Cr-Doped Hollow In2O3

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Alloys and Compounds 879 (2021) 160472

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Metal-organic framework-derived Cr-doped hollow In2O3 nanoboxes


with excellent gas-sensing performance toward ammonia ]]
]]]]]]
]]

⁎ ⁎
Jing Sun a, Yanzheng Wang b, Peng Song a, , Zhongxi Yang a, Qi Wang a,
a
School of Material Science and Engineering, University of Jinan, Jinan 250022, China
b
Shandong Academy of Building Research Co., LTD, China

a r t i cl e i nfo a bstr ac t

Article history: Transition metal ions doping is an effective method to improve the gas sensing performance of metal oxide
Received 21 December 2020 semiconductors. Herein, we report the synthesis of Cr-doped In2O3 samples by calcining metal-organic
Received in revised form 13 May 2021 frameworks (MOFs)-precursors prepared by a simple oil bath method. The experimental characterization
Accepted 17 May 2021
results show that the as-prepared samples exhibit uniform hollow nanoboxes with large specific surface
Available online 20 May 2021
areas. Through comparison, the 0.5 wt% Cr-doped In2O3 sample with complete and regular morphology was
selected for a series of gas-sensitive performance tests. The results show that the sensor has good detection
Keywords:
Indium oxide performance for ammonia (NH3). Gas-sensing tests revealed that the sensor exhibited excellent sensing
Cr-doping performance towards NH3 with low operating temperature (140 °C), high selectivity, good repeatability,
Nanobox relatively high response, low detection concentration (~ 1 ppm), and short response time (~ 1 s). The
Metal-organic framework generation of lattice defects and the increased conductivity after doping are the reasons for the excellent
Gas sensor performance of the sensor. This provides a new strategy for the research and preparation of high perfor­
Ammonia mance gas sensors.
© 2021 Elsevier B.V. All rights reserved.

1. Introduction However, sensors made solely from pure In2O3 often exhibit pro­
blems such as high operating temperatures, high detection limits,
In recent years, environmental pollution has become more and and low response values. To some extent, the application of indium
more serious. As a toxic gas pollutant, ammonia has caused wide­ oxide gas sensors is limited, so doping [15], compounding [16] or
spread concern [1,2]. Ammonia is a colorless gas with strong pun­ noble metal modification [17] is performed to improve the gas
gent odor. It is easy to liquefy and can be directly synthesized from sensitivity performance. Similarly, the preparation of nano-scale
nitrogen and hydrogen. Ammonia irritates the respiratory tract and samples is also conducive to the detection of gases [18–21].
corrodes the skin [3,4], and exposure to high concentrations of Metal-organic framework materials are composed of metal cations
ammonia may cause death. Due to ammonia is used in many ways, (divalent, trivalent, and tetravalent) and organic ligand molecules [22].
real-time monitoring of its concentration is particularly important. MOFs with different structures can be obtained by using different
As indicated in literatures that people should not be exposed to metal ions and organics, which is also in line with a major advantage
25 ppm ammonia atmosphere for more than 8 h or 35 ppm for more of MOFs – the controllable structure. At the same time, because of the
than 10 min [5–7]. Consequently, accurate detection of ammonia is developed pore structure, open active sites, and high specific surface
essential to protect human health. In the field of gas sensing, metal area [23,24], MOFs have been applied in more and more fields, espe­
oxides (MOS) have become a suitable material for the detection of cially gas detection and chemical sensing that are closely related to
toxic gases due to their advantages such as low production cost [8], human health [25]. Take an example, Wei et al. synthesized MOF-
small size [9], and long service life [10] compared to other materials. derived rGO/α-Fe2O3 nanocomposites with the simple solvothermal
As a typical n-type semiconductor with a wide bandgap [11–14], method, which reveal high-performance, excellent selectivity and
In2O3 shows very superior application in the field of gas sensing. quick response-recovery on triethylamine gas [26]. For instance,
Drobek et al. uniformly covered the ZIF-8 molecular sieve membrane
on the surface of ZnO nanowires. Alarmingly, the ZnO@ZIF-8-based
nanocomposite sensor prepared in this way exposed excellent se­

Corresponding authors. lectivity for H2 in comparison with the pristine ZnO nanowires sensor
E-mail addresses: mse_songp@ujn.edu.cn (P. Song),
mse_wangq@ujn.edu.cn (Q. Wang).
[27]. In view of this, the preparation of metal oxides using MOFs as

https://doi.org/10.1016/j.jallcom.2021.160472
0925-8388/© 2021 Elsevier B.V. All rights reserved.
J. Sun, Y. Wang, P. Song et al. Journal of Alloys and Compounds 879 (2021) 160472

Fig. 1. Growth process of the Cr-doped In2O3 hollow nanoboxes.

Fig. 2. Schematic diagram of a gas-sensitive element.

precursors or the combination of MOS and MOFs structure may have a pure In2O3 to acetone gas [28]. The Cr doped WO3 nanofibers syn­
promising effect on improving gas sensing performance. thesized by Zhu et al. can detect 3-hydroxy-2-butanone biomarker
Studies have also shown that doping rare earth elements or with a concentration as low as 0.05 ppm, indicating that the sensor
transition metal elements can effectively enhance the gas sensing prepared from this material is extremely promising, and provides
performance of metal oxides. For example, Wei et al. prepared Ce- new ideas for cheaper and portable 3H-2B detection devices [29].
doped hierarchical flower-like In2O3 microspheres by hydrothermal Meanwhile, it is also a good method to prepare nano-scale samples.
method and tested their gas-sensitive properties. The results in­ Now some people have prepared MOF-derived In2O3 hollow mi­
dicate that Ce-doped In2O3 exhibited a much higher response than crotubes [30–32], which have also been studied in the direction of

Fig. 3. (a) XRD patterns of pure In2O3 and Cr-doped In2O3 hollow nanoboxes with different doping levels; (b) partial enlarged detail in the range of 28°–53°.

