Journal Pre-proof Facile synthesis of ultrafine rGO/WO3 nanowire nanocomposites for highly sensitive toxic NH3 gas sensors Chu Manh Hung, Do Quang Dat, Nguyen Van Duy, Vu Van Quang, Nguyen Van Toan, Nguyen Van Hieu, Nguyen Duc Hoa
PII:
S0025-5408(19)32353-0
DOI:
https://doi.org/10.1016/j.materresbull.2020.110810
Reference:
MRB 110810
To appear in:
Materials Research Bulletin
Received Date:
13 September 2019
Revised Date:
18 January 2020
Accepted Date:
28 January 2020
Please cite this article as: Hung CM, Dat DQ, Van Duy N, Van Quang V, Van Toan N, Van Hieu N, Hoa ND, Facile synthesis of ultrafine rGO/WO3 nanowire nanocomposites for highly sensitive toxic NH3 gas sensors, Materials Research Bulletin (2020), doi: https://doi.org/10.1016/j.materresbull.2020.110810
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2019 Published by Elsevier.
Facile synthesis of ultrafine rGO/WO3 nanowire nanocomposites for highly sensitive toxic NH3 gas sensors Chu Manh Hung1**, Do Quang Dat1,2, Nguyen Van Duy1, Vu Van Quang1, Nguyen Van Toan1, Nguyen Van Hieu3,4, Nguyen Duc Hoa1* 1
International Training Institute for Materials Science (ITIMS), Hanoi University of Science and Technology (HUST), No 1 - Dai Co Viet Str. Hanoi, Vietnam Department of Natural Science, Hoa Lu University, Ninh Nhat commune, Ninh Binh city, Ninh
of
2
Binh province, Vietnam
Faculty of Electrical and Electronic Engineering, Phenikaa Institute for Advanced Study
ro
3
4
-p
(PIAS), Phenikaa University, Yen Nghia, Ha-Dong district, Hanoi 10000, Viet Nam; Phenikaa Research and Technology Institute (PRATI), A&A Green Phoenix Group, 167 Hoang
re
Ngan, Hanoi 10000, Viet Nam.
lP
Corresponding authors
Jo
ur na
* Nguyen Duc Hoa, PhD., Assoc. Professor ** Chu Manh Hung, PhD. International Training Institute for Materials Science (ITIMS), Hanoi University of Science and Technology (HUST) No. 1, Dai Co Viet Road, Hanoi, Vietnam Phone: 84 24 38680787 84 24 38692963 Fax:
[email protected] (N D Hoa)
[email protected] (C M Hung) E-mail: Post address:
No. 1, Dai Co Viet Road, Hanoi, Vietnam
1
Graphical abstract SMO
(A) Graphite
SMO
Adding Metal salts & Hydrothermal synthesis
(B) Graphene oxide
(C) Reduce Graphene oxide/SMO W eV0=m -n CB
B
EC Fn
FG
of
Chemical exfoliation
EV VB
11
10 o
@ 100 ppm & 300 C
8 6 4 1.8
2 NH3
CO
1.1 H2
Tested Gas
1.3
SO2
ur na
lP
re
0
-p
S =(Ra/Rg)
12
ro
(D) Graphene/SMO heterojunction 14
Highlights
Jo
Nanocomposites of rGO and WO3 nanowires were hydrothermally synthesized The rGO/WO3 nanocomposites showed excellent sensing performance with theoretical detection limit of 138 ppb
Gas-sensing mechanism was discussed based on the p-n heterojunction The rGO/WO3 nanocomposites are capable for monitoring highly toxic NH3 gas in air
2
Abstract: We introduce a facile and scalable synthesis of rGO/WO3 nanocomposites by hydrothermal method for gas sensing applications. The characterization of rGO/WO3 nanocomposites by some advanced techniques such as scanning electron microscopy, high resolution transmission electron microscopy and Raman spectroscopy revealed that high quality rGO/WO3 nanocomposites
of
included single crystal WO3 nanowires (average diameter of 10 nm) entangled by thin rGO layers.
ro
Sensing measurements demonstrated that the rGO/WO3 nanocomposite-based sensor can detect highly toxic NH3 gas at low concentrations ranging from 20 ppm to 500 ppm and fulfills practical
-p
applications. The developed gas sensors based on rGO/WO3 nanocomposites have significant
re
application prospects in environment pollution monitoring at detection limit of about 138 ppb. We also discussed the gas sensing mechanism of the rGO/WO3 nanocomposites based on the p-n
lP
junction.
Jo
ur na
Keywords: WO3 nanowires; rGO; Nanocomposites; Hydrothermal; Highly toxic NH3, Gas sensors
3
1. Introduction Since its first invention [1], resistive gas sensor has attracted massive attention from researchers worldwide [2–7] because it has large potential application in various fields [8], such as chemical processing, agriculture, environmental monitoring, gas exhaust control, industrial process control, military, security, smart house, food analysis, and lung cancer diagnosis [9,10].
of
Attempts have been made to fabricate novel nanomaterials [11] for gas sensing devices with high sensitivity and selectivity for practical applications in modern society [12–14]. Metal oxide-based
ro
gas sensors [15] are prospective candidates, which have been extensively studied due to their low
-p
cost, compact size, real-time detection, portability, and low power consumption [16,17]. Among the materials used as gas sensitive layers, pristine WO3 material with different morphologies is
re
used as a gas sensor because of its high sensitivity to NO2 [18,19]. Response value of WO3 nanowires to NH3 1500 ppm at 250 oC is only 9.7 [20]. Recently, WO3 synthesized by
lP
hydrothermal method and decorated with Pd nanoparticles has been found to enhance gas sensing performance [21]. However, the pristine metal oxide-based gas sensors have low sensitivity and
ur na
poor selectivity to some certain gases with low signal to noise due to its low conductivity [22–24]. In addition to metal oxides, 2D materials, such as graphene [25] and reduced oxide graphene (rGO) [26], have recently been used as sensing materials for the detection of hazardous and
Jo
explosive gases such as H2S, NO2, H2, and NH3, due to their large specific surface area and tunable electrical properties [27–29]. However, pristine graphene and nanocarbon-based sensors suffer from the slow response and recovery characteristics [27], thereby limiting their potential application [28]. Recently, the hybrids or composites of metal oxide and carbon allotropes [29] have been extensively studied as gas-sensitive materials and reported to exhibit significantly high sensitivity [30–32]. The composites of metal oxide and carbon allotropes like rGO were used to 4
enhance the gas sensing ability of sensors [33,34] by utilizing the heterojunction between nanocarbon and metal oxides [35,36]. Therefore, many efforts have focused on the synthesis of rGO and metal oxide composites for gas sensor applications [37]. For instance, the rGO/hexagonal WO3 nanosheets composites were synthesized for enhanced H2S sensing performance [38], whereas the WO3 nanorods/graphene nanocomposites were prepared for NO2 gas sensing
of
application [39]. The rGO/WO3 nanolamellae nanocomposites were prepared by hydrothermal method for acetone sensing application [40], while the Pd–WO3/rGO hierarchical nanostructures
ro
were used as efficient H2 gas sensors [41]. Composites of rGO and WO3 nanoparticles were also prepared by ultrasonicating WO3 nanospheres and rGO layers for the NH3 sensor [33]. It is clearly
-p
that the gas sensing characteristics of rGO based nanocomposites are strongly dependent on the
re
morphology and the size of WO3 nanostructures. The smaller diameter of WO3 is expected to show higher sensitivity [42]. However, none of the report dedicating on the synthesis and application of
lP
WO3 nanowires/rGO for gas sensors despite the nanowires were reported to exhibit better sensing performance. Advantages of WO3 nanowires/rGO include the small diameter of WO3 nanowires
ur na
which is comparable with the Debye length, and the p-n junction between rGO and WO3, thus enhancing the gas sensing performance [40]. Herein, we introduce our study on the synthesis of rGO/WO3 nanowire composites for gas sensor application. The design of gas sensor based on rGO/WO3 nanowire composites is shown in
Jo
Figure 1(A). The nanocomposites of rGO and WO3 nanowires were prepared by a green facile and scalable hydrothermal method for gas sensor applications. The advantages of this method include simple synthesis and good dispersion of rGO and WO3 nanowires. Given that the hydrothermal synthesis of WO3 requires low pH, we used ascorbic acid (C6H8O6) as a reducing agent to reduce GO into rGO, control the pH, and form ultrafine WO3 nanowires. Under these synthesis conditions,
5
the surface of rGO was highly functionalized and could be easily dispersed in solution for sensor fabrication. Homogenous WO3 nanowires with an average diameter of 10 nm were obtained, and the rGO layers were entangled over the WO3 nanowires. The gas sensing properties of the synthesized materials were tested using various gases, including of SO2, NH3, CO, and H2. We also discussed the gas sensing mechanism of the synthesized rGO/WO3 nanocomposites based on
of
the p-n heterojunctions.
