1. Introduction
Tungsten trioxide (WO
3) is an n-type wide bandgap semiconductor with various functionalities [
1,
2,
3,
4]. Among numerous applications, WO
3 with various morphologies has received extensive attention as a forward-looking gas-sensing material due to its high sensitivity and stability toward target gases [
1,
5]. For example, it has been used to detect methane vapor, NO
2 gas, and CO gas with distinct sensing responses [
1,
6]. However, WO
3 has various crystallographic structures [
2]. Most gas-sensing properties are reported for the monoclinic structured WO
3; by contrast, reports for the hexagonal structured WO
3 being used as gas sensor materials are limited in number.
Oxides in a one-dimensional architecture have the advantage of high sensitivity and fast response/recovery speed due to their high surface-to-volume ratio and great surface activity compared to bulk or thin-film form [
7,
8,
9]. Therefore, the application of one-dimensional WO
3 nanostructures is one of the main strategies for increasing their gas-sensing performances. Many methods such as hydrothermal and electrospinning techniques have been proposed to fabricate one-dimensional WO
3 nanostructures [
10,
11,
12]. The hydrothermal synthesis of one-dimensional WO
3 nanostructures yields a large amount of WO
3 nanostructures from a solution at process temperatures lower than 250 °C. This method has the advantages of large-scale amount fabrication, low process cost, and easy process parameter control; therefore, hydrothermal methods are promising for synthesizing one-dimensional WO
3 crystals for gas-sensing applications. Recently, an improved gas-sensing ability of nanostructured WO
3 was achieved through heterostructure engineering. The intrinsic gas-sensing abilities of nanostructured WO
3 toward various target gases can be substantially enhanced through the coupling with other semiconductor oxides. For example, microwave synthesized Fe
2O
3-decorated WO
3 nanostructures exhibit improved H
2S gas-sensing performance [
13]. Electrospinning method-derived NiO particles functionalized with WO
3 porous composites demonstrate enhanced acetone gas-sensing responses [
11]. WO
3 nanosheets loaded with SnO
2 nanoparticles exhibit enhanced methane-sensing performance. Moreover, it has been shown that the loading content of SnO
2 nanoparticles has an important influence on the sensing behavior of WO
3–SnO
2 nanocomposites [
6].
Among the various coupling oxides integrated into WO
3, SnO
2 is also an n-type wide bandgap semiconductor, widely used as a gas-sensing material. It has been used to detect methanol, ethanol, and ethylene glycol gases with desirable sensing performance [
14,
15,
16]. Although improvement in the gas-sensing performance of SnO
2 nanoparticle-decorated monoclinic WO
3 nanosheets and tetragonal SnO
2-monoclinic WO
3 composite films has been reported [
4,
6], gas-sensing properties of hexagonally structured WO
3 nanorods coupled with thin coverage layers of SnO
2 have not yet been proposed. This might hinder the potential applications of hexagonally structured WO
3-based composite nanorods in gas sensor devices. In this study, SnO
2 thin layers with various thicknesses were sputter coated onto hexagonally structured WO
3 nanorods. Sputtering has advantages for in situ growing crystalline oxides with tunable film thickness [
17]. The microstructure-dependent gas-sensing behaviors of the hydrothermally derived WO
3 nanorods sputter coated with thin layers of SnO
2 were systematically investigated in this study.
3. Results and Discussion
Figure 1a shows the XRD pattern of hydrothermally derived WO
3 nanorods. The distinct Bragg reflections are ascribed to the (001), (002), and (301) of hexagonal WO
3 phase according to JCPDS No. 00-033-1387. Noteworthy, the intense (001) Bragg reflection in
Figure 1a demonstrated that highly c-axis-oriented WO
3 crystals were formed.
