You are on page 1of 7

Journal of Solid State Chemistry 273 (2019) 25–31

Contents lists available at ScienceDirect

Journal of Solid State Chemistry


journal homepage: www.elsevier.com/locate/jssc

Study of WO3–In2O3 nanocomposites for highly sensitive CO and NO2


gas sensors
Yu.S. Haiduk a, **, A.A. Khort b, c, *, N.M. Lapchuk a, A.A. Savitsky a
a
Belarusian State University, 220030 Minsk, Belarus
b
A.V. Luikov Heat and Mass Transfer Institute of the National Academy of Sciences of Belarus, 220072 Minsk, Belarus
c
Center of Functional Nano-Ceramics, National University of Science and Technology “MISIS”, 119049 Moscow, Russia

A R T I C L E I N F O A B S T R A C T

Keywords: Pure semiconductor tungsten oxide (WO3), indium oxide (In2O3) and mixed nanocomposites with different WO3
WO3 to In2O3 ratios were successfully synthesized by simple sol-gel method following calcination at 600  C. The
In2O3 morphology, phase composition and features of crustal structure of the materials were studied by X-ray diffrac-
Sol-gel
tion, Fourier-transform infrared spectroscopy, scanning electron microscopy, transition electron microscopy and
Semiconductor gas-sensors
CO
electron paramagnetic resonance spectroscopy. It was found that the nanocomposite materials are characterized
NO2 by fine crystallinity of 8–38 nm, highly defective crystal cells and presence of delocalized electrons in their
structures, which can significantly affect gas sensitivity. The gas sensors based on WO3–In2O3 composite struc-
tures exhibited excellent CO and NO2 detecting performance at optimal operating temperature of ~240  C
and ~140  C, respectively.

1. Introduction semiconductors can be used to manufacture sensors for detection of both


oxidizing gases (such as NO2) [14] and reducing gases (such as CO) [15].
Semiconductor sensors are widely applied in various fields of gas One of the most widely studied way to increase sensitivity and
analysis: monitoring of air quality, control of technological processes, selectivity of semiconductor oxide sensors is to create sensing materials
toxic gas detecting, etc. [1]. The basic operating principle of semi- with a complex highly defected crystal structure, which is characterized
conductor gas sensors is chemical adsorption and desorption of detected by high specific surface area and low activation energy of the possible
gas molecules on the oxide surface, which leads to a change in the con- conversion reaction. This can be achieved by doping of sensing materials
centration of electrons in the conduction band and, accordingly, elec- with metals (Ag, Au, Pd, and Pt) or other oxide semiconductors [8,14,16,
trical conductivity of the sensing material [2–5]. It is highly desirable 17]. Earlier, numbers of studies were reported on the synthesis and
that semiconductor sensors have a large surface area, so as to adsorb on sensing applications of composite nanostructures such as SnO2/ZnO [18,
the surface as much of the target gas molecules as possible, giving a 19], SnO2/In2O3 [20,21], SnO2/NiO [22], Fe2O3/In2O3 [23], ZnO/In2O3
stronger and more measurable response [6]. Thanks to this feature, in [24], Fe2O3/WO3 [25], WO3/SnO2 [25,26] and others. At the same time,
some cases, sensitivity of a semiconductor surface to a gas can be as low despite the fact that pure In2O3 and WO3 individually are among the best
as parts per billion (ppb). gas sensing semiconductors, information on the study of the gas sensitive
In the past few years, great efforts have been devoted to improve gas properties and structural features of WO3–In2O3 complex materials of
sensor performances such as sensitivity, selectivity and reversibility [7]. various mass ratios is very limited [27,28].
The most common used semiconductor oxide materials include, but are In this study, we prepared pure tungsten oxide and indium oxide
not limited to tin oxide (SnO2), indium oxide (In2O3), zinc oxide (ZnO), nanopowders and their compositions by sol-gel method following calci-
molybdenum oxide (MoO3), titanium oxide(TiO2), tungsten oxide (WO3) nation at 600  C for 2 h. Various techniques were employed for studying
and others [7–9]. These materials are being widely researched owing to of synthesis process and characterization of the structure and
their low cost, simple synthesis process and possibility to cast them in morphology of the prepared nanopowders. The gas sensing properties of
devices of compact size [10–13]. Moreover, nanostructured metal oxide the pure and composite oxide nanomaterials were studied and the results

* Corresponding author. A.V. Luikov Heat and Mass Transfer Institute of the National Academy of Sciences of Belarus, 220072 Minsk, Belarus.
** Corresponding author.
E-mail addresses: j_hajduk@bk.ru (Yu.S. Haiduk), khort@hmti.ac.by (A.A. Khort).