ro
2. Experimental
The materials used in this study were graphite, NaNO3, H2SO4, KMnO4, H2O2, HCl,
-p
Na2WO4.2H2O, NaCl, C6H8O6, C2H5OH, and deionized (DI) water. All analytical grade reagents
re
were used without further purification. 2.1 rGO/WO3 nanowire nanocomposite synthesis
lP
Graphene oxide (GO) was synthesized using a modified Hummers method, as reported elsewhere [43]. Detail about the synthesis of GO is summarized in Fig. S1 (Supplementary). The
ur na
nanocomposites of rGO/WO3 nanowires were synthesized by a facile hydrothermal method using as-prepared GO, Na2WO4.2H2O and C6H8O6 as precursors. Schematic for the synthesis of the nanocomposites of rGO/WO3 nanowires is shown in Figure 1(B). In a typical synthesis, 1000 µL of the GO solution synthesized above was dispersed in a mixture of 80 g of DI water, 1.5 g of
Jo
Na2WO4.2H2O, 1.0 g of NaCl, and 2.5 g of C6H8O6 via ultrasonication. Herein, we used C6H8O6 as reducing and pH control agent. The mixture was further stirred for 15 min at room temperature, and then pH of the solution was adjusted to 2. The mixture solution was transferred into a 100 mL Teflon lined autoclave for hydrothermal treatment at 180oC for 12 h. The precipitate product was washed with DI water and ethanol, and then collected by centrifugation at a rate of 5800 rpm for
6
15 min. Finally, the product was dried at 60oC for 24 h before characterization. The crystal structures and morphology of the synthesized materials were characterized by X-ray diffraction (XRD; Advance D8, Bruker), field-emission scanning electron microscopy (SEM, JEOL 7600 F), and high-resolution transmission electron microscopy (TEM, JEM2100). The Raman spectra were studied using LabRAM HR (HORIBA Jobin Yvon (exc = 632.8 nm). Fourier transform infrared
of
(FTIR) spectra were obtained by Thermo Nicolet Nexus 670 Fourier Transform Infrared Spectroscopy. UV-vis spectra were obtained in the wavelength range 190–900nm by using a
-p
2.2 rGO/WO3 nanocomposite-based sensor fabrication
ro
spectrophotometer (PG-T90, UK).
Gas sensors were fabricated via a thick film technology as reported elsewhere [14]. In brief, the
re
synthesized rGO/WO3 nanocomposite was dispersed in solution and then coated onto a thermally
lP
oxidized Si substrate equipped with a pair of interdigitated Pt electrodes. The synthesized sensors were heat treated at 400oC for 2h in air to remove the binder before gas sensing measurements.
ur na
The gas sensing properties were measured at different temperatures under dynamic conditions, where the dry air and analytic gas was continuously flowed through the chamber while the resistance was recorded using a Keithley instrument (Model 2602). The sensor was placed on a heating plate to control the working temperatures of approximately 250, 300, 350, and 400 °C. The
Jo
gas sensing characteristics were tested for NH3, H2, CO and SO2 at various temperatures. The sensor response (S) is defined by the ratio of Ra/Rg (for reducing gases) or Rg/Ra (for oxidizing gases), where Ra and Rg are the resistances of the sensors in dry air and in the tested gas, respectively.
7
3. Results and discussion 3.1 Characterization of materials The morphology of the synthesized rGO/WO3 nanocomposites studied by SEM is shown in Figure 2. The low-magnification SEM image (Figure 2(A)) revealed that the sample was composed of rGO flakes and WO3 nanowire bundles. Given that the amount of rGO was low (less than 1% wt.),
of
the sample mainly looked like bundles of WO3. The average length of the WO3 bundles is about 1
ro
µm. The high-resolution SEM image (Figure 2(B)) demonstrated that WO3 bundles were composed of very fine nanowires with an average diameter of less than 10 nm. Given that the
-p
length of the WO3 nanowires in a bundle differed, the bundle size was small at the top but large at the center. In our previous study, microwheels composed of WO3 nanorods, which were
re
approximately 1.5 μm long and 10 nm in diameter, were obtained by hydrothermal method at
lP
temperature of 200oC [44]. In the study by Xiong et al. [45], nanowire bundle-like WO3-W18O49 was obtained through solvothermal method using WCl6 as precursor and isopropanol as reaction medium. The nanowire WO3 bundle was obtained without using isopropanol as reaction medium.
ur na
Compared with the hexagonal WO3 nanorods prepared by hydrothermal method reported in the literature [21], the diameter of our product was considerably smaller, which was advantageous to enhance gas sensing performance.
Jo
The crystal structures of the calcinated rGO/WO3 nanocomposites studied by XRD is shown in Figure 3(A). The XRD pattern showed the typical diffraction peaks of hexagonal structure, and those peaks were consistent with the standard card of WO3 (JCPDS, No. 33-1387), with lattice constants a = b = 0,7298 nm and c = 0,3899 nm [44]. These findings indicated that the hexagonal
8
structure of WO3 was stable with thermal treatment up to temperature of 400oC. This result was consistent with other report, where the authors pointed out that the XRD pattern of hexagonal WO3 remains nearly the same up to 450°C [46]. This phenomenon is crucial for gas sensing application because the high working temperature of the device does not influence to the stability of the sensor. In addition, no impurity peaks were observed in the XRD pattern, suggesting the high purity of the
of
hexagonal WO3 phase. The diffraction peaks with large and sharp intensity showed that the hexagonal phase of the WO3 crystal in the synthesized samples exhibited high crystallinity. This
ro
result was consistent with other reports on the hydrothermal synthesis of hexagonal rGO/WO3 nanosheets [38]. However, Figure 3(A) shows no obvious diffraction peak of rGO in the XRD
-p
pattern due to small amount of rGO (less than 1 wt.%) in the synthesized nanocomposites. Further
re
characterization by Raman spectroscopy, TEM images, and FTIR spectroscopy was performed to confirm the presence of rGO in the obtained nanocomposites.
