed that the nanoflowers were made up of individual nanosheets. The edges of the nanosheets are exposed and serve as active sites. The microspheres in Figure 7B are consistent with the nanoflower morphology recorded by SEM in Figure 6A,B.
UV–Vis absorption and photoluminescence spectroscopy further confirmed the reduced size of WS2. Figure 8 shows the blue-shifted 625–630 nm and 505–515 nm peaks corresponding to the excitons A and B (636 and 525 nm for bulk WS2, respectively) [33]. The existence of few layers was further suggested by the diminished photoluminescence peak, shown in Figure 8. A similar result was reported by Gutiérrez et al. and was said to be the result of a competition between indirect and direct transitions [34].

3.2. Gas Sensing Properties of the WS2 Nanoflowers

The sensing principle of thin films of semiconducting TMDCs is generally based on the change in conductance due to the reaction of the gas that is adsorbed on the surface. An electric charge is transferred between the gas (target analyte) and the active material (WS2 nanoflowers), causing changes in the electrical properties of the sensing material. Preliminary gas sensing measurements at RT under dry N2 (25% RH) revealed a fast decrease in the conductance, G, of the WS2 nanoflowers. The sharp decrease in conductance, G, as shown in Figure 9A (red curve), was due to the adsorption of NH3 gas molecules on the surface of the WS2 nanoflowers. The negative slope (sensitivity = −7.99 ppm−1) as shown in Figure 9B confirmed the decrease in conductance. The response was linearly proportional to the concentration of NH3. NH3 is a Lewis base and therefore serves as an electron donor. In this work, the NH3 lone pair electron was transferred to the conduction band of the WS2 nanoflower-based sensor upon adsorption. The reduction in the electrical conductance suggested that positive holes are the main charge carriers on the surface of WS2 nanoflowers, hence the p-type doping behavior. This is due to the depletion of positive (hole) charge carriers on the sensor surface by the negative charge carriers from NH3. A similar p-type behavior was observed previously for MoS2 and WS2 sensors by Järvinen et al. [35]. The gas response of the WS2 nanoflower-based sensor to NH3 vapors was determined from the variation in the conductance, G, by using ΔG/G0 in the equation below:
S = Δ G G 0
where ΔG = (GG0). G is the maximum conductance under NH3 vapors while G0 is conductance under dry N2. The sensitivity, S (in ppm−1), of the sensor was determined from the slope of the fit.
Specificity or selectivity is another parameter that determines the practical use of a sensor. It is the ability of a sensor to detect the target analyte in the presence of other contaminants. The specificity value for a sensor is between 0 and 1 with a value closer to 1 representing high selectivity of the sensor for the target analyte relative to the interferents. Consider a particular sensor and a set of n species; the specificity to the particular species i, δi can defined as follows:
δ i = ( Δ G G 0 ) i j = 1 n ( Δ G G 0 ) j
The equation above was adapted from Llobet et al. [36] and was used to calculate the specificity value (δ) of the sensor to NH3 and the interferents. In this work, the sensor showed higher selectivity for ammonia than to acetone, chloroform, ethanol and toluene, as shown in Figure 10. The specificity values of the vapors are summarized in Table 2. Therefore, WS2 nanoflowers have the potential for use as elements in chemical sensor arrays.
For practical purposes, a sensor must respond fast upon exposure to the target gas and recover just as fast when the gas is removed. The characteristic sensor curve (dynamic range) shown in Figure 11 was analyzed and used to estimate the response and recovery speed at incremental concentrations of NH3. The conductance of the sensor decreased sharply (~28 s) upon exposure to NH3, giving a negative response, and slowly recovered (~42 s) when removed from NH3. Slow recovery at RT is common to TMDC thin film-based NH3 sensors [11,37] due to strong interactions between NH3 molecules and the active layer. This leads to analyte accumulation on the surface of the sensor [38].
Theoretical calculations predicted that the analyte is physically adsorbed on the surface of a perfect 2D monolayer [39]. However, this is not the case for few-layered and bulk WS2 sensors where the NH3 molecule can be inserted into the inner layers and interact with the two adjacent layers, as shown in Figure 12 below [11].
Intercalation compounds such as (NH4+)X(NH3)Y(WS2)x− may be formed. The intercalated NH3 molecules are more difficult to desorb than the surface molecules. This behavior is similar to the intercalation and deintercalation processes of NH3 with layered TiS2 and TaS2 which were detailed by McKelvy and Glaunsinger [40]. Based on the confirmation of the existence of a few layers of WS2 by XRD, Raman and TEM, the argument of intercalated NH3 molecules can be applied in this study to explain the slow recovery of the sensor. Furthermore, the nanoflower morphology has interspaces which provide more reactive sites; thus, more NH3 molecules end up deeper inside the layers, leading to a slow recovery. It is worth mentioning that defects are introduced during colloidal synthesis, and this too provides more reactive sites than the perfect lattice of a 2D monolayer. The WS2 nanoflower-based sensor recovered completely after irradiation with UV light for 60 min or heating in an oven for 10 min at 100 °C.
Water molecules are easily adsorbed on the surface of the sensor, leading to an increase or decrease in its performance. This is common with RT chemical sensors. As the WS2 nanoflower-based sensor is designed for application in RT sensing of NH3, it was imperative to measure its performance under various levels of humidity. Figure 13 illustrates the response of the sensor at different RH conditions. The effect of humidity on the gas sensing performance of WS2 nanoflowers was found to be more pronounced at 97% RH.
The increase in NH3 response in the presence of humidity could be attributed to the increased acidity due to the water molecules on the surface of the WS2 nanoflower-based sensor [10]. Such an increase can be explained by the hydroxylation reaction below:
SO42− (WS2 surface) + H2O ↔ H+ + SO42− + OH
The NH3 molecules donated more electrons to the already acidic surface of the WS2 sensor, resulting in the observed increase in response.