https://doi.org/10.1016/j.jssc.2019.02.023
Received 14 October 2018; Received in revised form 6 February 2019; Accepted 15 February 2019
Available online 18 February 2019
0022-4596/© 2019 Published by Elsevier Inc.
Yu.S. Haiduk et al. Journal of Solid State Chemistry 273 (2019) 25–31

exhibited that the formation of highly defected and distorted crustal scanning electron microscope Leo-1420 Carl Zeiss. Transmission Electron
structure along with a high specific surface area of the sensors provided a Microscopy (TEM) images were taken on a Hitachi H-800 series oper-
significant enhanced sensing performance to carbon monoxide and ni- ating at 200 keV.
trogen dioxide, which can be useful for creation of efficient low-
consumption semiconductor gas sensors. 2.3. Gas tests

2. Experimental Sensor elements were fabricated in a form of gas-permeable tablets


(d ¼ 10 mm, h ¼ 3.5 mm) by pressing of the obtained experimental
2.1. Materials synthesis powders with an organic binder on a hydraulic press at 150 kPa, which
then were sintered in air at 450  C for 5 h. To enhance electrical con-
All chemicals were the analytical-grade reagents and used without ductivity Ag electrodes were deposited on parallel sides of the tablets.
any further purification. Gas sensing properties of the sensor elements were characterized using a
Nanopowders of pure indium and tungsten oxides (In2O3 and WO3) home-designed flow type sensing measurements system inside an
and their mixtures at In2O3 to WO3 ratios of 25:75, 50:50 and 75:25 aluminum chamber with precisely controlled temperature and atmo-
(samples 25InW, 50InW and 75InW, respectively) were prepared by sol- sphere. Sensitivity (S) was calculated according to equations: S ¼ (Ra/Rg)
gel method. In a typical synthesis route 1.23 M aqua solution of sodium ∙100%, in case of reducing gas mixtures, and S ¼ (Rg/Ra)∙100%, for
tungstate dihydrate (Na2WO3⋅2H2O) was added into 12 M aqua solution NO2/air mixture. Here Ra is a resistance of the sensor element in the air
of nitric acid under constant strong stirring. After stirring the prepared and Rg is resistance in the test gas mixture. The measurement procedure
sol of tungstic acid was washed in distilled water using multiply centri- was carried out as follows. Sensing element was placed into preheated
fugation. Indium hydroxide sol was prepared by adding of 9.24 M and thermostabilized chamber. Then testing gas mixtures (CO/air or CO/
ammonia aqua solution to 0.78 M aqua solution of In(NO3)∙4.5H2O N2 with CO concentration of 3000 ppm) were injected into the chamber
under constant stirring and washed by multiply centrifugation. Then the at a rate of 2 l/h. during 10 min. And, finally, the chamber with the
sols were mixed in the proportions in regard to obtain mixtures of oxides sensing element was refilled by air for another 10 min. After this, the
of required final compositions. The obtained mixtures and pure sols of measurement was repeated. The same procedure was carried out for
both compositions have been drying until xerogels have formed and, measurements of materials sensitivity to acetone (45000 ppm) and NH3
finally, calcined at 600  C for 2 h. (2000 ppm). Sensitivity to NO2 was measured using NO2/air mixture
with NO2 concentration of 1 ppm in a stationary regime in a precisely
2.2. Characterization controlled atmosphere.