lP
Figure 3(B) reveals the Raman spectrum of the rGO/WO3 nanocomposites. In addition to the vibrational modes between W and O (WO and WOW) of the crystalline hexagonal-WO3
ur na
nanowires in the range of 1501050 cm-1 [47,48], the spectrum exhibited two distinct Raman modes centered at approximately 1330 and 1593 cm-1, which correspond to the vibrational modes of the D and Gbands of rGO, respectively. The Dband originated from the vibration of the
Jo
CC bond induced by the disorder [49], whereas the Gband was attributed to the first-order scattering of the E2g vibration mode of sp2 carbon domains [48]. These results confirmed the existence of rGO in the synthesized nanocomposites. Intensity ratio between the D and Gband peaks (ID/IG) is often used as a parameter to estimate the defects and disorders of carbonaceous materials [48,50]. The calculated ID/IG intensity ratio (Figure 2(B)) was approximately 1.09, which was considerably smaller than that of rGO/WO3 nanosheets (1.28) synthesized through the 9
hydrothermal method in a previous report [48]. Thus, less carbonaceous defects and disorders were generated in the synthesized rGO presented in the current work. In our study, we used ascorbic acid as pH adjusting and reducing agent, thus during hydrothermal process the ascorbic further reduced GO to enhance the quality of rGO and decreased the ID/IG intensity ratio.
of
The representative TEM images of the rGO/WO3 nanocomposites are displayed in Figure 4. The
ro
low-magnification TEM image of Figure 4(A) revealed extremely thin WO3 nanowires of approximately 10 nm, which was exfoliated from the bundles of rGO/WO3 nanowires (as shown
-p
in the SEM image). Curled and corrugated rGO sheets were also observed in the TEM images. The extremely thin rGO sheets appeared to cover the WO3 nanowires, thereby enhancing the gas
re
sensing characteristics of the rGO/WO3 nanocomposites [51]. High-resolution TEM image of the
lP
rGO/WO3 nanocomposites (Figure 4(B)) showed homogeneous lattice structures along the WO3 nanowires. The d-spacing of the lattice planes was calculated to be approximately 0.39 nm, which
ur na
corresponded to the (001) lattice plane of WO3 [52].
The optical absorption properties and band gap energy of the rGO/WO3 nanocomposites were characterized by UV-vis diffuse reflectance spectroscopy (DRS). Figure 5(A) shows the UV-Vis
Jo
spectrum of the rGO/WO3 nanocomposites in the spectral range of 200800 nm, and the Tauc plot (inset graph) shows the effect of ℎ𝜈 on the corresponding (𝛼ℎ𝜈)1/2 . The UV-vis DRS spectrum revealed a characteristics absorption band at 390 nm, indicating that WO3 with high crystalline structure was formed in the rGO/WO3 nanocomposites [53]. The absorption intensity of the synthesized rGO/WO3 nanocomposites in the visible region from 400 nm to 800 nm was higher
10
than that of the pristine WO3 nanowires grown via a similar method as reported by Nagy et al. [54]. This result was due to the presence of rGO in the synthesized nanocomposites, which formed additional defects in the WO3 crystals. The absorption edge of the rGO/WO3 nanocomposites showed a blue shift compared with the pristine WO3 nanowires [54]; this result was in good agreement with a previous report [52]. The band gap energy of the rGO/WO3 nanocomposites was
of
estimated from the UV-vis spectrum using the following expression [55,56]: (𝛼ℎ𝜈)1/2 = 𝐾(ℎ𝜈 − 𝐸𝑔 ),
(1)
ro
where is the absorption coefficient, ℎ is Planck’s constant, 𝜈 is the incident photon’s frequency,
-p
K is a proportionality constant, and Eg is the band gap energy. The inset in Figure 5(A) displays the linear extrapolation of the Tauc plot in which the X-axis is ℎ𝜈 and the Y-axis is the Tauc term
re
((𝛼ℎ𝜈)1/2 ) in the case of indirect band gap rGO/WO3 nanocomposites. The linear extrapolation
lP
intercepts the X-axis at a certain value, which is generally known as the optical band-gap energy [55,56]. Therefore, the band gap energy of rGO/WO3 nanocomposites was approximately 2.9 eV, which was consistent with that of rGO-WO3 nanorod composites synthesized by the one-pot
ur na
hydrothermal method in a previous work [55].
The functional groups in rGO and rGO/WO3 nanocomposites were investigated by FTIR spectroscopy. Figure 5(B) shows the FTIR spectra of rGO and the synthesized rGO/WO3
Jo
nanocomposites. The FTIR spectrum of rGO (on top) showed characteristic absorption peaks at 3427.7, 1728.9, 1583.5, 1225.4, 1046.7, and 585.6 cm-1, which corresponded to the OH, CO, CC, phenolic COH, alkoxy CO, and epoxy CO stretching vibrations of rGO, respectively [57,58]. However, only the peaks of OH and CC stretching vibrations were not observed in the FTIR spectrum of rGO/WO3 nanocomposites (at the bottom). It indicates a reduction of oxygen-
11
containing functional groups of the rGO/WO3 nanocomposites caused by the thermal treatment. The results imply an occurrence of the interaction between rGO and WO3, which may enhance the electron transfer, and therefore improve the gas sensing properties of the rGO/WO3 nanocomposites [59]. In addition, the presence of a broad absorption peak at 820 cm-1 in the spectrum of the rGO/WO3 nanocomposites was ascribed to the WOW stretching mode in the
nanosheets in a previous report [38].
ro
of
ultrafine WO3 nanowires, which is consistent with the stretching mode of the crystalline WO3
-p
3.2 Gas sensing characteristics
re
The gas sensing properties of the rGO/WO3 nanowire composites were tested at 20500 ppm NH3 in a temperature range of 300450 °C. Figure 6(A) shows the transient resistances of the rGO/WO3
lP
compositebased sensor at different temperatures. The data presented a decrease in sensor resistance with increasing working temperature. This occurrence was attributed to the fact that the
ur na
density of carrier electron, and the electron mobility degree in the rGO/WO3 nanocomposites increased with increasing the working temperature, which was in good agreement with the general characteristics of semiconductor materials [60]. Moreover, the rGO/WO3 nanocomposite sensor revealed a typical n-type gas sensing response upon exposure to NH3 reducing gas within the
Jo
temperature range. This observation was consistent with the n-type semiconducting property of pristine WO3 materials [61]. The gas responses of the rGO/WO3 nanocomposite sensor at different temperature as a function of NH3 gas concentration are shown in Figure 6(B). The gas response of the sensor increased with increasing NH3 concentration from 20 ppm to 500 ppm. At high NH3 concentrations, more gas 12
molecules participated in the adsorption, diffusion, and reaction processes, resulting in elevated gas sensor response. The increase in the sensor response was rather linear with the increase in gas concentration; this factor is important to evaluate potential application in practical devices for NH3 monitoring [62]. Figure 6(B) also shows that the gas response of the rGO/WO3 nanocomposite sensor increased as the working temperature decreased. The highest sensing performance was achieved at 300 °C, and the response value to 500 ppm NH3 gas was 35. At a low gas concentration
of
of 250 ppm and operating temperature of 300 °C, the NH3 gas response of the sensor also reached
ro
the value of approximately 16, which is about 5 times higher than that of the sensor based on the Fe2O3/WO3 nanorod composites [63]. As the working temperature of the rGO/WO3
-p
nanocomposite sensor further decreased, the NH3 gas response decreased and the response and
re
recovery times were very long, which hampered the practical applications of devices (Fig. S2,
measurements.