2
J. Sun, Y. Wang, P. Song et al. Journal of Alloys and Compounds 879 (2021) 160472

Fig. 4. (a, b) SEM images of pure In2O3 hollow nanoboxes; (c, d) SEM images of 0.25 wt% Cr-doped In2O3 hollow nanoboxes; SEM images of 0.5 wt% Cr-doped In2O3 hollow
nanoboxes; (g, h) SEM images of 1 wt% Cr-doped In2O3 hollow nanoboxes.

gas sensing. But their length is about tens of microns, and the spe­ is smaller than In3+ (0.08 nm), the doping of Cr3+ to replace the po­
cific surface area is large. On the basis, it is necessary to improve the sition of In3+ in the unit cell and the occurrence of lattice distortion are
experimental scheme and reduce the sample size. the reasons for improving the sensitivity. In the following, the gas
In the paper, we use a simple oil bath method to prepare the sensing mechanism of Cr-doped In2O3 will be explained.
precursor nanoboxes, and obtained Cr-doped In2O3 hollow nanoboxes
by calcining them. After comparing the structure and morphological 2. Experimental
characterization of samples with different doping amounts, we se­
lected samples with a Cr doping amount of 0.5 wt% for gas-sensitivity 2.1. Preparation of precursor nanoboxes and Cr-doped In2O3 hollow
testing. The results indicate that 0.5 wt% Cr-doped In2O3 show ex­ nanoboxes
cellent gas sensing response to 10 ppm ammonia at 140 °C. In addition,
the sample exposed quick response-recovery and low detection limit In the typical process, 160 mL N, N-dimethylformamide was
(can detect 1 ppm ammonia). Since the ion radius of Cr3+ (0.0615 nm) measured and transferred into a 250 mL beaker labeled A, then

Fig. 5. (a) EDS energy spectrum diagram of 0.5 wt% Cr-doped In2O3 hollow nanoboxes; (b–d) EDS mapping images of 0.5 wt% Cr-doped In2O3 hollow nanoboxes.

3
J. Sun, Y. Wang, P. Song et al. Journal of Alloys and Compounds 879 (2021) 160472

Fig. 6. (a, b) TEM image of 0.5 wt% Cr-doped In2O3 hollow nanoboxes; (c) high-resolution TEM image of 0.5 wt% Cr-doped In2O3 hollow nanoboxes.

0.0730 g 1,4-benzenedicarboxylic acid and 0.0700 g Indium nitrate therein was poured out. The powder was collected by centrifugation
hydrate was dissolved in the N, N-dimethylformamide that has been and washing 3 times with absolute ethanol or deionized water, thus
measured. In addition, 200 mL N, N-dimethylformamide was mea­ we can obtain the precursor nanoboxes. Finally, Cr-doped In2O3
sured and transferred into a 250 mL beaker labeled B, then 200 mg hollow nanoboxes are obtained by calcining the precursor nano­
Chromium nitrate nonahydrate was added in the N, N-di­ boxes at 400 °C for 3 h in air atmosphere. The growth process of the
methylformamide. Ultimately, measure a certain amount of solution Cr-doped In2O3 hollow nanoboxes is shown in the Fig. 1.
from the B beaker (the corresponding Cr doping amounts are 0.25,
0.5 and 1 wt%) and transfer it into the A beaker, and stir until well. 2.2. Characterization
Pure samples can be prepared without adding chromium nitrate
solution. The Powder X-ray diffractometer (XRD, Bruker D8 Advance) was
Subsequent to this procedure, A beaker was placed in an oil bath used to determine the crystal structure of the sample. The micro­
for heating and heated to 120 °C for 30 min. After reaction, the A scopic morphology of the sample was characterized by a field
beaker was slowly cooled to room temperature and the mixture emission scanning electron microscope (FESEM, FEI Company,

Fig. 7. (a–d) Pore-size distribution (inset) and N2 adsorption-desorption isotherm of pure and 0.25, 0.5, 1 wt% Cr-doped In2O3 hollow nanoboxes.

4
J. Sun, Y. Wang, P. Song et al. Journal of Alloys and Compounds 879 (2021) 160472

Fig. 8. XPS spectra of 0.5 wt% Cr-doped In2O3 hollow nanoboxes: (a) survey spectrum, (b) In 3d spectrum, (c) Cr 2p spectrum.

QUANTA FEG 250). The element mapping spectra was exported also measured by a UV–vis (UV–vis, SHIMADZU UV-3600) spectro­
through the FESEM attachment. In order to obtain high-magnifica­ photometer.
tion images and the interplanar spacing of the sample, a transmis­
sion electron microscope (TEM, Hitachi H-800) was used for further 2.3. Sensor fabrication and test
testing. X-ray photoelectron spectroscopy (XPS, Thermo Scientific
Escalab 250Xi) can determine the valence state and proportions of The production process of the gas sensor is shown in Fig. 2. First
elements. The specific surface area of the sample is calculated using of all, take an alumina ceramic tube with four platinum wires (length
the Brunauer–Emmett–Teller (BET) method, which depends on N2 7 mm, diameter 1.5 mm) using tweezer, unfold the platinum wire
adsorption-desorption isotherms (Micromeritics Instrument Corp., with another tweezer, and connect the platinum wire to the com­
ASAP2460). Photoluminescence (PL) spectra were characterized on a ponent base with conductive silver paste. Afterwards, insert a small
PL spectrophotometer (Hitachi Model F-7000). The band gaps were Ni-Cr alloy coil into the ceramic tube, and also connect it to the base.