Thermal evaluation of the xerogels was conducted by simultaneous 3. Results and discussion
thermal gravimetric analysis (DTA-TG) NETZSCH STA 449 F3 Jupiter
from room-temperature to 600  C in nitrogen with a heating rate of 10 C/ In order to investigate features of In2O3 and WO3 nanooxides syn-
min. X-ray powder diffraction (XRD) analysis was conducted on a DRON- thesis process thermal behavior of the initial In(OH)3 and H2WO4⋅H2O
3 X-ray diffractometer with Cu Kα1 radiation (α ¼ 1.5405A). The refer- xerogels were studied by TG–DTA analysis from room temperature to
ence data was used from the PDF2 database. The average crystallite size 600  C. The TG-DTA curves of the both samples are shown in Fig. 1.
(D) was determined from the broadening of the diffraction peak using the There are several stages can be clearly distinguished at TG-DTA
Scherrer formula D ¼ Kλ/βcosθ where D is the average particle size in curves during In(OH)3 xerogel thermal decomposition (Fig. 1, left
(nm), K is the Scherrer constant (0.89), λ is the wavelength of X-ray image). At first, dehydration of xerogel from room temperature to
source, β is the full width at half-maximum and θ is the Bragg's angle. The ~175  C resulted in mass loss of 5.35%. Then, the main mass reduction of
experimental XRD data were processed with HighScore Plus software. 15.77% and 2.49% was detected from ~175  C to ~250  C and from
Fourier-transform infrared (FTIR) spectra were recorded on Avatar 330 ~250  C to 340  C, respectively, which can be attributed to thermal
FTIR spectrometer equipped with Smart Diffuse Reflectance system at decomposition of indium hydroxide to indium trioxide. The next stage of
room temperature using KBr pellet method. Electron paramagnetic decomposition, which gave only 0.74% of mass loss, was connected with
resonance (EPR) spectra of the experimental samples were recorded at removal of absorbed gases, like CO2, N2, etc. [30].
room temperature on a RadioPan SE/X-2543 with a H102 resonator in the In case of H2WO4⋅H2O (Fig. 1, right image), the observed mass loss of
X-band spectrometer operating at 100 kHz with amplitude of 1 G and 6.51% from room temperature to ~100  C can be attributed to dehy-
microwave radiation frequency in the resonator of about 9.3 GHz. dration of xerogel, followed by thermal decomposition of H2WO4 with
Quality of the resonator factor was controlled by an oriented ruby single total mass reduction of about 8.5%. After finishing of formation of
crystal. Electron microscopy study was performed by means of the tungsten oxide (~270  C), only removal of absorbed gases and WO3

Fig. 1. TG–DTA curves of In(OH)3 (left) and H2WO4 (right) xerogels.

26
Yu.S. Haiduk et al. Journal of Solid State Chemistry 273 (2019) 25–31

changes in a similar way: the largest is for WO3 (~24 m2/g), lower is for
mixed compositions (~11–13 m2/g) and the lowest is for pure indium
oxide (~10 m2/g).
The appearance of four sharp bands peaks around 410, 540, 565 and
600 cm1 at FTIR curves of all In2O3–containing samples (Fig. 3) can be
assigned to the phonon vibrations of In–O bonds and are characteristic of
cubic In2O3 [32]. The band located at 710–880 cm1 was attributed to
W–O–W stretching vibration of WO3 in tungsten oxide–containing sam-
ples [33]. In addition, the broad structureless absorption at
3600–3100 cm1, as well as peaks around 1625–1630, 1425–1500 and
1330–1350 cm1 correspond to the absorbed hydroxyl groups on the
surfaces of the oxide nanomaterials.
FTIR spectra of the samples are in a good agreement with the results
of XRD and TG-DTA analysis, showing full thermal decomposition of
xerogels and formation only two oxide phases, without any by-phases.
Fig. 4 shows typical microstructure of the powder samples at low
magnification. Grains of different sizes from hundreds nanometers to
several micrometers can be observed in the microstructures. Shape of the
most WO3 grains is close to rhombohedral with some degree of distor-
Fig. 2. XRD patterns of the experimental samples. tion. Smaller grains are uniformly distributed on the surfaces of larger
grains. On the other hand, geometric shape of grains of 25InW sample is
not defined clearly, which is confirmed by TEM image (Fig. 4 bottom
recrystallization took place at higher temperatures [31].
right).
To study features of phase compositions and crystal structures of
There are many distorted polygons with fuzzy edges and angles,
prepared oxides and their mixtures XRD and FTIR analysis were carried
which can be due to distortion of crystal structures in mixed oxide ma-
out.
terial. The sizes of grains are more uniform in comparison with WO3
XRD data analysis (Fig. 2) showed that non-mixed samples are pure
powder. Further increase in In2O3 content leads to appearance of large
indium and tungsten oxides, respectively.
(up to 15 μm) grains and their aggregates, covered by smaller grains. The
In their turn, samples 25InW, 50InW and 75InW are mixtures of both
element mapping analysis (Fig. 5) shows that larger grains are indium
oxides with peak intensities that correspond to the relative content of
reach phase, while smaller one are tungsten oxide grains uniformly
each of the oxide phases. Peaks of any minor phases were not detected.
distributed on their surface. The observed addiction is in a good agree-
In2O3 phase in all samples is characterized by cubic (Ia3) crystal structure
ment with the specific surface area measurements, predicting grains size
(JCPDS card no. 71–2194), the calculated cell parameters of which
increase along with increase of indium oxide content. In general, the
decrease with increase of the content of tungsten oxide in the composi-
most uniform distribution of the elements is observed in 25InW sample,
tion of the materials (Table 1). In case of tungsten oxide the same trend is
where grains of both oxides are the smallest, while 50InW and 75InW
observed: calculated values of cell parameters of orthorhombic (Pnma)
samples contain relatively large grains of both oxides uncovered by the
WO3 (JCPDS card no. 43–0679) increase with increasing content of the
other one.
phase (Table 1).
The study of sensitivity of 25InW sample on CO/air, CO/N2 and NO2
The cell shrinkage of both oxide phases in the mixed samples can be
at different temperatures are shown in Fig. 6a. The sensitivity on the NO2
explained by their reciprocal influence. We suppose that during heat
containing gas mixture increases dramatically from 32% to its maximum
treatment indium and tungsten incorporate into crustal structures of
value in a temperature range from 60  C to 130–140  C. Further increase
oxides of each other, which leads to structure distortion. Also, due to the
in temperature leads to its gradual decrease up to 15% at 240  C. In case
fact that both indium and tungsten are metals with variable valence, the
of CO/air and CO/N2 gas mixtures the sensitivity significantly increases
distortion of their crystal cells may be the result of formation of small
from 68% at 168  C to 96% at 222  C. At higher temperatures only a
amount of mixed oxides during heat treatment. The both factors
slight sensitivity increase with the maximum sensitivity value at 242  C
described above promote formation of oxygen vacancies that also effects
was detected. It should be noted, sensitivity vs temperature curves are
on cell parameters of the oxides and can significantly changes electrical
quite the same for all In2O3–WO3 samples with almost the same peak
characteristics of the materials.
sensitivity at 140  C and 242  C for NO2 and CO, respectively. This fact
It also should be noticed, that In2O3 phase in all the samples has
makes it possible to compare maximal sensitivity of different materials at
practically the same calculated average crystalline sizes of 8–10 nm and
the same optimal operating temperature.
is not affected by the WO3 content. In the other hand, average crystalline
Typically, operating temperature significantly influences on the
sizes of WO3 phase sequentially change as the content of indium oxide
response of a semiconductor gas sensor. The low sensitivity at low tem-
decreases from 15 to 21 nm, for composite materials, to 38 nm, for pure
peratures is mainly attributed to CO and NO2 molecules having no
oxide sample (Table 1). The specific surface area of the compositions
enough thermal energy to react on the sensor surface The energy,