lP
Supplementary). Thus, temperatures ranging from 300 °C to 450 °C were selected for gas sensing
Figures 6(C) and (D) show the response and recovery times of the sensors as function of NH3 gas
ur na
concentrations at different working temperatures. As shown in Figure 6(C), at all working temperatures, the response time of the sensor decreased as the NH3 gas concentration increased. At high gas concentration, more NH3 molecules were absorbed on the active sites of the sensing
Jo
layers, leading to the rapid saturation of sensor resistance; hence, the response time of the sensor was shortened. However, a longer time is needed for the sensor to recover to its initial state after stopping exposure to high gas concentration, as shown in Figure 6(D). It is clearly attributed to the fact that at high gas concentration, more NH3 molecules can be absorbed on the active sites of the sensing layers, leading to quickly reach to the saturated resistance of the sensor. Thus, the response time of the sensor was shorter. Whereas, longer time 13
was needed to recover to the initial state after stopped exposure of the sensor to high gas concentration for all working temperatures as shown in Figure 6(D). In addition, the response/recovery times increased with decreasing working temperature; at 300 °C, the response and recovery times were approximately 37 and 711 s, respectively.
of
ro
Table 1 summarizes the recent studies on the NH3 sensing performance of sensors based on different WO3 materials and composites. The sensors based on pristine WO3 demonstrated low
-p
sensitivity to NH3 gas. For instance, the WO3 particles prepared by thermal decomposition of
re
ammonium wolframate showed the response value of 6 to 1000 ppm NH3 at 300 oC [64], while the WO3 nanowire prepared by sputtering and calcination using carbon nanotubes as templates
lP
exhibited the response value of 9.7 to 1500 ppm NH3 at 250 oC [20]. Table 1 clearly documents that the NH3 gas sensing performance of the sensors improved with surface decoration, doping,
ur na
and/or formation of composite compounds with other materials. The rGO/WO3 nanoparticles prepared by ultrasonication showed the response value of 15.83 to 100 ppm NH3 at room temperature [33], whereas the rGO/WO3 nanocomposite prepared by hydrothermal method showed the response value of 27.7 to 100 ppm NH3 at 150 oC [65]. The rGO/WO3 nanowires
Jo
prepared in this study showed the response value of approximately 11 to 100 ppm NH3, which is higher than the response of GO/WO3 nanorods [66]. Anyhow, on the basis of our study, the rGO/WO3 nanowire composites demonstrated the enhanced NH3 gas response compared to the pristine WO3 nanorods, nanowires, and nanoparticles.
14
Table 1. Comparison of NH3 sensing performance between the present work and other literature reports Materials
NH3
Temp.
(ppm)
(°C)
Response
Detection
Method
Ref.
range
100
350
3.2
20-100
Two-step hydrothermal
[67]
1000
250
25
2-1000
Hydrothermal
[45]
500
300
12.5
25-1000
Reactive sputtering
[68]
Fe doped WO3 film
20
250
0.15
20
Electron beam evaporation
WO3 nanocrystals
74
325
3.2
74
ro
[69]
Hydrothermal
[70]
WS2/WO3 composites
10
150
4.9
1-10
Oxidation
[71]
Ru loaded WO3 nanosheets
20
300
17.8
re
(ppm) Closed pack WO3
Acidification with impregnation
[72]
microspheres WO3-W18O49 nanowire
Bilayer SnO2-WO3
-p
nanofilms
of
bundle
lP
process Hydrothermal
WO3 nanowires
1500
250
9.7
300-1500
Sputtering & Calcination
[20]
WO3 particles
1000
300
6
10-1000
Thermal decomposition
[64]
rGO/WO3 nanoparticles
100
RT
15.83
10-100
ultrasonication
[33]
rGO/WO3 nanocomposite
100
150
27.7
1-100
hydrothermal
[65]
Jo
3-20
GO/WO3 nanorods
100
200
1.17
10-100
Hydrothermal
[66]
rGO/WO3 nanowire
100
300
11
20-500
Hydrothermal
This
Fe2O3/WO3 nanorod
300
composites
composites
100
ur na
Ag-doped -Fe2O3/SiO2
375
4
25-300
[63]
work
15
In addition to the high gas response and fast response/recovery times, good limit of detection (LOD), high gas selectivity, and good stability are also important characteristics of the sensor for practical applications. As mentioned above, the response of the rGO/WO3 nanocomposite sensor increased when the NH3 gas concentration increased. Figure 7 (A) shows the representative response of the sensor to 25500 ppm NH3 at 300 oC, which indicated a linear response to the
of
increase in NH3 concentration. Through linear fitting, the rGO/WO3 nanocomposite sensor exhibited a slope (S) of 0.0615 and coefficient of determination (R2 value) of 0.9916. The LOD of
ro
the rGO/WO3 nanocomposite sensor was calculated using the equation (2), where and S are the root-mean square deviation calculated by the fifth-order polynomial fit of base response versus
-p
time, and slope value of the linear fit of the gas response versus gas concentration, respectively
re
[73]. Detail about the fifth-order polynomial fit of 15 data points taken on the base of the response curve as a function of time for calculation of detection limit is shown in Figure 7 (B). 3𝜎 𝑆
lP 𝐿𝑂𝐷 =
(2)
ur na
The LOD for the rGO/WO3 nanocomposite sensor was calculated to be 138 ppb, which demonstrated the potential application sensors with low LOD toward NH3 gas.
Jo
To evaluate gas selectivity, the rGO/WO3 nanocomposite sensor was exposed to 100 ppm CO, H2, and SO2 gases at 300 oC. The transient resistances of the sensors upon exposure to these gases are provided in the supporting information (Figure S35, Supplementary), which showed that the resistance of the sensor decreased and increased upon exposure to reducing gases (CO and H2) and oxidizing gas SO2, respectively. This finding again confirmed the typical n-type sensing behavior of rGO/WO3 nanocomposite sensor under different working temperature ranges, as shown in the 16
supporting information (Figure S35, Supplementary). The highest sensor response to 5100 ppm CO and 25500 ppm H2 was as low as 1.6 and 1.42 at 300 °C and 350 °C, which were similar optimal working temperatures to that of the NH3 sensor based on the rGO/WO3 nanocomposite. By contrast, the optimal working temperature of the rGO/WO3 nanocomposite sensor to 110 ppm SO2 gas was room temperature with a sensor response of only 1.31. For comparison, Figure 8(A)
of
presents the gas responses of the rGO/WO3 nanocomposite sensor to NH3, CO, H2, and SO2 gases under identical working conditions. At the same concentration and working temperature, the
ro
response of the sensor toward NH3 gas was considerably higher than that toward the other gases, indicating good selectivity to NH3. Figure 8(B) illustrates the short-term stability of the rGO/WO3
-p
nanocomposite sensor to NH3 gas at 300 °C with 10 response/recovery cycles. The data showed
re
that the sensor nearly maintained its original response and recovery even after 10 cycles, thereby confirming the good short-term stability and reproducibility of the rGO/WO3 nanocomposite
lP
sensor. In addition, the long-term stability of the sensor was also studied by measuring the NH3 response characteristics of the sensor six months storage in ambient environment and a week
ur na
continuously working at 300oC, as shown in Figure 8(C). The sensor distorted its performance after a week of continuously working at high temperature. Such data confirms that the rGO/WO3 nanocomposite is suitable for disposable sensor for monitoring of toxic NH3 gas.