Fig. 9. (a) Sensor response of pure In2O3 and Cr-doped In2O3 hollow nanoboxes to 10 ppm ammonia gas corresponds to the operating temperature (100–200 °C); (b) Dynamic
response-recovery curve (10 ppm ammonia) of pure In2O3 at 160 °C and 0.5 wt% Cr-doped In2O3 hollow nanoboxes at 140 °C; (c) Response and recovery curves and the stability of
the sensors based on pure In2O3 and 0.5 wt% Cr-doped In2O3 hollow nanoboxes exposure to 10 ppm ammonia at 160 °C and 140 °C, respectively; (d) Concentration-dependent
response curves of pure In2O3 and 0.5 wt% Cr-doped In2O3 hollow nanoboxes sensors to 1–50 ppm ammonia.

5
J. Sun, Y. Wang, P. Song et al. Journal of Alloys and Compounds 879 (2021) 160472

After it is completely fixed, take about 30 mg Cr-doped In2O3 powder Table 1


and pour it into an agate mortar, add appropriate amount of absolute Compared ammonia-sensing properties of different sensors.
ethanol, fully grind and mix to make a slurry, then dip it with a fine Sensor material T (°C) NH3 Response Ref.
brush and coat it on a ceramic tube (need to completely cover the (ppm) (Ra/Rg)
gold electrode). Finally, after the coating layer is dried, the gas 0.5 wt% Cr-doped In2O3 140 10 11 This
sensing elements are aged at 200 °C for 24 h [26,33]. work
The gas sensing tests use the WS-30A system (Winsen 10% Pd-WO3 300 50 8.6 [46]
Co3O4 hierarchical 160 100 11.2 [52]
Electronics Co. Ltd., Zhengzhou, China) and the sensitivity of the
nanorods
material can be obtained by calculation under computer control. Co3O4/SnO2 Nanospheres 200 50 13.6 [53]
Formaldehyde and ammonia aqueous solution, methanol, ethanol, WO3@SnO2 nanosheets 200 15 1.5 [54]
and acetone are selected from analytically pure liquids that are in­ WO3-W18O49 nanowire 250 100 10.3 [55]
jected into the heating plate through a micro-injector and heated, bundle
Pr-SnS2/ZnS nanoflower 160 50 14.03 [56]
and mixed with air after evaporation to achieve the desired gas Flower-shaped SnS2 200 100 7.4 [57]
environment. The relationship between the injected liquid volume SnO2 hierarchical NWs 300 500 10.9 [58]
and the gas concentration is calculated by the following for­ 1.0% Pt-WO3 125 200 13.61 [59]
mula [34]:
V·C·M comparing with standard cards, we found that these peaks corre­
Vx =
24.2·d·p spond to chromium oxides (including C2CrO4, Cr5O12 and Cr8O21),
which may be caused by excessive Cr3+ content. At the same time, as
where Vx (μL) is the injection volume, V (L) is the volume of sensing the amount of doping increases, the sharpness of the diffraction
chamber, C (ppm) is the concentration of target gas, M (g/mol) is the peak decreases significantly, indicating that doping will cause a
molecular weight, d (g/mL) is the density and P is the weight ratio in decrease in crystallinity. In Fig. 3(b), the diffraction peaks corre­
aqueous solution. sponding to the (222) and (440) crystal planes move to a large angle,
When the gas sensor is exposed to the air atmosphere, the re­ because the ion radius of Cr3+ is smaller than that of In3+.
sistance value is expressed by Ra, and when exposed to the atmo­ The morphology and nanostructure of pure In2O3 and Cr-doped
sphere of target gas, the resistance value is expressed by Rg. The In2O3 samples were investigated by SEM. As can be seen in Fig. 4(a
ratio of these two resistance values (Ra/Rg) is defined as the sensi­ and b), the pure In2O3 nanoboxes are in the shape of a hexagonal
tivity of the sensor [35]. The time required for the change in re­ short column with a diameter of about 465 nm and uniform size.
sistance value to reach 90% of the final equilibrium value during gas Fig. 4(c and d) presents the typical SEM images of 0.25 wt% Cr-doped
diffusion and volatilization is defined as the response and recovery In2O3, showing hollow nanoboxes with the diameter of approxi­
time [36]. mately 507 nm. As shown in the Fig. 4(e and f) which corresponds to
0.5 wt% Cr-doped In2O3 with the diameter of about 663 nm, the sides
3. Results and discussion of the nanobox began to roughen, no longer smooth, and a hole
structure appeared at the same time. The SEM images of 1 wt% Cr-
3.1. Structural and morphological characteristics doped In2O3 is shown in Fig. 4(g and h). It can be clearly seen that
the diameter is about 874 nm, the edges of the nanobox shrink and
For obtain the phase purity and crystal structure of the materials, the thickness becomes larger. In particular, the dispersibility of the
X-ray diffraction analysis was performed on Cr-doped In2O3 samples. sample begin to deteriorate. Therefore, 0.5 wt% Cr-doped In2O3
From the bottom to the top in Fig. 3(a), the In2O3 samples with Cr sample was selected for other characterization.
doping amounts of 0 wt%, 0.25 wt%, 0.5 wt% and 1 wt% are shown in. According to the EDS results (Fig. 5(a–d)), which clearly indicate
It can be seen that when the doping amount is small, the diffraction the distribution of elements in the 0.5 wt% Cr-doped In2O3 hollow
peak position corresponding to the samples are consistent with the nanobox. Fig. 5(a) shows that the product consists of three elements
indium oxide standard card (JCPDS No. 06-0416). However, when the (O, Cr, and In), the peak of Cr element verifies the presence of Cr. In
Cr doping exceeds 1 wt%, will appear many miscellaneous peaks. By