4. Conclusions

WS2 nanoflowers were successfully synthesized after 45 min via a simple colloidal route. The colloidal nanoflowers displayed a good response and higher selectivity towards NH3 vapors relative to acetone, ethanol, toluene and chloroform. The gas sensing performance exhibited by the nanoflowers is evidence that the partial oxidation did not adversely compromise it. Humidity interference was established and its significance was observed at very high % RH values. Evidence of incomplete recovery due to the strong interaction between NH3 molecules and the sensor was observed. The sensor has potential for use as an active material in RT chemiresistive sensors and sensor arrays. Studies are underway to develop strategies to improve the overall gas sensing performance of the WS2 nanoflowers by doping with metal oxides, metals, carbon materials, perovskites and other TMDCs. The focus is on reducing the recovery time, enhancing the sensitivity and selectivity to NH3 and slowing down oxidation. The sensor will also be tested for sensitivity to NO2 and CO.

Author Contributions

S.S.G., N.M., R.R. and I.A.H. conceptualized the experiments. S.S.G., S.L.M., Z.N., P.O.F. and R.R. performed the experiments. S.S.G. and R.R. analyzed the data. S.S.G. and R.R. prepared the original draft. The review and editing of the manuscript were done by S.S.G., R.R., M.A., I.A.H., N.M., E.C.L., M.J.M., N.M. and I.A.H. supervised the research project. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research Foundation (TTK180419322895), the University of the Witwatersrand and the Department of Higher Education.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge the Department of Physics, Universidade Federal do Paranà, Brazil for technical support and materials used for experiments.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ko, K.Y.; Park, K.; Lee, S.; Kim, Y.; Woo, W.J.; Kim, D.; Song, J.-G.; Park, J.; Kim, H. Recovery Improvement for Large-Area Tungsten Diselenide Gas Sensors. ACS Appl. Mater. Interfaces 2018, 10, 23910–23917. [Google Scholar] [CrossRef]
  2. Kaewmaraya, T.; Ngamwongwan, L.; Moontragoon, P.; Jarernboon, W.; Singhd, D.; Ahujad, R.; Karton, A.; Hussain, T. Novel green phosphorene as a superior chemical gas sensing material. J. Hazard. Mater. 2021, 401, 123340. [Google Scholar] [CrossRef] [PubMed]
  3. Zheng, Y.; Sun, L.; Liu, W.; Wang, C.; Dai, Z.; Ma, F. Tungsten oxysulfide nanosheets for highly sensitive and selective NH3 sensing. J. Mater. Chem. C 2020, 8, 4206–4214. [Google Scholar] [CrossRef]
  4. Kumar, R.; Goel, N.; Hojamberdiev, M.; Kumar, M. Transition metal dichalcogenides-based flexible gas sensors. Sens. Actuators A 2020, 303, 111875. [Google Scholar] [CrossRef]
  5. Wang, W.; Cao, J.; Zhou, J.; Chen, J.; Liu, J.; Deng, H.; Zhang, Y.; Li, X. Highly enhanced performance for sensing by monolayer 1T’ WS2 with atomic vacancy. Microelectron. Eng. 2020, 223, 111215. [Google Scholar] [CrossRef]
  6. Gatensby, R.; McEvoy, N.; Lee, K.; Hallam, T.; Berner, N.C.; Rezvani, E.; Winters, S.; O’Brien, M.; Duesberg, G.S. Controlled synthesis of transition metal dichalcogenide thin films for electronic applications. Appl. Surf. Sci. 2014, 297, 139–146. [Google Scholar] [CrossRef]