Jo
Gas sensing mechanism of pristine metal oxide was based on the surface reaction between the preadsorbed oxygen species and analytic gas molecules, which modulated the depletion region and altered sensor resistance [18]. However, herein, rGO/WO3 nanocomposite was used; so the heterojunction formed between rGO and WO3 also determined the gas sensing performance [43]. The gas sensing mechanism of the rGO/WO3 nanocomposites-based sensor is proposed in Figure 17
9 [40]. Hexagonal WO3 is an n-type wide band gap (2.9 eV) semiconductor with a work function of approximately 4.7–6.4 eV [74]. By contrast, rGO is a p-type semiconductor [75] with low work function ranging from 4.2 eV to 4.45 eV [76]. The band gap of rGO can vary from 1.00 eV to 1.69 eV, depending on reduction method [77]. Therefore, when rGO came into contact with WO3 in the form of rGO/WO3 nanocomposites, the p-n junction formed, which was highly sensitive to
of
gas adsorption/desorption [40]. Upon exposure to NH3, the gas molecule further reduced rGO until zero bandgap was reached or metallic graphene formed. The p-n junction was changed to Ohmic
ro
contact, which significantly reduced sensor resistance and improved sensing performance.
-p
re
4. Conclusion
We introduced a facile hydrothermal method of synthesizing high quality rGO/WO3
lP
nanocomposites for gas sensor applications. The nanocomposites composed of single crystal WO3 nanowires (average size of 10 nm) entangled by thin rGO layers. The diameter of the WO3
ur na
nanowires was very small and comparable with the Debye length, which ensured the total depletion of the oxide crystal. Meanwhile, the entanglement of rGO over the nanowires provided additional adsorption sites, thereby enhancing sensing performance. Gas sensing measurements demonstrated that the rGO/WO3 nanocomposite could monitor NH3 at low concentrations with
Jo
detection limit of 138 ppb. The gas sensing mechanism of the of the rGO/WO3 nanocomposites was also discussed based on the heterojunction between highly conductive p-type rGO and n-type semiconducting WO3 and the enhancement of gaseous adsorption sites for maximizing sensing performance.
18
Authors statement: All authors equally contributed to the manuscript. Professors Nguyen Duc Hoa On behalf of all authors
Declaration of interests
ro
of
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
-p
Acknowledgment
Research under Vingroup Innovation Foundation (VINIF) annual research support program in project
lP
Conflicts of Interest
re
code VINIF.2019.DA10.
ur na
The authors declare that they have no conflicts of interest.
References
T. Seiyama, A. Kato, K. Fujiishi, M. Nagatani, A New Detector for Gaseous Components
Jo
[1]
Using Semiconductive Thin Films., Anal. Chem. 34 (1962) 1502–1503. doi:10.1021/ac60191a001.
[2]
A. Mirzaei, S.S. Kim, H.W. Kim, Resistance-based H2S gas sensors using metal oxide nanostructures: A review of recent advances, J. Hazard. Mater. 357 (2018) 314–331. doi:10.1016/j.jhazmat.2018.06.015. 19
[3]
Y. Gui, K. Tian, J. Liu, L. Yang, H. Zhang, Y. Wang, Superior triethylamine detection at room temperature by {-112} faceted WO3 gas sensor, J. Hazard. Mater. (2019) 120876. doi:10.1016/j.jhazmat.2019.120876.
[4]
S. Xu, H. Zhao, Y. Xu, R. Xu, Y. Lei, Carrier Mobility-Dominated Gas Sensing: A RoomTemperature Gas-Sensing Mode for SnO2 Nanorod Array Sensors, ACS Appl. Mater.
S. Xu, Y. Xu, H. Zhao, R. Xu, Y. Lei, Sensitive Gas-Sensing by Creating Adsorption
ro
[5]
of
Interfaces. 10 (2018) 13895–13902. doi:10.1021/acsami.8b03953.
Active Sites: Coating an SnO2 Layer on Triangle Arrays, ACS Appl. Mater. Interfaces. 10
H.-Y. Li, C.-S. Lee, D.H. Kim, J.-H. Lee, Flexible Room-Temperature NH3 Sensor for
re
[6]
-p
(2018) 29092–29099. doi:10.1021/acsami.8b08078.
Ultrasensitive, Selective, and Humidity-Independent Gas Detection, ACS Appl. Mater.
[7]
lP
Interfaces. 10 (2018) 27858–27867. doi:10.1021/acsami.8b09169. A.V. Raghu, K.K. Karuppanan, J. Nampoothiri, B. Pullithadathil, Wearable, Flexible
ur na
Ethanol Gas Sensor Based on TiO2 Nanoparticles-Grafted 2D-Titanium Carbide Nanosheets, ACS Appl. Nano Mater. 2 (2019) 1152–1163. doi:10.1021/acsanm.8b01975. [8]
G. Korotcenkov, Handbook of Gas Sensor Materials, Springer New York, New York, NY,
Jo
2013. doi:10.1007/978-1-4614-7165-3. [9]
S. Schon, S.J. Theodore, A.T. Güntner, Versatile breath sampler for online gas sensor analysis, Sensors Actuators B Chem. 273 (2018) 1780–1785. doi:10.1016/j.snb.2018.07.094.
[10]
A.T. Güntner, S. Abegg, K. Königstein, P.A. Gerber, A. Schmidt-Trucksäss, S.E.
20
Pratsinis, Breath Sensors for Health Monitoring, ACS Sensors. 4 (2019) 268–280. doi:10.1021/acssensors.8b00937. [11]
L.G. Bach, M.L.N. Thi, Q.B. Bui, H.-T. Nhac-Vu, Hierarchical cobalt nanorods shelled with nickel oxide vertically attached 3D architecture as non-binder and free-standing sensor for sensitive non-enzymatic glucose detection, Mater. Res. Bull. 118 (2019)
X. Liu, K. Zhao, X. Sun, C. Zhang, X. Duan, P. Hou, G. Zhao, S. Zhang, H. Yang, R.
ro
[12]
of
110504. doi:10.1016/j.materresbull.2019.110504.
Cao, X. Xu, Rational design of sensitivity enhanced and stability improved TEA gas
-p
sensor assembled with Pd nanoparticles-functionalized In2O3 composites, Sensors
[13]
re
Actuators B Chem. 285 (2019) 1–10. doi:10.1016/j.snb.2019.01.029. Y. Xu, T. Ma, L. Zheng, L. Sun, X. Liu, Y. Zhao, J. Zhang, Rational design of Au/Co3O4-
lP
functionalized W18O49 hollow heterostructures with high sensitivity and ultralow limit for triethylamine detection, Sensors Actuators B Chem. 284 (2019) 202–212.
[14]
ur na
doi:10.1016/j.snb.2018.12.119.
N.D. Hoa, C.M. Hung, N. Van Duy, N. Van Hieu, Nanoporous and crystal evolution in nickel oxide nanosheets for enhanced gas-sensing performance, Sensors Actuators B
Jo
Chem. 273 (2018) 784–793. doi:10.1016/j.snb.2018.06.095. [15]
N.D. Hoa, V. Van Quang, D. Kim, N. Van Hieu, General and scalable route to synthesize nanowire-structured semiconducting metal oxides for gas-sensor applications, J. Alloys Compd. 549 (2013) 260–268. doi:10.1016/j.jallcom.2012.09.051.