Fig. 10. (a) Sensor responses to different gases (10 ppm) of 0.5 wt% Cr-doped In2O3 hollow nanoboxes sensor at 140 °C; (b) Long-term stability of 0.5 wt% Cr-doped In2O3 hollow
nanoboxes sensor on successive exposure to 10 ppm ammonia at 140 °C.

6
J. Sun, Y. Wang, P. Song et al. Journal of Alloys and Compounds 879 (2021) 160472

Fig. 11. Schematic diagrams for possible gas sensing mechanism of Cr-doped In2O3 hollow nanoboxes.

addition, Fig. 5(d) further proves the presence of Cr in the sample. It 444.3 eV and 531.3 eV, corresponding to In 3d5/2 and In 3d3/2 [34,39].
can be seen that Cr is very uniformly distributed in the resulting Moreover, the XPS spectrum of Cr 2p is depicted in Fig. 8(c). It il­
product, indicating that Cr successfully entered the In2O3 lattice. In lustrates that the Cr 2p2/3 at 576.2 eV and Cr 2p2/3 at 586.1 eV, thus
order to observe the morphological structure at a more microscopic determining the valence state of the dopant as Cr3+ [40].
scale, the samples need to be tested by TEM.
As indicated in the TEM images of 0.5 wt% Cr-doped In2O3 sam­ 3.2. Gas sensing properties
ples (Fig. 6(a and b)), it can be observed that the hollow structure of
the nanobox and the wall of the nanobox is found to be very thin, The sensor was made using all samples. It is well known that
only about 36 nm. The thickness of the entire nanobox is approxi­ operating temperature is a principal factor to affect the response of
mately 660 nm. The hollow structure is conducive to gas diffusion the sensor. Fig. 9(a) presents the responses of sensors to 10 ppm
and transmission, so it can produce high gas sensitivity. According to ammonia at various operating temperatures, the responses were
the HRTEM images (Fig. 6(c)) we can clearly see that the interplanar measured in the range of 100–200 °C with 20 °C as interval. It is clear
spacing exposed by Cr-doped In2O3 hollow nanobox was 0.308 nm, that as the temperature increases, the response of sensor experience
and it was found that it matches well with the (222) crystal plane by a climb firstly and then decreases gradually. Among them, the sensor
comparing with In2O3 standard card. prepared from the 0.5 wt% Cr-doped In2O3 shows the best perfor­
The Brunauer–Emmett–Teller (BET) method was usually used to mance for ammonia, and has a response value of 11.11 at 140 °C,
calculate the specific surface area of the material. The Fig. 7(a–d) which is higher than sensors based on other samples (the sensitivity
indicates that the pore-size distribution (inset) and N2 adsorption- of the pure In2O3 sensor is 3.9, the sensitivity of the 0.25 wt% Cr-
desorption isotherm, respectively. The N2 adsorption-adsorption doped In2O3 sensor is 9.2, the sensitivity of the 1 wt% Cr-doped In2O3
isotherm could be appeared as typical type IV isotherm which sensor is 8.8, and the optimum working temperature for the pure
shown that the obtained samples contained the mesoporous struc­ In2O3 sensor and 0.25 wt% Cr-doped In2O3 sensor is 160 °C). This is
ture [37,38]. The Brunauer–Emmett–Teller surface areas of the pure because at the optimum operating temperature, the adsorbed
In2O3, 0.25, 0.5 and 1 wt% Cr-doped In2O3 hollow nanoboxes were oxygen on the surface of the material has sufficient capacity to
calculated to be 46.4, 66.7, 51.3 and 58.2 m3/g. At the same time, in participate in the reaction. When the temperature is lower than the
the figure of pore-size distribution, it can be obtained that the pore optimal temperature, the surface activity of the material is low and
sizes with the largest volume ratio are located at 6.68, 6.26, 5.79, and the amount of adsorbed oxygen that can react with ammonia is
4.53 nm. For MOF-derived oxides, the volatilization of organics small. If the temperature exceeds the optimal temperature, the
during the calcination process is responsible for its large specific desorption rate eventually exceeds the adsorption rate, resulting in a
surface area. When the specimen are used for gas sensing, this high decrease in the amount of adsorbed oxygen, which leads to the
specific surface area can promote gas diffusion, adsorption and decrease in sensitivity [26]. It is worth noting that compared with
desorption process. Therefore, we can infer that the Cr-doped In2O3 pure In2O3 sensors, the optimal operating temperature has dropped
hollow nanoboxes structure derived from MOF are highly likely to from 160 °C to 140 °C, and the sensitivity to ammonia has also been
possess great gas sensing performance. greatly improved.
To further explore the chemical state of the elements, the sam­ As shown in Fig. 9(b), we plot the typical response/recover curve
ples were subjected to XPS analysis. From the full survey spectrum toward 10 ppm ammonia at 140 °C. In this situation, the results show
(Fig. 8(a)), only three elements present in the as-prepared samples. that the sensor based on the 0.5 wt% Cr-doped In2O3 has rapid re­
In Fig. 8(b), there are two peaks where the binding energy is sponse (1 s)/recover times (18 s).