  7. Samadi, M.; Sarikhani, N.; Zirak, M.; Zhang, H.; Zhang, H.-L.; Moshfegh, A.Z. Group 6 transition metal dichalcogenide nanomaterials: Synthesis, applications and future perspectives. Nanoscale Horiz. 2018, 3, 90–204. [Google Scholar]
  8. Pumera, M.; Loo, A.H. Layered transition-metal dichalcogenides (MoS2 and WS2) for sensing and biosensing. Trends Anal. Chem. 2014, 61, 49–53. [Google Scholar] [CrossRef]
  9. Huo, N.; Yang, S.; Wei, Z.; Li, S.-S.; Xia, J.-B.; Li, J. Photoresponsive and Gas Sensing Field-Effect Transistors based on Multilayer WS2 Nanoflakes. Sci. Rep. 2014, 4, 5209. [Google Scholar] [CrossRef] [Green Version]
  10. Li, X.; Li, X.; Li, X.; Wang, J.; Zhang, Z. WS2 Nanoflakes Based Selective Ammonia Sensors at Room Temperature. Sens. Actuators B Chem. 2017, 240, 273–277. [Google Scholar] [CrossRef]
  11. Qin, Z.; Zeng, D.; Zhang, J.; Wu, C.; Wen, Y.; Shan, B.; Xie, C. Effect of layer number on recovery rate of WS2 nanosheets for ammonia detection at room temperature. Appl. Surf. Sci. 2017, 414, 244–250. [Google Scholar] [CrossRef]
  12. Perrozzi, F.; Emamjomeh, S.M.; Paolucci, V.; Taglier, 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]
  13. Ouyang, C.; Chen, Y.; Qin, Z.; Zeng, D.; Zhang, J.; Wang, H.; Xi, C. Two-dimensional WS2-based nanosheets modified by Pt quantum dots for enhanced room-temperature NH3 sensing properties. Appl. Surf. Sci. 2018, 455, 45–52. [Google Scholar] [CrossRef]
  14. Qin, Z.; Ouyang, C.; Zhang, J.; Wan, L.; Wang, S.; Xie, C.; Zeng, D. 2D WS2 nanosheets with TiO2 quantum dots decoration for high-performance ammonia gas sensing at room temperature. Sens. Actuators B 2017, 253, 1034–1042. [Google Scholar] [CrossRef]
  15. Wu, Y.-C.; Liu, Z.-M.; Chen, J.-T.; Cai, X.-J.; Na, P. Hydrothermal fabrication of hyacinth flower-like WS2 nanorods and their photocatalytic properties. Mater. Lett. 2017, 189, 282–285. [Google Scholar] [CrossRef]
  16. Li, X.; Zhang, J.; Liu, Z.; Fu, C.; Niu, C. WS2 nanoflowers on carbon nanotube vines with enhanced electrochemical performances for lithium and sodium-ion batteries. J. Alloys Compd. 2018, 766, 656–662. [Google Scholar] [CrossRef]
  17. Nguyen, T.P.; Kim, S.Y.; Lee, T.H.; Jang, H.W.; Led, Q.V.; Kim, I.T. Facile synthesis of W2C@WS2 alloy nanoflowers and their hydrogen generation performance. Appl. Surf. Sci. 2020, 504, 144389. [Google Scholar] [CrossRef]
  18. Srinivaas, M.; Wu, C.-Y.; Duh, J.-G.; Wu, J.M. Highly Rich 1T Metallic Phase of Few-Layered WS2 Nanoflowers for Enhanced Storage of Lithium-Ion Batteries. ACS Sustain. Chem. Eng. 2019, 7, 10363–10370. [Google Scholar] [CrossRef]
  19. Tekalgne, M.; Hasani, A.; Le, Q.V.; Nguyen, T.P.; Choi, K.S.; Lee, T.H.; Jang, H.W.; Luo, Z.; Kim, S.Y. CdSe Quantum Dots Doped WS2 Nanoflowers for Enhanced Solar Hydrogen Production. Phys. Status Solidi A 2019, 216, 180085. [Google Scholar] [CrossRef]
  20. Hasani, A.; Nguyen, T.P.; Tekalgne, M.; Le, Q.V.; Choi, K.S.; Lee, T.H.; Park, T.J.; Jang, H.W.; Kim, S.Y. The role of metal dopants in WS2 nanoflowers in enhancing the hydrogen evolution reaction. Appl. Catal. A Gen. 2018, 567, 73–79. [Google Scholar] [CrossRef]
  21. Cao, C.Z.; Peng, L.; Han, T. Synthesis of uniform WS2 nanoflowers via a sodium silicate-assisted hydrothermal process. J. Mater. Sci. Mater. Electron. 2016, 27, 3821–3825. [Google Scholar] [CrossRef]