[16]
Y. Xu, T. Ma, L. Zheng, Y. Zhao, X. Liu, J. Zhang, Heterostructures of hematitesensitized W18O49 hollow spheres for improved acetone detection with ultralow detection 21
limit, Sensors Actuators B Chem. 288 (2019) 432–441. doi:10.1016/j.snb.2019.03.015. [17]
T.M. Ngoc, N. Van Duy, C.M. Hung, N.D. Hoa, N.N. Trung, H. Nguyen, N. Van Hieu, Ultralow power consumption gas sensor based on a self-heated nanojunction of SnO2 nanowires, RSC Adv. 8 (2018) 36323–36330. doi:10.1039/C8RA06061D.
[18]
N.D. Hoa, S.A. El-Safty, Gas nanosensor design packages based on tungsten oxide:
of
mesocages, hollow spheres, and nanowires, Nanotechnology. 22 (2011) 485503.
[19]
ro
doi:10.1088/0957-4484/22/48/485503.
N.D. Hoa, N. Van Duy, N. Van Hieu, Crystalline mesoporous tungsten oxide nanoplate
-p
monoliths synthesized by directed soft template method for highly sensitive NO2 gas
re
sensor applications, Mater. Res. Bull. 48 (2013) 440–448. doi:10.1016/j.materresbull.2012.10.047.
N. Van Hieu, V. Van Quang, N.D. Hoa, D. Kim, Preparing large-scale WO3 nanowire-like
lP
[20]
structure for high sensitivity NH3 gas sensor through a simple route, Curr. Appl. Phys. 11
[21]
ur na
(2011) 657–661. doi:10.1016/j.cap.2010.11.002. P. Van Tong, N.D. Hoa, N. Van Duy, D.T.T. Le, N. Van Hieu, Enhancement of gassensing characteristics of hydrothermally synthesized WO3 nanorods by surface
Jo
decoration with Pd nanoparticles, Sensors Actuators B Chem. 223 (2016) 453–460. doi:10.1016/j.snb.2015.09.108.
[22]
A. Staerz, S. Somacescu, M. Epifani, T. Russ, U. Weimar, N. Barsan, WO3 Based Gas Sensors, Proceedings. 2 (2019) 826. doi:10.3390/proceedings2130826.
[23]
R. Godbole, V.P. Godbole, P.S. Alegaonkar, S. Bhagwat, Effect of film thickness on gas
22
sensing properties of sprayed WO 3 thin films, New J. Chem. 41 (2017) 11807–11816. doi:10.1039/C7NJ00963A. [24]
N. Van Duy, N.D. Hoa, N.T. Dat, D.T.T. Le, N. Van Hieu, Ammonia-Gas-Sensing Characteristics of WO3/Carbon Nanotubes Nanocomposites: Effect of Nanotube Content and Sensing Mechanism, Sci. Adv. Mater. 8 (2016) 524–533. doi:10.1166/sam.2016.2716. V. Van Quang, N.S. Trong, N.N. Trung, N.D. Hoa, N. Van Duy, N. Van Hieu, Full-Layer
of
[25]
ro
Controlled Synthesis and Transfer of Large-Scale Monolayer Graphene for Nitrogen Dioxide and Ammonia Sensing, Anal. Lett. 47 (2014) 280–294.
N. Sharma, V. Sharma, Y. Jain, M. Kumari, R. Gupta, S.K. Sharma, K. Sachdev,
re
[26]
-p
doi:10.1080/00032719.2013.832270.
Synthesis and Characterization of Graphene Oxide (GO) and Reduced Graphene Oxide
lP
(rGO) for Gas Sensing Application, Macromol. Symp. 376 (2017) 1700006. doi:10.1002/masy.201700006.
Z. Xiao, L.B. Kong, S. Ruan, X. Li, S. Yu, X. Li, Y. Jiang, Z. Yao, S. Ye, C. Wang, T.
ur na
[27]
Zhang, K. Zhou, S. Li, Recent development in nanocarbon materials for gas sensor applications, Sensors Actuators B Chem. 274 (2018) 235–267.
Jo
doi:10.1016/j.snb.2018.07.040. [28]
A.R. Cadore, E. Mania, A.B. Alencar, N.P. Rezende, S. de Oliveira, K. Watanabe, T. Taniguchi, H. Chacham, L.C. Campos, R.G. Lacerda, Enhancing the response of NH3 graphene-sensors by using devices with different graphene-substrate distances, Sensors Actuators B Chem. 266 (2018) 438–446. doi:10.1016/j.snb.2018.03.164.
[29]
M.Y. Lone, A. Kumar, N. Ansari, S. Husain, M. Zulfequar, R.C. Singh, M. Husain, 23
Structural effect of SWCNTs grown by PECVD towards NH3 gas sensing and field emission properties, Mater. Res. Bull. 119 (2019) 110532. doi:10.1016/j.materresbull.2019.110532. [30]
N.D. Hoa, N. Van Quy, D. Kim, Nanowire structured SnOx–SWNT composites: High performance sensor for NOx detection, Sensors Actuators B Chem. 142 (2009) 253–259.
N.D. Hoa, N. Van Quy, M. An, H. Song, Y. Kang, Y. Cho, D. Kim, Tin-Oxide Nanotubes
ro
[31]
of
doi:10.1016/j.snb.2009.07.053.
for Gas Sensor Application Fabricated Using SWNTs as a Template, J. Nanosci.
P.-G. Su, S.-L. Peng, Fabrication and NO2 gas-sensing properties of reduced graphene
re
[32]
-p
Nanotechnol. 8 (2008) 5586–5589. doi:10.1166/jnn.2008.1387.
oxide/WO3 nanocomposite films, Talanta. 132 (2015) 398–405.
[33]
lP
doi:10.1016/j.talanta.2014.09.034.
G. Jeevitha, R. Abhinayaa, D. Mangalaraj, N. Ponpandian, P. Meena, V. Mounasamy, S.
ur na
Madanagurusamy, Porous reduced graphene oxide (rGO)/WO3 nanocomposites for the enhanced detection of NH3 at room temperature, Nanoscale Adv. (2019). doi:10.1039/C9NA00048H.
S. Gupta Chatterjee, S. Chatterjee, A.K. Ray, A.K. Chakraborty, Graphene–metal oxide
Jo
[34]
nanohybrids for toxic gas sensor: A review, Sensors Actuators B Chem. 221 (2015) 1170– 1181. doi:10.1016/j.snb.2015.07.070.
[35]
Q.T. Minh Nguyet, N. Van Duy, N.T. Phuong, N.N. Trung, C.M. Hung, N.D. Hoa, N. Van Hieu, Superior enhancement of NO2 gas response using n - p - n transition of carbon nanotubes/SnO2 nanowires heterojunctions, Sensors Actuators B Chem. 238 (2017) 1120– 24
1127. doi:10.1016/j.snb.2016.07.143. [36]
V. Van Quang, N. Van Dung, N. Sy Trong, N. Duc Hoa, N. Van Duy, N. Van Hieu, Outstanding gas-sensing performance of graphene/SnO2 nanowire Schottky junctions, Appl. Phys. Lett. 105 (2014) 13107. doi:10.1063/1.4887486.
[37]
S. Zhang, B. Zhang, G. Sun, Y. Li, B. Zhang, Y. Wang, J. Cao, Z. Zhang, One-step
of
synthesis of Ag/SnO2/rGO nanocomposites and their trimethylamine sensing properties,
[38]
ro
Mater. Res. Bull. 114 (2019) 61–67. doi:10.1016/j.materresbull.2019.02.019.