7
J. Sun, Y. Wang, P. Song et al. Journal of Alloys and Compounds 879 (2021) 160472

Fig. 12. O 1s spectra of (a) pure In2O3 hollow nanoboxes, (b) 0.25 wt% Cr-doped In2O3 hollow nanoboxes, (c) 0.5 wt% Cr-doped In2O3 hollow nanoboxes and (d) 1 wt% Cr-doped
In2O3 hollow nanoboxes; (e) Room-temperature PL spectra pure In2O3 and 0.5 wt% Cr-doped In2O3 hollow nanoboxes; (f) UV–vis diffuse reflectance spectra of pure In2O3 and
0.5 wt% Cr-doped In2O3 hollow nanoboxes.

We can reach conclusion that the Cr-doped In2O3 gas sensor shows the dynamic response curve to ammonia with concentra­
displays excellent reproducibility from three reversible cycles of tions of 1–50 ppm. Visibly, as the gas concentration increases, the
response-recovery curve (Fig. 9(c)) which is due to the maximum sensitivity also gradually increases. The sensor has a sensitivity of
response value under 10 ppm ammonia. 2 when bathed at a gas concentration of 1 ppm, proving that the
At the optimum temperature, the sensitivity of Cr-doped In2O3 Cr-doped In2O3 gas sensor can detect lower concentrations of
to different concentrations of ammonia gas can be tested. Fig. 9(d) ammonia. When the sensor is at 50 ppm test gas, the sensitivity

8
J. Sun, Y. Wang, P. Song et al. Journal of Alloys and Compounds 879 (2021) 160472

The Fig. 11 describes the gas-sensing mechanism. In fact, the


improvement of In2O3 gas sensing performance by Cr doping can be
attributed to the defects generated after Cr3+ enters the In2O3 lattice.
These defects enable the material to form more active sites to cap­
ture oxygen in the air. Fig. 12(a) shows the O 1s spectrum of pure
In2O3. It is worth noting that the binding energies of 529.9, 531.0 and
531.9 eV exist three high intense peaks, attribute to the lattice
oxygen (OL), oxygen vacancy (OV) and chemisorbed oxygen (OC),
respectively [47,48]. Fig. 12(b–d) shows O 1s spectra of In2O3 with Cr
doping of 0.25, 0.5 and 1 wt%, respectively. It is known from calcu­
lation that the proportion of oxygen adsorption and oxygen vacancy
of the 0.5 wt% Cr-doped In2O3 is about 47.0% and 31.2%, which are
higher than other samples (the adsorbed oxygen proportions of the
pure, 0.5, and 1 wt% Cr-doped In2O3 are about 28.02%, 16.1% and
19.0%; the oxygen vacancy proportions of pure, 0.5, and 1 wt% Cr-
doped In2O3 are about 21.1%, 27.9% and 27.0%). In order to analyze
the difference in oxygen vacancies between In2O3 and 0.5 wt% Cr-
doped In2O3, we supplemented the photoluminescence (PL) mea­
Fig. 13. Change curves of the resistance of pure In2O3 and 0.5 wt% Cr-doped In2O3 surement and plotted in Fig. 12(c). The peaks located at the range of
hollow nanoboxes. 450–530 nm are caused by oxygen vacancies. Free electrons are in
contact with oxygen vacancies and react with photo-excited holes,
value can reach more than 25. The reason is that as the con­ resulting in green emission in this wavelength range. Obviously, the
centration increases, more ammonia gas molecules adsorb to the peak of the doped sample is stronger than that of pure indium oxide,
sensor surface and react with the sensor material, thereby in­ indicating that the sample contains more oxygen vacancies after
creasing the sensitivity. doping [49]. Fig. 12(d) shows the UV–vis diffuse reflectance spectra
In order to clearly reflect the good response of the 0.5 wt% Cr- of pure In2O3 and 0.5 wt% Cr-doped In2O3 samples. It can be seen
doped In2O3 to ammonia, the sensor toward 10 ppm other gases that the transition of electrons from the valence band to the con­
such as formaldehyde, acetone, methanol and ethanol and the duction band causes the samples to have strong absorption in the
graph were investigated at the optimal working temperature and wavelength range of 350–450 nm. The calculation shows that the
plotted in Fig. 10(a). Obviously, the sample has a sensitivity of band gap value of pure In2O3 is 2.92, while that of the doped sample
11–10 ppm ammonia and less than 4 for other gases. Furthermore, is 2.87. According to the literature [50], an increase in the oxygen
we also tested the long-term stability of the sensor based on 0.5 wt vacancy concentration will reduce the band gap, which can also
% Cr-doped In2O3 (Fig. 10(b)). The result shows that the response prove that the doped sample has high oxygen vacancies. Oxygen
value obtained is stable after 15 days of continuous exposure vacancies containing unpaired electrons can be used as sites for gas
under 10 ppm ammonia, indicating that the long-term stability is adsorption and reaction, resulting in more oxygen adsorption on the
superior. As expected, 0.5 wt% Cr-doped In2O3 is a promising am­ surface. Secondly, the introduction of Cr surprisingly increases the
monia detection material compared to other materials (Table 1). conductivity of the material [51]. Fig. 13 is the change curve of the
resistance value of 0.5 wt% Cr-doped In2O3 and pure In2O3. The dif­
ference between the resistance value of 0.5 wt% Cr-doped In2O3 in
3.3. Gas sensing mechanism air and its resistance value in ammonia is greater than that of pure
In2O3, so more electrons are involved in the transfer during the gas
The gas sensing mechanism of In2O3 (an n-type semiconductor adsorption and desorption process. It can be proved that the con­
material) depends on its different resistance values when exposed to ductivity has increased.
the air and the target gas environment. Upon exposure to air, a part
of oxygen molecules will be adsorbed on the surface of the material,
4. Conclusions
while capturing free electrons, thus forming adsorbed oxygen. The
reaction that occurs when oxygen ions (O2−, O−) are formed is re­
In this work, Cr-doped In2O3 hollow nanoboxes with different Cr
flected by the following formulas [41,42]:
doping amounts were successfully synthesized via a simple oil bath
O2 (gas) ↔ O2 (ads) (1) method and annealing treatment. It can be found that all samples
− − have well hollow structure, which is conducive to gas adsorption and
O2 (ads) + e ↔ O2 (ads) (below 100 °C) (2)
transmission. In addition, the incorporation of Cr3+ ions increases the
O2− (ads) + e− ↔ 2O− (ads) (100–300 °C) (3) number of oxygen vacancy defects and active sites on the surface.
The gas-sensing experiments displayed that the as-obtained 0.5 wt%
In the process, the resistance value of the material will increase, Cr-doped In2O3 exhibited the extremely fast response (1 s), high-
due to reduction of electrons, so that an electron depletion layer is performance gas sensing property for ultra-low concentration
formed on the surface of the material. After the Cr-doped In2O3 (1 ppm) NH3 detections at low temperature (140 °C). Therefore, Cr-
sensors are placed in ammonia surroundings, electrons will be re­ doped In2O3 hollow nanoboxes could be a promising material to
leased when the ammonia gas reacts with oxygen ions generated on detect NH3.
the surface of the material (Eq. (4)). In this case, the resistance value
of the material will decrease, and the thickness of the electron de­
pletion layer of the material will decrease [4,43–46]. CRediT authorship contribution statement