Figure 1b–d shows the XRD patterns of the WO
3 nanorods sputter coated with various thicknesses of SnO
2 layers. The corresponding XRD patterns exhibited a visible Bragg reflection centered at approximately 34.1°, which can be ascribed to tetragonal SnO
2 (101) (JCPDS No. 00-002-1337). Notably, with an increase in the sputtering duration of SnO
2 thin films, the intensity of the SnO
2 (101) peak gradually increased, revealing an increase in the thickness of the SnO
2 layers on the composite nanorods (
Figure 1b–d). No trace of other evident Bragg reflections from impurity phase were observed. Obviously, crystalline WO
3–SnO
2 composite nanorods with favorable properties were successfully synthesized by sputter-assisted coating of thin layers of SnO
2 on the WO
3 nanorods.
SEM micrographs and the corresponding high magnification images of as-synthesized WO
3 nanorods and various WO
3–SnO
2 composite nanorods are shown in
Figure 2. The micrographs of the WO
3 nanorods in
Figure 2a show that the pristine WO
3 nanorods feature obvious stripes on their surfaces extending along their growth directions and sharply structured heads. Moreover, the body is conically shaped.
Figure 2b–d shows SEM images of WO
3 nanorods sputter coated with SnO
2 shell layers with various thicknesses. These WO
3–SnO
2 composite nanorods had a different morphology compared with those of the pristine WO
3 nanorods. Notably, in
Figure 2b, the WS-1 composite nanorods had a polycrystalline structure on the surfaces of the WO
3 nanorods. Further increasing the SnO
2 sputtering duration, from
Figure 2c to
Figure 2d, the morphology of the sputter-coated SnO
2 shell layer gradually changed from tiny particle-feature-coverage layer to the coverage layer consisting of large crystal agglomerates. Notably, the shape of the top region of the composite nanorods in
Figure 2d also changed from the original conical pileup to a cylindrical shape and all nanorods took on uniform cylindrical shapes. The SEM images showed that the surface morphology of the WO
3–SnO
2 composite nanorods varied with the sputtering duration of the SnO
2. The roughening of the SnO
2 coverage layer of the composite nanorods with prolonged sputtering duration was clearly demonstrated herein. The similar phenomenon has also been demonstrated in the ZnO–ZnS composite nanorod system synthesized by sputter-assisted coating of the ZnS shell layer on ZnO nanorods with different sputtering durations [
18].
The detailed morphology, the SnO
2 coverage thickness, and the elemental distribution of various WO
3–SnO
2 composite nanorods (WS-1, WS-2, and WS-3) were further examined by TEM.
Figure 3a shows a low magnification WS-1 nanorod. The composite nanorod exhibited a conical pileup morphology. The coverage layer consisted of tiny particle crystals and the thickness of the coverage layer was estimated to be approximately 13 nm at the top region of the nanorod.
Figure 3b,c shows the high-resolution TEM (HRTEM) images of the WS-1 composite nanorod taken from the local interfacial regions of WO
3/SnO
2. The lattice fringes with a spacing of 0.39 nm in the inside region of the composite nanorods were assigned to the interplanar distance of hexagonal WO
3 (001). In addition, the lattice fringes with a spacing of approximately 0.26 nm in the outside region of the composite nanorods were attributed to the interplanar distance of tetragonal SnO
2 (101).
Figure 3d demonstrates the Sn, W, and O elemental mapping images of the WS-1 composite nanorod. The W element was located inside the composite nanorod, revealing the position of the WO
3 rod template. The Sn element spatially enclosed the whole rod body, demonstrating a homogeneous coverage of the SnO
2 on the WO
3 nanorods through sputtering SnO
2 deposition. Similarly in the cross-sectional EDS line-scanning profiles (
Figure 3d), the Sn and O signals were mainly distributed through the whole composite nanorod and the marked W signal was confined to the inner region of the composite nanorod, indicating that the composite nanorod consisted of a WO
3 core and a SnO
2 shell coverage layer.