Table 1
Calculated crystallographic parameters and crystalline sizes of oxides of experimental samples.
Sample name Crystallographic parameters d, nm

In2O3 WO3 In2O3 WO3

a, Å V, Å Space group a, Å b, Å c, Å V, Å Space group

In2O3 10.585 1185.97 Ia3 (cubic) – – – – – 8 –


75InW 10.224 1068.72 7.402 7.664 7.849 445.27 Pnma (orthorhombic) 8 21
50InW 10.216 1066.21 7.416 7.683 7.859 447.78 8 17
25InW 10.199 1060.90 7.414 7.687 7.852 447.50 10 15
WO3 – – – 7.422 7.693 7.870 449.36 – 38

27
Yu.S. Haiduk et al. Journal of Solid State Chemistry 273 (2019) 25–31

oxygen molecules from air, which can capture free electron from the
conduction band of semiconductor oxide. The adsorbed oxygen forms the
chemisorbed oxygen species O, which turns to an electron depletion
layer on the surface and forms the potential barrier. Molecules of NO2 can
be directly adsorbed onto oxide surface by capturing free electron from
the conduction band or interact with the chemisorbed oxygen. Both
mechanisms will lead to increase of thickness of electron depletion layer
and substantial increase of the resistance of semiconductor oxide [34].
On the other hand CO molecules interact with chemisorbed oxygen
leading to electron transfer back into the nanowire, thereby increasing
materials conductivity.
It can be seen (Fig. 6b and c), that pure tungsten oxide is more sen-
sitive than In2O3 to the both CO-containing gas-mixtures. While the
sensitivity of pure oxides to NO2 is almost the same [35]. On the other
hand, mixed oxide samples are characterized by higher sensitivity than
the pure oxides. From the data analysis one can conclude, that decrease
in In2O3 content leads to a slight increase of the gas sensitivity sample
and 25InW demonstrates the highest sensitivity in all cases. To check the
trend an additional sample containing 5% of In2O3 (sample 5InW,
marked by asterisk (*) in Fig. 6b and c) were synthesized and tested.
According to the study the sensitivity of 5InW sample to the CO is
significantly lower (~8–32% depending on the gas-mixture), than for
25InW sample. At the same time, in case of NO2 detection, the sensitivity
of the 5InW is about 35% higher, than for 25InW and comparable with
previously published data on the absolute value [28,29]. As the specific
surface area of WO3 is higher than in mixed samples their higher sensi-
tivity can't be explained by simple difference in amount of adsorbed
molecules of gases. We suppose, joint heat treatment leads to formation
of complex defective structure in both oxides, which was confirmed by
XRD and FTIR data. The presence of defects of a different nature can lead
to a significant change in the nature of the conductivity associated with
an increase in the number of free or delocalized electrons, hence the
Fig. 3. FTIR spectra of experimental samples of pure tungsten and indium ox- effective carrier concentration near the surface of sensing materials.
ides and their mixtures. Another factor could potentially impact on the sensitivity is decrease of
specific surface area as indium oxide content increases.
Comparative study of sensitivity of 25InW sample to CO-based gas
required to overcome activation energy barrier on the materials surface
mixtures, acetone and ammonia at the same operating temperature
has a specific value for each molecule, mostly specified by its nature. This
shows that sensitivity of the sensors to CO/N2 is 3.18 and 2.14 times
leads to difference in optimal operating temperature for different gases
higher than for acetone and NH3 respectively (Fig. 6d). From the data one
even on the surface of the same material.
can conclude the experimental semiconductor gas sensor is characterized
Increase in operating temperature leads to increase in thermal energy
by superior selectivity at optimal operating temperature.
of absorbed gas molecules to the values sufficient to overcome the acti-
The study showed that the maximal sensitivity of the sensing ele-
vation energy barrier. After this point further operating temperature in-
ments in CO-containing gas mixtures is achieved in ~ 9–10 min in all
crease can possible leads to decrease of sensitivity due to the conversion
cases, while it takes only ~ 4–5 min to recover completely. In case of NO2
of adsorbed oxygen species which will capture more electrons from the
detection, maximal sensitivity is reached in 16 min. This indicates that
sensing material and decrease its gas adsorption ability. Also, the con-
the limiting factor of saturation time is adsorption/desorption of CO and
ductivity of oxide samples increases with temperature, due to increased
NO2 molecules on the surface of the sensory material and the sensitivity
amount of high mobile electrons [7]. Materials surface can adsorb