J. Shi, Z. Cheng, L. Gao, Y. Zhang, J. Xu, H. Zhao, Facile synthesis of reduced graphene
-p
oxide/hexagonal WO3 nanosheets composites with enhanced H2S sensing properties,
[39]
re
Sensors Actuators B Chem. 230 (2016) 736–745. doi:10.1016/j.snb.2016.02.134. X. An, J.C. Yu, Y. Wang, Y. Hu, X. Yu, G. Zhang, WO3 nanorods/graphene
lP
nanocomposites for high-efficiency visible-light-driven photocatalysis and NO2 gas sensing, J. Mater. Chem. 22 (2012) 8525. doi:10.1039/c2jm16709c. J. Kaur, K. Anand, K. Anand, R.C. Singh, WO3 nanolamellae/reduced graphene oxide
ur na
[40]
nanocomposites for highly sensitive and selective acetone sensing, J. Mater. Sci. 53 (2018) 12894–12907. doi:10.1007/s10853-018-2558-z. A. Esfandiar, A. Irajizad, O. Akhavan, S. Ghasemi, M.R. Gholami, Pd–WO3/reduced
Jo
[41]
graphene oxide hierarchical nanostructures as efficient hydrogen gas sensors, Int. J. Hydrogen Energy. 39 (2014) 8169–8179. doi:10.1016/j.ijhydene.2014.03.117.
[42]
P. Van Tong, N.D. Hoa, D.D. Trung, N.D. Quang, N. Van Hieu, Tungsten Oxide UrchinFlowers and Nanobundles: Effect of Synthesis Conditions and Heat Treatment on
25
Assembly and Gas-Sensing Characteristics, Sci. Adv. Mater. 6 (2014) 1081–1090. doi:10.1166/sam.2014.1854. [43]
N. Van Hoang, C.M. Hung, N.D. Hoa, N. Van Duy, I. Park, N. Van Hieu, Excellent detection of H2S gas at ppb concentrations using ZnFe2O4 nanofibers loaded with reduced graphene oxide, Sensors Actuators B Chem. 282 (2019) 876–884.
P. Van Tong, N.D. Hoa, N. Van Duy, N. Van Hieu, Micro-wheels composed of self-
ro
[44]
of
doi:10.1016/j.snb.2018.11.157.
assembled tungsten oxide nanorods for highly sensitive detection of low level toxic
Y. Xiong, Z. Zhu, T. Guo, H. Li, Q. Xue, Synthesis of nanowire bundle-like WO3-W18O49
re
[45]
-p
chlorine gas, RSC Adv. 5 (2015) 25204–25207. doi:10.1039/C5RA00916B.
heterostructures for highly sensitive NH3 sensor application, J. Hazard. Mater. 353 (2018)
[46]
lP
290–299. doi:10.1016/j.jhazmat.2018.04.020.
I.M. Szilágyi, J. Pfeifer, C. Balázsi, A.L. Tóth, K. Varga-Josepovits, J. Madarász, G.
ur na
Pokol, Thermal stability of hexagonal tungsten trioxide in air, J. Therm. Anal. Calorim. 94 (2008) 499–505. doi:10.1007/s10973-007-8601-y. [47]
G.L. Frey, A. Rothschild, J. Sloan, R. Rosentsveig, R. Popovitz-Biro, R. Tenne,
Jo
Investigations of Nonstoichiometric Tungsten Oxide Nanoparticles, J. Solid State Chem. 162 (2001) 300–314. doi:10.1006/JSSC.2001.9319.
[48]
J. Shi, Z. Cheng, L. Gao, Y. Zhang, J. Xu, H. Zhao, Facile synthesis of reduced graphene oxide/hexagonal WO3 nanosheets composites with enhanced H2S sensing properties, Sensors Actuators, B Chem. 230 (2016) 736–745. doi:10.1016/j.snb.2016.02.134.
26
[49]
S. Srivastava, K. Jain, V.N. Singh, S. Singh, N. Vijayan, N. Dilawar, G. Gupta, T.D. Senguttuvan, Faster response of NO2 sensing in graphene–WO3 nanocomposites, Nanotechnology. 23 (2012) 205501. doi:10.1088/0957-4484/23/20/205501.
[50]
A. Morais, C. Longo, J.R. Araujo, M. Barroso, J.R. Durrant, A.F. Nogueira, Nanocrystalline anatase TiO2 /reduced graphene oxide composite films as photoanodes for
of
photoelectrochemical water splitting studies: the role of reduced graphene oxide, Phys.
[51]
ro
Chem. Chem. Phys. 18 (2016) 2608–2616. doi:10.1039/C5CP06707C.
J. Zhang, H. Lu, C. Yan, Z. Yang, G. Zhu, J. Gao, F. Yin, C. Wang, Fabrication of
-p
conductive graphene oxide-WO3 composite nanofibers by electrospinning and their
138. doi:10.1016/j.snb.2018.02.026.
X. Hu, P. Xu, H. Gong, G. Yin, Synthesis and characterization of WO3/graphene
lP
[52]
re
enhanced acetone gas sensing properties, Sensors Actuators, B Chem. 264 (2018) 128–
nanocomposites for enhanced photocatalytic activities by one-step in-situ hydrothermal
[53]
ur na
reaction, Materials (Basel). 11 (2018). doi:10.3390/ma11010147. L. Fu, T. Xia, Y. Zheng, J. Yang, A. Wang, Z. Wang, Preparation of WO3-reduced graphene oxide nanocomposites with enhanced photocatalytic property, Ceram. Int. 41
Jo
(2015) 5903–5908. doi:10.1016/J.CERAMINT.2015.01.022. [54]
D. Nagy, T. Firkala, E. Drotár, Á. Szegedi, K. László, I.M. Szilágyi, Photocatalytic WO3 /TiO 2 nanowires: WO 3 polymorphs influencing the atomic layer deposition of TiO2, RSC Adv. 6 (2016) 95369–95377. doi:10.1039/C6RA18899K.
[55]
B. Ahmed, A.K. Ojha, F. Hirsch, I. Fischer, D. Patrice, A. Materny, Tailoring of enhanced interfacial polarization in WO3 nanorods grown over reduced graphene oxide synthesized 27
by a one-step hydrothermal method, RSC Adv. 7 (2017) 13985–13996. doi:10.1039/c7ra00730b. [56]
D. Chen, L. Gao, A. Yasumori, K. Kuroda, Y. Sugahara, Size- And shape-controlled conversion of tungstate-based inorganic-organic hybrid belts to WO3 nanoplates with high specific surface areas, Small. 4 (2008) 1813–1822. doi:10.1002/smll.200800205. Q. Li, B. Guo, J. Yu, J. Ran, B. Zhang, H. Yan, J.R. Gong, Highly efficient visible-light-
of
[57]
ro
driven photocatalytic hydrogen production of CdS-cluster-decorated graphene nanosheets, J. Am. Chem. Soc. 133 (2011) 10878–10884. doi:10.1021/ja2025454.
H. Hajishafiee, P. Sangpour, N.S. Tabrizi, Facile Synthesis and Photocatalytic
-p
[58]
re
Performance of WO3/rGO Nanocomposite for Degradation of 1-Naphthol , Nano. 10 (2015) 1550072. doi:10.1142/s1793292015500721.
B. Chai, J. Li, Q. Xu, K. Dai, Facile synthesis of reduced graphene oxide/WO3 nanoplates
lP
[59]
composites with enhanced photocatalytic activity, Mater. Lett. 120 (2014) 177–181.
[60]
ur na
doi:10.1016/j.matlet.2014.01.094.