Figure 4a shows a low magnification TEM image of the WS-2 nanorod. Similar to the WS-1 nanorod shown in
Figure 3a, the SnO
2 coverage layer of the WS-2 still consisted of tiny SnO
2 crystals. The SnO
2 coverage layer, however, was denser and the crystal size of the SnO
2 was larger than that in WS-1. The morphology of the WS-2 composite nanorod was more cylindrically shaped. The SnO
2 coverage thickness at the top region of the composite nanorod was evaluated to be approximately 25 nm.
Figure 4b,c shows the HRTEM images of the composite nanorods taken from the interfacial regions. The analysis the lattice fringes confirmed the crystal structures of the SnO
2 coverage layer and WO
3 nanorod. Furthermore, in
Figure 4d, the Sn, W, and O elemental mapping images of the WS-2 composite nanorod revealed the Sn and O elements to be homogeneously distributed over the whole composite nanorod. The W element was confined in the inner regions of the composite nanorod. The EDS line-scanning profiles across the composite nanorod in
Figure 4e supported the elemental mapping analysis results that the composite nanorods demonstrated a good WO
3–SnO
2 core–shell structure.
Figure 5a shows a low magnification image of the WS-3 composite nanorod. The morphology of the WS-3 nanorod took on a fully cylindrical shape after the SnO
3 sputtering deposition for 30 min. The crystal size of the SnO
2 coverage layer of the WS-3 nanorod was substantially increased through the prolonged sputtering duration of SnO
2 in comparison with those of the WS-1 and WS-2 nanorods. Moreover, these large SnO
2 crystals or aggregates resulted in a rugged surface morphology of the WS-3 nanorods. The SnO
2 coverage layer thickness at the top region of the composite nanorod was approximately 34 nm. The arrangements of local lattice fringes of the SnO
2 coverage layer and of the WO
3 core were also characterized in the HRTEM images (
Figure 5b,c).
Figure 5d,e demonstrates the homogeneous surface coverage of the Sn element through the whole WO
3 nanorod. Notably, the intensity of the Sn signal in the EDS line-scanning profiles of various WO
3–SnO
2 composite nanorods increased with SnO
2 sputtering duration as exhibited in the corresponding EDS spectra profiles, revealing an increased thickness of the SnO
2 coverage layers on the composite nanorods. The TEM results herein revealed a good coverage of the sputter-deposited SnO
2 thin films on the surface of the WO
3 nanorods. A schematic summary of the morphology changes of the WO
3–SnO
2 composite nanorods prepared with various sputtering durations of SnO
2 based on the TEM analysis results are also shown in
Figure 5f.
The peak intensity of W4f core-level doublets originating from the WO
3 nanorods decreased and the peak intensity of the Sn 3d core-level doublets of the sputtered SnO
2 coverage layers increased with the increase of the sputtering duration of SnO
2 in
Figure 6a,b. This reveals that the variation of the SnO
2 shell layer thickness on the composite nanorods is controlled by the change in the sputtering duration of the SnO
2 thin films.
Figure 6a displays the W4f
7/2 and W4f
5/2 peaks centered at 35.9 and 37.9 eV, respectively, for the WO
3 nanorods coated with various thicknesses of SnO
2 thin films. The Gaussian deconvolution results of the W4f spectra of the various composite nanorods illustrated the contributions corresponding to the W
5+ and W
6+ states in the WO
3. The main peaks centered at 35.9 and 37.9 eV correspond to W
6+ binding states; moreover, the weaker intensity and lower binding energies for the subpeaks centered at 35.2 and 36.5 eV correspond to W
5+ binding states.
Figure 6b displays XPS Sn3d spectra of various WO
3–SnO
2 composite nanorods. A correspondence Sn3d
5/2 peak centered at 487.1 eV and a Sn3d
3/2 peak centered at approximately 495.4 eV were observed. The binding energy difference between the Sn3d
5/2 and Sn3d
3/2 corresponded to the chemical binding component of Sn
4+ in the SnO
2 [
15,
17]. The O1s XPS spectra of various composite nanorods in
Figure 6c demonstrated three subcomponents. The component with a binding energy of approximately 529.8 eV is assigned to the tungsten oxide that formed the W–O bonds [
19]. The second component with a binding energy of 530.5 eV is assigned to the tin oxide that formed the strong Sn–O bonds [
15,
17], and the small degree of oxygen vacancies and/or oxygen species chemisorbed from the ambient air were demonstrated in the component centered at approximately 531.9 eV.