Fig. 4. SEM images of the experimental samples and TEM image of the 25InW sample (right bottom).

28
Yu.S. Haiduk et al. Journal of Solid State Chemistry 273 (2019) 25–31

Fig. 5. Element mapping of the mixed oxide samples 25InW, 50InW and 50InW.

Fig. 6. Sensitivity of In2O3–WO3 tablets (a) at different operation temperatures (for sample 25InW) and as function of In2O3 content at (b) CO- and (c) NO2-containing
gas mixtures. (d) Comparison of sensitivity value of 25InW sample to different gases at 242  C.

does not depend on sensitivity of the specific material. and width h ¼ 43.93 G), associated to the existence of a hyperfine
For further investigation of the defect structure, studying of EPR structure. The anisotropic shape of the second peak is characteristic for
spectra of the In2O3–WO3 nanopowders were carried out, which are hole centers, their associates and F-centers.
useful to detect the unpaired electrons from paramagnetic materials The peak anisotropy along with the existence of hyperfine structure
(Fig. 7). and noticeable deviation of the g-factor from the value of ge (i.e. g-factor
In the spectrum of WO3 (Fig. 7a) there are two peaks, near for the free electron) indicate partially reduced valence states of tungsten
1500 G (g ¼ 4.28252 and width h ¼ 79.93 G) and 3200 G (g ¼ 2.0753 ions. The weak resolution of the hyperfine structure can be associated

29
Yu.S. Haiduk et al. Journal of Solid State Chemistry 273 (2019) 25–31

Fig. 7. Room-temperature EPR spectra of the In2O3–WO3 samples.

with delocalization of an unpaired electron over several tungsten atoms with pure oxide samples.
at the lattice sites, which indicates defected crustal structure of WO3. Special attention to EPR spectrum of the sample 50InW is required
However, the nature of possible structural defects and their associates (Fig. 7c). The spectrum could be characterized by two peaks: the first at
can't be described reliably due to the low intensity and weak resolution of 1547 G (g ¼ 4.27802) is quite typical for pure oxide materials and the
signals of the hyperfine structure. On the other hand, two peaks on the second one, at 2815 G (g ¼ 2.35875) is highly distorted in comparison
EPR spectrum of In2O3 (Fig. 7e) at 1507 G (g ¼ 4.34211 and h ¼ 51.86 G) with typical peaks for indium and tungsten oxides. We suppose, the
and 3320 G (g ¼ 2.00256 and h ¼ 9.27 G) respectively show almost distorted peak could appear as a result of a resonance fusion of close in
complete absence of traces of hyperfine structure. Moreover, value of the energy peaks of the oxides. Another possible explanation is peak distor-
second g-factor close to ge indicates possible presence of radicals and tion as a result of paramagnetic transitions with a width and intense peak
atoms with an excess electrons in the structure of the indium oxide, at 2400 G. However, the effect should be more carefully considered for a
which can share their electrons with the indium atoms in the oxide more precise description of its nature and causes, that is not the aim of
structure [36]. In case of mixed oxide compositions of samples 25InW this work. One can definitely say that this effect does not influence
and 75InW EPR spectra show peaks at around 1500 G with g ¼ 4.2998 noticeably on the sensitivity of the materials to the gas mixtures
and 4.3013 respectively, and another highly symmetric peaks at around investigated.
3220 G with g value of 2.0011 close to ge (Fig. 7b and d). The data It should be noted the EPR data confirmed presence of the highly
indicate a complication of the defective oxide structures in comparison defected crustal structure saturated by delocalized electrons, which could