2NH3 + 3O− → N2 + 3H2O + 3e− (4) Jing Sun: Writing - original draft, Methodology, Investigation.
Normally, the resistance will gradually return to the initial value Yanzheng Wang: Validation, Formal analysis. Peng Song: Writing -
after the target gas reaction ends. review & editing, Resources. Zhongxi Yang: Project administration.
Qi Wang: Supervision, Funding acquisition.

9
J. Sun, Y. Wang, P. Song et al. Journal of Alloys and Compounds 879 (2021) 160472

Declaration of Competing Interest [22] X.F. Wang, X.Z. Song, K.M. Sun, L. Cheng, W. Ma, MOFs-derived porous nano­
materials for gas sensing, Polyhedron 152 (2018) 155–163.
[23] Y.S. Liu, H. Jiang, J.Y. Hao, Y.L. Liu, H.B. Shen, W.Z. Li, J. Li, Metal-organic frame­
The authors declare that they have no known competing fi­ workderived reduced graphene oxide-supported ZnO/ZnCo2O4/C hollow nano­
nancial interests or personal relationships that could have appeared cages as cathode catalysts for aluminum-O2 batteries, ACS Appl. Mater.
to influence the work reported in this paper. Interfaces 9 (2017) 31841–31852.
[24] J.N. Zhang, H. Lu, L.Z. Zhang, D.Y. Leng, Y.Y. Zhang, W. Wang, Y. Gao, H.B. Lua,
J.Z. Gao, G.Q. Zhu, Z.B. Yang, C.L. Wang, Metal–organic framework-derived ZnO
Acknowledgment hollow nanocages functionalized with nanoscale Ag catalysts for enhanced
ethanol sensing properties, Sens. Actuators B Chem. 291 (2019) 458–469.
[25] W.H. Li, X.F. Wu, N. Han, J.Y. Chen, X.H. Qian, Y.Z. Deng, W.X. Tang, Y.F. Chen,
This work was financially supported by National Natural Science MOF-derived hierarchical hollow ZnO nanocages with enhanced low-con­
Foundation of China (No. 61102006), and Natural Science Foundation centration VOCs gas-sensing performance, Sens. Actuators B Chem. 225 (2016)
of Shandong Province, China (Nos. ZR2018LE006 and ZR2015EM019). 158–166.
[26] Q. Wei, J. Sun, P. Song, J. Li, Z.X. Yang, Q. Wang, MOF-derived α-Fe2O3 porous
spindle combined with reduced graphene oxide for improvement of TEA sensing
References performance, Sens. Actuators B Chem. 304 (2020) 127306.
[27] M. Drobek, J.H. Kim, M. Bechelany, C. Vallicari, A. Julbe, S.S. Kim, MOF-based
[1] Marcelo Eising, Carlos Eduardo Cava, Rodrigo Villegas Salvatierra, Aldo José membrane encapsulated ZnO nanowires for enhanced gas sensor selectivity, ACS
Gorgatti Zarbin, Lucimara Stolz Roman, Doping effect on self-assembled films of Appl. Mater. Interfaces 8 (2016) 8323–8328.
polyaniline and carbon nanotube applied as ammonia gas sensor, Sens. [28] D.D. Wei, Z.S. Huang, L.W. Wang, X.H. Chuai, S.M. Zhang, G.Y. Lu, Hydrothermal
Actuators B Chem. 245 (2017) 25–33. synthesis of Ce-doped hierarchical flower-like In2O3 microspheres and their
[2] D.Z. Zhang, C.X. Jiang, Y.E. Sun, Room-temperature high-performance ammonia excellent gas-sensing properties, Sens. Actuators B Chem. 255 (2018) 1211–1219.
gas sensor based on layer-by-layer self-assembled molybdenum disulfide/zinc [29] Z.Y. Zhu, L.J. Zheng, S.Z. Zheng, J. Chen, M.H. Liang, Y.T. Tian, D.C. Yang, Cr doped
oxide nanocomposite film, J. Alloy. Compd. 698 (2017) 476–483. WO3 nanofibers enriched with surface oxygen vacancies for highly sensitive
[3] Q.Q. Wang, H.W. Fu, J.J. Ding, C. Yang, S. Wang, Sensitivity enhanced microfiber detection of the 3-hydroxy-2-butanone biomarker, J. Mater. Chem. A 6 (2018)
interferometer ammonia gas sensor by using WO3 nanorods coatings, Opt. Laser 21419–21427.
Technol. 125 (2020) 1–5. [30] D.Z. Zhang, Z.M. Yang, P. Li, X.Y. Zhou, Ozone gas sensing properties of metal-
[4] Madhukar Poloju, Nagabandi Jayababu, M.V. Ramana Reddy, Improved gas organic frameworks-derived In2O3 hollow microtubes decorated with ZnO na­