Figure 7a–c demonstrates the variation of the temperature-dependent gas-sensing responses of the pristine WO
3 nanorods, a 50 nm thick SnO
2 film, and various WO
3–SnO
2 composite nanorods upon exposure to 100 ppm acetone vapor. The optimum operating temperature of the composite nanorods is lower than that of the WO
3 nanorods and SnO
2 thin film, in which they exhibited an optimal operating temperature of 325 °C. Previously, Zhang et al. showed that the La
2O
3-decorated SnO
2 gas sensors feature a 2-fold higher gas-sensing performance improvement at 250 °C, which is lower than the optimum operating temperature of pristine SnO
2 of 300 °C [
20]. A decreased operating temperature of the oxide semiconductor through a heterostructure structure has also been shown in the ZnO–SnO
2 system [
21]. The optimal operating temperature for acetone gas-sensing examinations of various WO
3–SnO
2 composite nanorods was chosen to be 300 °C for this study.
Figure 8a–d shows the dynamic acetone gas-sensing response of pristine WO
3 nanorods and various WO
3–SnO
3 nanorods upon exposure to various acetone vapor concentrations. The acetone gas-sensing responses of gas sensors made from various nanorods increased with the acetone vapor concentration, revealing that an increased number of acetone molecule numbers interacted with the absorbed oxygen species on the surfaces of the nanorods [
15]. A plot of the acetone gas-sensing response vs. acetone vapor concentration is shown in
Figure 8e. The acetone gas-sensing response of the pristine WO
3 nanorods on exposure to 100–1000 ppm acetone vapor ranged from 1.25 to 1.35. The smooth and distinct dynamic response curves on exposure to various acetone vapor concentrations are demonstrated for the WO
3 nanorods in
Figure 8a, revealing that the WO
3 nanorods are responding to acetone vapor; however, the response values are not high enough for practical use. By contrast, in
Figure 8e, the WO
3 nanorods coated with various thicknesses of SnO
2 coverage layers exhibit an improved acetone gas-sensing capability compared with the pristine WO
3 nanorods. The highest level of enhancement in acetone gas-sensing response was observed for the WS-1 nanorods. Their responses ranged from 6.3 to 12.1 upon exposure to 100 to 1000 ppm acetone vapor, respectively. An approximately 5-fold increase in the acetone gas-sensing response upon exposure to 100 ppm acetone vapor was observed for WO
3 nanorods coated with 13 nm thick SnO
2 films. Notably, thicker SnO
2 coatings on the WO
3 nanorods did not further enhance the acetone gas-sensing response of the composite nanorods. The acetone gas-sensing response decreases with SnO
2 layer thickness as shown in
Figure 8e. The optimal coating thickness of the SnO
2 thin layer is approximately 13 nm for the WO
3–SnO
2 composite nanorod system herein. The response and recovery times for the gas sensors made from various nanorod samples are defined as the duration required to drop the 90% resistance on exposure to the target gas and that to increase 90% resistance with the removal of the target gas. The response times for the pristine WO
3 nanorods exposed to 100–1000 ppm acetone vapor concentrations ranged from 8 to 17 s, whereas recovery times ranged from 33 to 58 s in the same acetone vapor concentration range. By contrast, the response times for the sensors made from the WS-2 and WS-3 composite nanorods ranged from 6 to 16 s and 20 to 32 s, respectively. The recovery times of the WS-2 and WS-3 composite nanorods were 38–83 s and 35–100 s in the acetone concentration range of 100–1000 ppm, respectively. Substantially increased response and recovery times were observed for the composite nanorods with the thickest SnO
2 coverage layer. Notably, the response times and recovery times for the WS-1 composite nanorods in the same test acetone vapor concentration range ranged between 3–13 s and 28–51 s, respectively. The slightly improved response and recovery speeds together with the substantial enhancement in gas-sensing responses revealed an improved acetone gas-sensing performance of the WO
3–SnO
2 composite nanorods with an optimal SnO
2 thin layer thickness of approximately 13 nm.