30
Yu.S. Haiduk et al. Journal of Solid State Chemistry 273 (2019) 25–31

be the reason of significant enhance of gas-sensing properties of the [13] D. Chen, L. Ge, L. Yin, H. Shi, D. Yang, J. Yang, Solvent-regulated solvothermal
synthesis and morphology-dependent gas-sensing performance of low-dimensional
25InW, 50InW and 75InW powders.
tungsten oxide nanocrystals, Sensor. Actuator. B Chem. 205 (2014) 391–400,
https://doi.org/10.1016/j.snb.2014.09.007.
[14] B. Urasinska-Wojcik, T.A. Vincent, M.F. Chowdhury, J.W. Gardner, Ultrasensitive
4. Conclusion
WO3 gas sensors for NO2 detection in air and low oxygen environment, Sensor.
Actuator. B Chem. 239 (2017) 1051–1059, https://doi.org/10.1016/
In summary, semiconductor WO3, In2O3 and WO3–In2O3 nano- j.snb.2016.08.080.
composites were successfully prepared assisted by a sol-gel method. The [15] N.P. Zaretskiy, L.I. Menshikov, A.A. Vasiliev, Theory of gas sensitivity of
nanostructured MOX layers: charge carrier self-exhaustion approach, Sensor.
careful studies showed the appearance of highly defected crustal struc- Actuator. B Chem. 175 (2012) 234–245, https://doi.org/10.1016/
tures of both oxide phases in nanocomposite while simultaneous absence j.snb.2012.08.068.
of any solid by-products. The gas sensitivity indicates that the gas sensors [16] J. Zhang, X. Liu, G. Neri, N. Pinna, Nanostructured materials for room-temperature
gas sensors, Adv. Mater. 28 (2016) 795–831, https://doi.org/10.1002/
based on WO3/In2O3 nanocomposites exhibit a high response to carbon adma.201503825.
monoxide and nitrogen dioxide at optimal operating temperatures of [17] D. Chen, L. Yin, L. Ge, B. Fan, R. Zhang, J. Sun, G. Shao, Low-temperature and
~240  C and ~140  C. Comparative gas sensitivity between nano- highly selective NO-sensing performance of WO3 nanoplates decorated with silver
nanoparticles, Sensor. Actuator. B Chem. 185 (2013) 445–455, https://doi.org/
composites of different In2O3 to WO3 ratios and pure oxides indicates 10.1016/j.snb.2013.05.006.
that the nanocomposites exhibit a better sensitivity to the investigated [18] K.W. Kim, P.S. Cho, S.J. Kim, J.H. Lee, C.Y. Kang, J.S. Kim, S.J. Yoon, The selective
gas mixtures then the individual oxides. At the same time, an increase in detection of C2H5OH using SnO2-ZnO thin film gas sensors prepared by
combinatorial solution deposition, Sensor. Actuator. B Chem. 123 (2007) 318–324,
the content of indium oxide in the nanocomposite over 5% in case of NO2
https://doi.org/10.1016/j.snb.2006.08.028.
detection and over 25% in case of CO detection led to a decrease in [19] I.S. Hwang, S.J. Kim, J.K. Choi, J. Choi, H. Ji, G.T. Kim, G. Cao, J.H. Lee, Synthesis
sensitivity. This could be attributed to degree of defectiveness of crystal and gas sensing characteristics of highly crystalline ZnO-SnO2 core-shell nanowires,
Sensor. Actuator. B Chem. 148 (2010) 595–600, https://doi.org/10.1016/
structure of the materials and the resulting amount of delocalized elec-
j.snb.2010.05.052.
trons on the surface of sensing elements. Although nanocomposites based [20] H. Du, J. Wang, M. Su, P. Yao, Y. Zheng, N. Yu, Formaldehyde gas sensor based on
on tungsten and indium oxides require further detailed study, they could SnO2/In2O3 hetero-nanofibers by a modified double jets electrospinning process,
be very promising materials for energy-efficient CO and NO2 gas sensors. Sensor. Actuator. B Chem. 166–167 (2012) 746–752, https://doi.org/10.1016/