sensing performance of Al doped ZnO/CuO nanocomposite based ammonia gas noparticles, Sens. Actuators B Chem. 301 (2019) 127081.
sensor, Mater. Sci. Eng. B Solid 227 (2018) 61–67. [31] Y.S. Liu, X.P. Liu, Y.B. Wang, R. Wang, T. Zhang, Metal-organic-framework-derived
[5] M.L. Yin, Z. Zhu, Mesoporous NiO as an ultra-highly sensitive and selective gas In2O3 microcolumnar structures embedded with Pt nanoparticles for NO2 de­
sensor for sensing of trace ammonia at room temperature, J. Alloy. Compd. 789 tection near room temperature, Ceram. Int. 45 (2019) 9820–9828.
(2017) 941–947. [32] D. Han, P. Song, S. Zhang, H.H. Zhang, Q. Xu, Q. Wang, Enhanced methanol gas-
[6] R. Sankar Ganesh, M. Navaneethan, V.L. Patil, S. Ponnusamy, sensing performance of Ce-doped In2O3 porous nanospheres prepared by hy­
C. Muthamizhchelvan, S. Kawasaki, P.S. Patil, Y. Hayakawa, Sensitivity enhance­ drothermal method, Sens. Actuators B 216 (2015) 488–496.
ment of ammonia gas sensor based on Ag/ZnO flower and nanoellipsoids at low [33] L. Liu, Y.T. Zhao, P. Song, Z.X. Yang, Q. Wang, ppb level triethylamine detection of
temperature, Sens. Actuators B Chem. (2018) 672–683. yolk-shell SnO2/Au/Fe2O3 nanoboxes at low-temperature, Appl. Surf. Sci. 476
[7] I.P. Liu, C.H. Chang, T.C. Chou, K.W. Lin, Ammonia sensing performance of a (2019) 391–401.
platinum nanoparticle-decorated tungsten trioxide gas sensor, Sens. Actuators B [34] B.J. Wang, S.Y. Ma, S.T. Pei, X.L. Xu, P.F. Cao, J.L. Zhang, R. Zhang, X.H. Xu, T. Han,
Chem. 291 (2019) 148–154. Cellulolytic bacterium characterization and genome functional analysis: an at­
[8] L. Morawska, P.K. Thai, X. Liu, A.A. Sakyi, G. Ayoko, B. Christensen, Applications of tempt to lay the foundation for waste management, Bioresour. Technol. 321
low-cost sensing technologies for air quality monitoring and exposure assess­ (2021) 124462.
ment: how far have they gone? Environ. Int. 116 (2018) 286–299. [35] P. Song, D. Han, H.H. Zhang, J. Li, Z.X. Yang, Q. Wang, Hydrothermal synthesis of
[9] J. Liu, S. Li, B. Zhang, Y.L. Wang, Y. Gao, X.S. Liang, Y. Wang, G.Y. Lu, Flower-like porous In2O3 nanospheres with superior ethanol sensing properties, Sens.
In2O3 modified by reduced graphene oxide sheets serving as a highly sensitive Actuators B 196 (2014) 434–439.
gas sensor for trace NO2 detection, J. Colloid Interface Sci. 504 (2017) 206–213. [36] F.E. Annanouch, G. Bouchet, P. Perrier, N. Morati, C. Reynard-Carettea, K. Aguir,
[10] F. Aliyu, T. Sheltami, Development of an energy-harvesting toxic and combus­ V. Martini-Laithier, M. Bendahan, Hydrodynamic evaluation of gas testing
tible gas sensor for oil and gas industries, Sens. Actuators B Chem. 231 (2016) chamber: simulation, experiment, Sens. Actuators B Chem. 290 (2019) 598–606.
265–275. [37] F. Li, T. Zhang, X. Gao, R. Wang, B.H. Li, Coaxial electrospinning heterojunction
[11] D.M. Han, L.L. Zhai, F.B. Gu, Z.H. Wang, Highly sensitive NO2 gas sensor of ppb- SnO2/Au-doped In2O3 core-shell nanofibers for acetone gas sensor, Sens.
level detection based on In2O3 nanobricks at low temperature, Sens. Actuators B Actuators B Chem. 252 (2017) 822–830.
Chem. 262 (2018) 655–663. [38] B.Q. Han, J.H. Wang, W.Y. Yang, X.W. Chen, H.R. Wang, J.L. Chen, C. Zhang,
[12] S. Zhang, P. Song, J. Zhang, H.H. Yan, J. Li, Z.X. Yang, Q. Wang, Highly sensitive J.H. Sun, X.Y. Wei, Hydrothermal synthesis of flower-like In2O3 as a chemir­
detection of acetone using mesoporous In2O3 nanospheres decorated with Au esistive isoprene sensor for breath analysis, Sens. Actuators B Chem. 309 (2020)
nanoparticles, Sens. Actuators B Chem. 242 (2017) 983–993. 127788.