Y. Wang, Z. Chen, Y. Li, Z. Zhou, X. Wu, Electrical and gas-sensing properties of WO3 semiconductor material, Solid. State. Electron. 45 (2001) 639–644. S. Park, S. Park, J. Jung, T. Hong, S. Lee, H.W. Kim, C. Lee, H2S gas sensing properties
Jo
[61]
of CuO-functionalized WO3 nanowires, Ceram. Int. 40 (2014) 11051–11056. doi:10.1016/j.ceramint.2014.03.120.
[62]
S.M. Kanan, O.M. El-kadri, I.A. Abu-yousef, M.C. Kanan, Semiconducting Metal Oxide Based Sensors for Selective Gas Pollutant Detection, (2009) 8158–8196.
28
doi:10.3390/s91008158. [63]
N.D. Dien, L.H. Phuoc, V.X. Hien, D.D. Vuong, N.D. Chien, Hydrothermal Synthesis and Ammonia Sensing Properties of WO3/Fe2O3 Nanorod Composites, J. Electron. Mater. 46 (2017) 3309–3316. doi:10.1007/s11664-017-5283-5.
[64]
E. Llobet, G. Molas, P. Molinàs, J. Calderer, X. Vilanova, J. Brezmes, J.E. Sueiras, X.
[65]
ro
Electrochem. Soc. 147 (2000) 776. doi:10.1149/1.1393270.
of
Correig, Fabrication of Highly Selective Tungsten Oxide Ammonia Sensors, J.
D. Punetha, S.K. Pandey, Sensitivity Enhancement of Ammonia gas sensor based on
-p
hydrothermally synthesized rGO/WO3 nanocomposites, IEEE Sens. J. (2019) 1–1.
[66]
re
doi:10.1109/JSEN.2019.2950781.
T.M. Salama, M. Morsy, R.M. Abou Shahba, S.H. Mohamed, M.M. Mohamed, Synthesis
lP
of Graphene Oxide Interspersed in Hexagonal WO3 Nanorods for High-Efficiency Visible-Light Driven Photocatalysis and NH3 Gas Sensing, Front. Chem. 7 (2019).
[67]
ur na
doi:10.3389/fchem.2019.00722.
C.Y. Wang, X. Zhang, Q. Rong, N.N. Hou, H.Q. Yu, Ammonia sensing by closely packed WO3 microspheres with oxygen vacancies, Chemosphere. 204 (2018) 202–209.
Jo
doi:10.1016/j.chemosphere.2018.04.050. [68]
N. Van Toan, C.M. Hung, N. Van Duy, N.D. Hoa, D.T.T. Le, N. Van Hieu, Bilayer SnO2– WO3 nanofilms for enhanced NH3 gas sensing performance, Mater. Sci. Eng. B SolidState Mater. Adv. Technol. 224 (2017) 163–170. doi:10.1016/j.mseb.2017.08.004.
[69]
T. Tesfamichael, A. Ponzoni, M. Ahsan, G. Faglia, Gas sensing characteristics of Fe-
29
doped tungsten oxide thin films, Sensors Actuators, B Chem. 168 (2012) 345–353. doi:10.1016/j.snb.2012.04.032. [70]
M. D’Arienzo, L. Armelao, C.M. Mari, S. Polizzi, R. Ruffo, R. Scotti, F. Morazzoni, Surface interaction of WO3 nanocrystals with NH3. Role of the exposed crystal surfaces and porous structure in enhancing the electrical response, RSC Adv. 4 (2014) 11012–
F. Perrozzi, S.M. Emamjomeh, V. Paolucci, G. Taglieri, L. Ottaviano, C. Cantalini,
ro
[71]
of
11022. doi:10.1039/c3ra46726k.
Thermal stability of WS2 flakes and gas sensing properties of WS2/WO3 composite to H2,
-p
NH3 and NO2, Sensors Actuators, B Chem. 243 (2017) 812–822.
[72]
re
doi:10.1016/j.snb.2016.12.069.
Z. Qiu, X. Tian, Y. Li, Y. Zeng, C. Fan, M. Wang, Z. Hua, NH3 sensing properties and
lP
mechanism of Ru-loaded WO3 nanosheets, J. Mater. Sci. Mater. Electron. 29 (2018) 11336–11344. doi:10.1007/s10854-018-9221-y. D. Maity, R.T.R. Kumar, Polyaniline Anchored MWCNTs on Fabric for High
ur na
[73]
Performance Wearable Ammonia Sensor, ACS Sensors. 3 (2018) 1822–1830. doi:10.1021/acssensors.8b00589.
G. Halek, I.D. Baikie, H. Teterycz, P. Halek, P. Suchorska-Woźniak, K. Wiśniewski,
Jo
[74]
Work function analysis of gas sensitive WO3 layers with Pt doping, Sensors Actuators B Chem. 187 (2013) 379–385. doi:10.1016/j.snb.2012.12.062.
[75]
D.-T. Phan, G.-S. Chung, P–n junction characteristics of graphene oxide and reduced graphene oxide on n-type Si(111), J. Phys. Chem. Solids. 74 (2013) 1509–1514. doi:10.1016/j.jpcs.2013.02.007. 30
[76]
L. Sygellou, G. Paterakis, C. Galiotis, D. Tasis, Work Function Tuning of Reduced Graphene Oxide Thin Films, J. Phys. Chem. C. 120 (2016) 281–290. doi:10.1021/acs.jpcc.5b09234.
[77]
Abid, P. Sehrawat, S.S. Islam, P. Mishra, S. Ahmad, Reduced graphene oxide (rGO) based wideband optical sensor and the role of Temperature, Defect States and Quantum
Jo
ur na
lP
re
-p
ro
of
Efficiency, Sci. Rep. 8 (2018) 3537. doi:10.1038/s41598-018-21686-2.
31
ur na
lP
re
-p
ro
of
Figure caption
Jo
Figure 1. (A) Deign of gas sensor based on rGO/WO3 nanowire composites, and (B) Materials and sensor fabrication processes
32
of ro -p re lP
ur na
Figure 2. (A) low and (B) high magnification SEM images of the synthesized rGO/WO3
Jo
nanowire nanocomposites
33
of ro -p re lP
ur na
Figure 3. (A) XRD pattern, and (B) Raman spectrum of the synthesized rGO/WO3 nanowire
Jo
nanocomposites
34
of ro -p re lP
ur na
Figure 4. (A) low and (B) high magnification HRTEM images of the synthesized rGO/WO3
Jo
nanowire nanocomposites
35
of ro -p re lP
ur na
Figure 5. (A) UV-Vis spectrum; and (B) FTIR spectra of the synthesized rGO/WO3 nanowire
Jo
nanocomposites
36
of ro -p re lP
ur na
Figure 6. NH3 gas sensing characteristics of the synthesized rGO/WO3 nanowire nanocomposites measured at different temperatures: (A) transient resistance vs. time upon exposure to different NH3 concentrations; (B) sensor response, (C) response time, (D) recovery
Jo
time as a function of NH3 concentrations
37
of ro -p re lP
ur na
Figure 7. Sensor response plot of the synthesized rGO/WO3 nanowire nanocomposites on
Jo
exposure to 25-500 ppm NH3 and its linear fitting
38
of ro -p re lP
ur na
Figure 8. (A) Selectivity and (B) Stability of the as-made rGO/WO3 gas sensor; (C) stability
Jo
after storage of six months in ambient environment and a week working at 300oC
39
of ro -p re lP
Jo
ur na
Figure 9. Proposed gas sensing mechanism of the rGO/WO3 nanocomposites
40