j.snb.2012.03.055.
[21] A. Chen, S. Bai, B. Shi, Z. Liu, D. Li, C.C. Liu, Methane gas-sensing and catalytic
Acknowledgments oxidation activity of SnO2-In2O3 nanocomposites incorporating TiO2, Sensor.
Actuator. B Chem. 135 (2008) 7–12, https://doi.org/10.1016/j.snb.2008.06.050.
[22] L. Liu, Y. Zhang, G. Wang, S. Li, L. Wang, Y. Han, X. Jiang, A. Wei, High toluene
The authors gratefully acknowledge the financial support of the Re- sensing properties of NiO-SnO2 composite nanofiber sensors operating at 330  C,
public of Belarus research program “Energetic systems, processes and Sensor. Actuator. B Chem. 160 (2011) 448–454, https://doi.org/10.1016/
j.snb.2011.08.007.
technologies”, task 2.80 and the Ministry of Science and Higher Educa- [23] M. Ivanovskaya, D. Kotsikau, G. Faglia, P. Nelli, Influence of chemical composition
tion of the Russian Federation in the framework of Increase Competi- and structural factors of Fe2O3/In2O3 sensors on their selectivity and sensitivity to
tiveness Program of NUST «MISiS» (К2-2019-007), implemented by a ethanol, Sensor. Actuator. B Chem. 96 (2003) 498–503, https://doi.org/10.1016/
S0925-4005(03)00624-5.
governmental decree dated 16th of March 2013, N 211.
[24] R.K. Gupta, K. Ghosh, R. Patel, S.R. Mishra, P.K. Kahol, Band gap engineering of
ZnO thin films by In2O3 incorporation, J. Cryst. Growth 310 (2008) 3019–3023,
https://doi.org/10.1016/j.jcrysgro.2008.03.004.
References
[25] L. Yin, D. Chen, M. Feng, L. Ge, D. Yang, Z. Song, Hierarchical Fe2O3@WO3
nanostructures with ultrahigh specific surface areas: microwave-assisted synthesis
[1] J. Zhang, Z. Qin, D. Zeng, C. Xie, Metal-oxide-semiconductor based gas sensors: and enhanced H2S-sensing performance, RSC Adv. 5 (2014) 328–337, https://
screening, preparation, and integration, Phys. Chem. Chem. Phys. 19 (2017) doi.org/10.1039/C4RA10500A.
6313–6329, https://doi.org/10.1039/c6cp07799d. [26] N. Van Toan, C.M. Hung, N. Van Duy, N.D. Hoa, D. Thi, T. Le, Bilayer SnO2–WO3
[2] I. Elmi, S. Zampolli, E. Cozzani, F. Mancarella, G.C. Cardinali, Development of ultra- nanofilms for enhanced NH3 gas sensing performance, Mater. Sci. Eng. B 224
low-power consumption MOX sensors with ppb-level VOC detection capabilities for (2017) 163–170, https://doi.org/10.1016/j.mseb.2017.08.004.
emerging applications, Sensor. Actuator. B Chem. 135 (2008) 342–351, https:// [27] L. Yin, D. Chen, H. Zhang, G. Shao, B. Fan, R. Zhang, In situ formation of Au/SnO2
doi.org/10.1016/j.snb.2008.09.002. nanocrystals on WO3 nanoplates as excellent gas-sensing materials for H2S
[3] Y. Guo, X. wen Zhang, G. rong, Han, Investigation of structure and properties of N- detection, Mater. Chem. Phys. 148 (2014) 1099–1107, https://doi.org/10.1016/
doped TiO2 thin films grown by APCVD, Mater. Sci. Eng. B Solid State Mater. Adv. j.matchemphys.2014.09.025.
Technol. 135 (2006) 83–87, https://doi.org/10.1016/j.mseb.2006.08.031. [28] M.A. Gondal, M.A. Dastageer, L.E. Oloore, U. Baig, Laser induced selective photo-
[4] A. Gurlo, N. B^arsan, M. Ivanovskaya, U. Weimar, W. G€ opel, In2O3 and MoO3–In2O3 catalytic reduction of CO2 into methanol using In2O3-WO3 nano-composite,
thin film semiconductor sensors: interaction with NO2 and O3, Sensor. Actuator. B J. Photochem. Photobiol. Chem. 343 (2017) 40–50, https://doi.org/10.1016/
Chem. 47 (1998) 92–99, https://doi.org/10.1016/S0925-4005(98)00033-1. j.jphotochem.2017.04.016.