Figure 8f displays cyclic acetone gas-sensing tests for the WS-1 composite nanorods exposed to 500 ppm acetone vapor. Steady gas-sensing activity under five test cycles was observed, confirming that WS-1 nanorods were reproducible and stable for detecting acetone vapor.
Figure 8g shows the gas-sensing selectivity of CH
3COCH
3, H
2, and NH
3 gases with concentrations of 100 and 10 ppm for NO
2 gas for the WS-1 composite nanorods. The WS-1 composite nanorods exhibited the best gas-sensing response toward acetone vapor among the various target gases.
Table 1 summarizes the acetone gas-sensing responses of the WO
3-based composites operating at the temperature range of 280–400 °C [
22,
23,
24,
25,
26,
27]. In comparison, the WS-1 composite nanorods presented the best acetone vapor detection performance among various reported works.
The acetone gas-sensing mechanism of WO
3–SnO
2 composite nanorods with an optimum thickness of SnO
2 layer is illustrated in
Figure 9. In our assessment of the gas-sensing mechanism, we follow the approach outlined in reference [
28]. When thin SnO
2 coverage layers are sputtered onto the surfaces of the WO
3 nanorods, an interfacial depletion layer forms as a WO
3/SnO
2 heterojunction develops. According to the band alignment structure of the WO
3/SnO
2 in
Figure 9a [
29,
30], an electron depletion layer initially forms on the WO
3 side. Upon exposure of the WO
3–SnO
2 composite nanorods to ambient air, oxygen molecules remove surface electrons from the composite nanorods, thus forming adsorbed negatively charged oxygen species (
Figure 9b). In this way, the sputter-coated SnO
2 thin layers in the composite nanorods become partially depleted at some degrees under the sensor operating conditions. Notably, the extent of this depletion is highly dependent on the sensor operating temperature. The thickness of the surface depletion layer of the SnO
2 at 327 °C in ambient air has been shown to be approximately 21 nm [
31]. In ZnO–SnO
2 core–shell nanowires, it has been proposed that the SnO
2 shell layers with a thickness of 15–20 nm was fully depleted in the range of sensor operating temperatures of 200–400 °C [
32]. The complete depletion of the SnO
2 shell layer significantly affects the variation of the electron depletion layer width at the ZnO/SnO
2 heterointerface during gas exposure, thus leading to a substantially enhanced gas-sensing performance [
30]. In the current work, the extraction of surface electrons widened the thickness of the interfacial depletion layer and thus increased the interfacial potential barrier. In this way, the width of the electron conduction path through the single crystalline WO
3 nanorods is reduced and the resistance of the composite nanorods is increased (
Figure 9b). When acetone vapor is introduced into the test chamber, the interaction between the reducing acetone vapor and the adsorbed oxygen species of the surfaces of the composite nanorods can be expressed as follows [
33]:
The release of free electrons from the adsorbed surface oxygen species eliminates the surface depletion regions in the SnO
2 coverage layers and further narrows the size of interfacial depletion regions, thus causing the substantial drop in resistance of the composite nanorods. The marked resistance variation of the composite nanorods before and after introducing acetone vapor results in a distinct gas-sensing response of the composite nanorods upon exposure to acetone vapor. The efficiency of this process is critically dependent on the thickness of the SnO
2 coverage layers. Our work suggests that in WO
3–SnO
2 composite nanorods this optimum thickness is close to 13 nm. In its order of magnitude, this value agrees with the results of previous investigations into core–shell SnO
2–ZnO nanowires [
34] and SnO
2–ZnO nanofibers [
35]. It appears that in all cases a substantial improvement in gas response is obtained when the thickness of the coverage layers is close to the respective Debye length.