  22. Liu, S.; Shen, B.; Niu, Y.; Xu, M. Fabrication of WS2-nanoflowers@rGO composite as an anode material for enhanced electrode performance in lithium-ion batteries. J. Colloid Interface Sci. 2017, 488, 20–25. [Google Scholar] [CrossRef]
  23. Cao, S.; Liu, T.; Zeng, W.; Hussain, S.; Peng, X.; Pan, F. Synthesis and characterization of flower-like WS2 nanospheres via a facile hydrothermal route. J. Mater. Sci. Mater. Electron. 2014, 25, 4300–4305. [Google Scholar] [CrossRef]
  24. Prabakaran, A.; Dillon, F.; Melbourne, J.; Jones, L.; Nicholls, R.J.; Holdway, P.; Britton, J.; Koos, A.A.; Crossley, A.; Nellist, P.D.; et al. WS2 2D nanosheets in 3D nanoflowers. Chem. Commun. 2014, 50, 12360–12362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Zhang, X.; Wang, J.; Xu, H.; Tan, H.; Ye, X. Preparation and Tribological Properties of WS2 Hexagonal Nanoplates and Nanoflowers. Nanomaterials 2019, 9, 840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Javed, R.; Zia, M.; Naz, S.; Aisida, S.O.; Ain, N.u.; Ao, Q. Role of capping agents in the application of nanoparticles in biomedicine and environmental remediation: Recent trends and future prospects. J. Nanobiotechnol. 2020, 18, 172. [Google Scholar] [CrossRef]
  27. Campisi, S.; Schiavoni, M.; Chan-Thaw, C.E.; Villa, A. Untangling the Role of the Capping Agent in Nanocatalysis: Recent Advances and Perspectives. Catalysts 2016, 6, 185. [Google Scholar] [CrossRef] [Green Version]
  28. Phan, C.H.; Nguyen, H.M. Role of Capping Agent in Wet Synthesis of Nanoparticles. J. Phys. Chem. A 2017, 121, 3213–3219. [Google Scholar] [CrossRef]
  29. Gulati, S.; Sachdeva, M.; Bhasin, K.K. Capping Agents in Nanoparticle Synthesis: Surfactant and Solvent System. AIP Conf. Proc. 2018, 1953, 030214. [Google Scholar]
  30. Atewologun, A.; Ge, W.; Stiff-Roberts, A.D. Characterization of Colloidal Quantum Dot Ligand Exchange by X-ray Photoelectron Spectroscopy. J. Electron. Mater. 2013, 42, 809–814. [Google Scholar] [CrossRef]
  31. Wang, X.; 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]
  32. Lee, J.-H. Gas Sensors Using Hierarchical and Hollow Oxide Nanostructures: Overview. Sens. Actuators B Chem. 2009, 140, 319–336. [Google Scholar] [CrossRef]
  33. Li, Y.-X.; Guo, Z.; Su, Y.; Jin, X.-B.; Tang, X.-H.; Huang, J.-R.; Huang, X.-J.; Li, M.-Q.; Liu, J.-H. Hierarchical Morphology-Dependent Gas-Sensing Performances of Three-Dimensional SnO2 Nanostructures. ACS Sens. 2017, 2, 102–110. [Google Scholar] [CrossRef]
  34. Gutiérrez, H.R.; Perea-López, N.; Elías, A.L.; Berkdemir, A.; Wang, B.; Lv, R.; López-Urías, F.; Crespi, V.H.; Terrones, H.; Terrones, M. Extraordinary Room-Temperature Photoluminescence in Triangular WS2 Monolayers. Nano Lett. 2013, 13, 3447–3454. [Google Scholar] [CrossRef] [Green Version]
  35. Järvinen, T.; Lorite, G.S.; Peräntie, J.; Toth, G.; Saarakkala, S.; Virtanen, V.K.; Kordas, K. WS2 and MoS2 thin film gas sensors with high response to NH3 in air at low temperature. Nanotechnology 2019, 30, 405501. [Google Scholar] [CrossRef] [Green Version]
  36. Llobet, E.; Molas, G.; Molinàs, P.; Calderer, J.; Vilanova, X.; Brezmes, J.; Sueiras, J.E.; Correiga, X. Fabrication of Highly Selective Tungsten Oxide Ammonia Sensors. J. Electrochem. Soc. 2000, 147, 776–779. [Google Scholar] [CrossRef]