[5] T. Kida, A. Nishiyama, M. Yuasa, K. Shimanoe, N. Yamazoe, Highly sensitive NO2 [29] L. Yin, D. Chen, M. Hu, H. Shi, D. Yang, B. Fan, Microwave-assisted growth of In2O3
sensors using lamellar-structured WO3 particles prepared by an acidification nanoparticles on WO3 nanoplates to improve H2S-sensing performance, J. Mater.
method, Sensor. Actuator. B Chem. 135 (2009) 568–574, https://doi.org/10.1016/ Chem. A. 2 (2014) 18867–18874, https://doi.org/10.1039/C4TA03426K.
j.snb.2008.09.056. [30] K.W. Goh, M.R. Johan, Y.H. Wong, Enhanced structural properties of In2O3
[6] G.F. Fine, L.M. Cavanagh, A. Afonja, R. Binions, Metal oxide semiconductor gas nanoparticles at lower calcination temperature synthesised by co-precipitation
sensors in environmental monitoring, Sensors 10 (2010) 5469–5502, https:// method, Micro & Nano Lett. (2017) 1–6, https://doi.org/10.1049/mnl.2017.0540.
doi.org/10.3390/s100605469. [31] J. Cao, B. Luo, H. Lin, B. Xu, S. Chen, Thermodecomposition synthesis of WO3/
[7] A. Dey, Semiconductor metal oxide gas sensors : a review, Mater. Sci. Eng. B 229 H2WO4 heterostructures with enhanced visible light photocatalytic properties,
(2018) 206–217, https://doi.org/10.1016/j.mseb.2017.12.036. Appl. Catal., B 111–112 (2012) 288–296, https://doi.org/10.1016/
[8] P.T. Moseley, Progress in the development of semiconducting metal oxide gas j.apcatb.2011.10.010.
sensors : a review, Meas. Sci. Technol. 28 (2017), 082001, https://doi.org/ [32] C. Wang, D. Chen, X. Jiao, Flower-like In2O3 nanostructures derived from novel
10.1088/1361-6501/aa7443. precursor: synthesis, characterization, and formation mechanism, J. Phys. Chem. C
[9] D. Chen, X. Hou, H. Wen, Y. Wang, H. Wang, X. Li, R. Zhang, H. Lu, H. Xu, S. Guan, 113 (2009) 7714–7718, https://doi.org/10.1021/jp809855d.
J. Sun, L. Gao, The enhanced alcohol-sensing response of ultrathin WO3 nanoplates, [33] K. Jothivenkatachalam, S. Prabhu, A. Nithya, K. Jeganathan, Facile synthesis of
Nanotechnology 21 (2010) 035501–035512, https://doi.org/10.1088/0957-4484/ WO3 with reduced particle size on zeolite and enhanced photocatalytic activity,
21/3/035501. RSC Adv. 4 (2014) 21221–21229, https://doi.org/10.1039/c4ra01376j.
[10] N. Barsan, D. Koziej, U. Weimar, Metal oxide-based gas sensor research: how to? [34] D. Han, L. Zhai, F. Gu, Z. Wang, Highly sensitive NO2 gas sensor of ppb-level
Sensor. Actuator. B Chem. 121 (2007) 18–35, https://doi.org/10.1016/ detection based on In2O3 nanobricks at low temperature, Sensor. Actuator. B Chem.
j.snb.2006.09.047. 262 (2018) 655–663, https://doi.org/10.1016/j.snb.2018.02.052.
[11] H.J. Kim, J.H. Lee, Highly sensitive and selective gas sensors using p-type oxide [35] Z. Zhang, Z. Wen, Z. Ye, L. Zhu, Ultrasensitive ppb-level NO2 gas sensor based on
semiconductors: Overview, Sensor. Actuator. B Chem. 192 (2014) 607–627, WO3 hollow nanosphers doped with Fe, Appl. Surf. Sci. 434 (2018) 891–897,
https://doi.org/10.1016/j.snb.2013.11.005. https://doi.org/10.1016/j.apsusc.2017.10.074.
[12] J. Zhao, T. Yang, Y. Liu, Z. Wang, X. Li, Y. Sun, Y. Du, Y. Li, G. Lu, Enhancement of [36] F. Gu, C. Li, D. Han, Z. Wang, Manipulating the defect structure (VO) of In2O3
NO2 gas sensing response based on ordered mesoporous Fe-doped In2O3, Sensor. nanoparticles for enhancement of formaldehyde detection, ACS Appl. Mater.
Actuator. B Chem. 191 (2014) 806–812, https://doi.org/10.1016/ Interfaces 10 (2018) 933–942, https://doi.org/10.1021/acsami.7b16832.
j.snb.2013.09.118.

31

You might also like