  37. Late, D.J.; Huang, Y.-K.; Liu, B.; Acharya, J.; Shirodkar, S.N.; Luo, J.; Yan, A.; Charles, D.; Waghmare, U.V.; Dravid, V.P.; et al. Sensing behavior of atomically thin-layered MoS2 transistors. ACS Nano 2013, 7, 4879–4891. [Google Scholar] [CrossRef]
  38. Marr, I.; Groß, A.; Moos, R. Overview on Conductometric Solid-state Gas Dosimeters. J. Sens. Sens. Syst. 2014, 3, 29–46. [Google Scholar] [CrossRef] [Green Version]
  39. Zeng, Y.; Lin, S.; Gu, D.; Li, X. Two-Dimensional Nanomaterials for Gas Sensing Applications: The Role of Theoretical Calculations. Nanomaterials 2018, 8, 851. [Google Scholar] [CrossRef] [Green Version]
  40. McKelvey, M.J.; Glaunsinger, W.S. On the Intercalation and Deintercalation Mechanisms for Ammoniated Titanium Disulfide. Solid State Ion. 1987, 25, 287–294. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the gas chamber used for sensor measurements.
Figure 1. Schematic representation of the gas chamber used for sensor measurements.
Processes 09 01491 g001
Figure 2. X-ray diffraction pattern of WS2 nanoflowers. The peak at 29.5° (denoted by *) was assigned to WO3.
Figure 2. X-ray diffraction pattern of WS2 nanoflowers. The peak at 29.5° (denoted by *) was assigned to WO3.
Processes 09 01491 g002
Figure 3. XPS survey spectrum of WS2 nanoflowers.
Figure 3. XPS survey spectrum of WS2 nanoflowers.
Processes 09 01491 g003
Figure 4. High-resolution core level spectra of WS2 nanoflowers with focus on C1s, O1s, W4f and S2p. The blue, green, red and purple lines are their respective deconvoluted spectra.
Figure 4. High-resolution core level spectra of WS2 nanoflowers with focus on C1s, O1s, W4f and S2p. The blue, green, red and purple lines are their respective deconvoluted spectra.
Processes 09 01491 g004
Figure 5. Raman spectra of WS2 nanoflowers.
Figure 5. Raman spectra of WS2 nanoflowers.
Processes 09 01491 g005
Figure 6. (A,B) SEM characterization of WS2 nanoflowers.
Figure 6. (A,B) SEM characterization of WS2 nanoflowers.
Processes 09 01491 g006
Figure 7. (A,B) TEM images of WS2 nanoflowers.
Figure 7. (A,B) TEM images of WS2 nanoflowers.
Processes 09 01491 g007
Figure 8. UV–Vis absorption (black line) and photoluminescence spectra (red line) of WS2 nanoflowers.
Figure 8. UV–Vis absorption (black line) and photoluminescence spectra (red line) of WS2 nanoflowers.
Processes 09 01491 g008
Figure 9. Response of the WS2 nanoflower-based sensor at RT. (A) Conductance under dry N2 conditions (black curve) and after exposure to 958 ppm of NH3 (red curve) as a function of time. (B) Sensitivity of the WS2 nanoflowers to NH3 concentrations ranging from 240 to 958 ppm. Relative variation in the conductance to NH3 concentrations ranging from 240 to 958 ppm. The solid red line represents the linear fit to the experimental data (black squares). The slope of the fit is the sensitivity of the WS2 nanoflower-based sensor.
Figure 9. Response of the WS2 nanoflower-based sensor at RT. (A) Conductance under dry N2 conditions (black curve) and after exposure to 958 ppm of NH3 (red curve) as a function of time. (B) Sensitivity of the WS2 nanoflowers to NH3 concentrations ranging from 240 to 958 ppm. Relative variation in the conductance to NH3 concentrations ranging from 240 to 958 ppm. The solid red line represents the linear fit to the experimental data (black squares). The slope of the fit is the sensitivity of the WS2 nanoflower-based sensor.
Processes 09 01491 g009
Figure 10. Response of the WS2 nanoflowers to 1.5 µL of each analyte at 25% RH.
Figure 10. Response of the WS2 nanoflowers to 1.5 µL of each analyte at 25% RH.
Processes 09 01491 g010
Figure 11. Response and recovery curve for WS2 sensor to 240 ppm of NH3 in ambient atmosphere.
Figure 11. Response and recovery curve for WS2 sensor to 240 ppm of NH3 in ambient atmosphere.
Processes 09 01491 g011
Figure 12. Schematic illustration of the interfacial interaction of NH3 molecules with the surface and interlayer of WS2. The illustration is adapted from Qin, et al. [11].
Figure 12. Schematic illustration of the interfacial interaction of NH3 molecules with the surface and interlayer of WS2. The illustration is adapted from Qin, et al. [11].
Processes 09 01491 g012
Figure 13. Response of the WS2 sensor towards 240 ppm of NH3 at room temperature under various humidity conditions (25, 41, 62, 75, 87 and 97%).
Figure 13. Response of the WS2 sensor towards 240 ppm of NH3 at room temperature under various humidity conditions (25, 41, 62, 75, 87 and 97%).
Processes 09 01491 g013
Table 1. Summary of the atomic composition stoichiometric assignments obtained from the fitting of the XPS spectra of WS2 nanoflowers.
Table 1. Summary of the atomic composition stoichiometric assignments obtained from the fitting of the XPS spectra of WS2 nanoflowers.
ElementAtomic %Peak Binding Energy (eV)AssignmentsPeak Area %
C77.4284.8C-C91
286.3C-O6
288.9O-C=O3
O10.8532.2C-O100
W3.730.9WS241
33.2WS238
34.4W5+6
36.3WO315
S6.2160.5S2W def.9
161.6S2W60
162.8S2W31
Table 2. Specificity values of the sensor to each chemical vapor.
Table 2. Specificity values of the sensor to each chemical vapor.
AnalyteConcentration (ppm)Specificity
Acetone5020.086
Ammonia2400.23
Chloroform4440.031
Ethanol6320.026
Toluene3520.045
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gqoba, S.S.; Rodrigues, R.; Mphahlele, S.L.; Ndala, Z.; Airo, M.; Fadojutimi, P.O.; Hümmelgen, I.A.; Linganiso, E.C.; Moloto, M.J.; Moloto, N. Hierarchical Nanoflowers of Colloidal WS2 and Their Potential Gas Sensing Properties for Room Temperature Detection of Ammonia. Processes 2021, 9, 1491. https://doi.org/10.3390/pr9091491

AMA Style

Gqoba SS, Rodrigues R, Mphahlele SL, Ndala Z, Airo M, Fadojutimi PO, Hümmelgen IA, Linganiso EC, Moloto MJ, Moloto N. Hierarchical Nanoflowers of Colloidal WS2 and Their Potential Gas Sensing Properties for Room Temperature Detection of Ammonia. Processes. 2021; 9(9):1491. https://doi.org/10.3390/pr9091491

Chicago/Turabian Style

Gqoba, Siziwe S., Rafael Rodrigues, Sharon Lerato Mphahlele, Zakhele Ndala, Mildred Airo, Paul Olawale Fadojutimi, Ivo A. Hümmelgen, Ella C. Linganiso, Makwena J. Moloto, and Nosipho Moloto. 2021. "Hierarchical Nanoflowers of Colloidal WS2 and Their Potential Gas Sensing Properties for Room Temperature Detection of Ammonia" Processes 9, no. 9: 1491. https://doi.org/10.3390/pr9091491

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