[13] F.L. Gong, Y.Y. Gong, H.Z. Liu, M.L. Zhang, Y.H. Zhang, F. Li, Porous In2O3 nano­ [39] R.K. Chava, H.Y. Cho, J.M. Yoon, Y.T. Yu, Fabrication of aggregated In2O3 nano­
cuboids modified with Pd nanoparticles for chemical sensors, Sens. Actuators B spheres for highly sensitive acetaldehyde gas sensors, J. Alloy. Compd. 772
Chem. 223 (2016) 384–391. (2019) 834–842.
[14] W.H. Zhang, W.C. Zhang, B. Chen, R. Shao, R.F. Guan, W.D. Zhang, Q.F. Zhang, [40] K. Bunpang, A. Wisitsoraat, A. Tuantranont, S. Singkammo, S. Phanichphant,
G.H. Hou, L. Yue, Controllable biomolecule-assisted synthesis and gas sensing C. Liewhiran, Highly selective and sensitive CH4 gas sensors based on flame-
properties of In2O3 micro/nanostructures with double phases, Sens. Actuators B spray-made Cr-doped SnO2 particulate films, Sens. Actuators B Chem. 291 (2019)
Chem. 239 (2017) 270–278. 177–191.
[15] Asieh Montazeri, Farid Jamali-Sheini, Enhanced ethanol gas-sensing perfor­ [41] Q. Wei, J. Sun, P. Song, J. Li, Z.X. Yang, Q. Wang, Spindle-like Fe2O3/ZnFe2O4
mance of Pb-doped In2O3 nanostructures prepared by sonochemical method, porous nanocomposites derived from metal-organic frameworks with excellent
Sens. Actuators B Chem. 242 (2017) 778–791. sensing performance towards triethylamine, Sens. Actuators B Chem. 317 (2020)
[16] Q. Chen, S.Y. Ma, X.L. Xu, H.Y. Jiao, G.H. Zhang, L.W. Liu, P.Y. Wang, D.J. Gengzang, 128205.
H.H. Yao, Optimization ethanol detection performance manifested by gas sensor [42] S.D. Zhang, M.J. Yang, K.Y. Liang, A. Turak, B.X. Zhang, D. Meng, C.X. Wang,
based on In2O3/ZnS rough microspheres, Sens. Actuators B Chem. 264 (2018) F.D. Qu, W.L. Cheng, M.H. Yang, An acetone gas sensor based on nanosized Pt-
263–278. loaded Fe2O3 nanocubes, Sens. Actuators B Chem. 290 (2019) 59–67.
[17] Z.H. Wang, G.l. Men, R.x. Zhang, F.b. Gu, D.m. Han, Pd loading induced excellent [43] D.D. Trung, N.D. Cuong, K.Q. Trung, T.D. Nguyen, N.V. Toan, C.M. Hung, N.V. Hieu,
NO2 gas sensing of 3DOM In2O3 at room temperature, Sens. Actuators B Chem. Controlled synthesis of manganese tungstate nanorods for highly selective NH3
263 (2018) 218–228. gas sensor, J. Alloy. Compd. 735 (2018) 787–794.
[18] Y. Meng, G. Liu, A. Liu, Z. Guo, W. Sun, F. Shan, Photochemical activation of [44] Y.L. Pan, W.C. Liu, Study of a platinum (Pt) nanoparticle (NP)/vanadium pent­
electrospun In2O3 nanofibers for high-performance electronic devices, ACS Appl. oxide (V2O5) thin film-based ammonia gas sensor, IEEE Trans. Electron Devices
Mater. Interfaces 9 (2017) 10805–10812. 67 (2020) 2126–2132.
[19] R.K. Chava, H.Y. Cho, J.M. Yoon, Y.T. Yu, Fabrication of aggregated In2O3 nanospheres [45] S. Büyükköse, Highly selective and sensitive WO3 nanoflakes based ammonia
for highly sensitive acetaldehyde gas sensors, J. Alloy. Compd. 772 (2019) 834–842. sensor, Mater. Sci. Semicond. Process. 110 (2020) 104969.
[20] Q. Mi, D.Z. Zhang, X.X. Zhang, D.Y. Wang, Highly sensitive ammonia gas sensor [46] C. Castillo, G. Cabello, B. Chornik, Y. Huentupil, G.E. Buono-Core, Characterization
based on metal-organic frameworks-derived CoSe2@nitrogen-doped amorphous of photochemically grown Pd loaded WO3 thin films and its evaluation as am­
carbon decorated with multi-walled carbon nanotubes, J. Alloy. Compd. 860 monia gas sensor, J. Alloy. Compd. 825 (2020) 154166.
(2021) 158252. [47] D. Bekermann, A. Gasparotto, D. Barreca, C. Maccato, E. Comini, C. Sada,
[21] R.K. Chava, S.Y. Oh, Y.T. Yu, Enhanced H2 gas sensing properties of Au@In2O3 G. Sberveglieri, A. Devi, R.A. Fischer, Co3O4/ZnO nanocomposites: from plasma
core–shell hybrid metal–semiconductor heteronanostructures, CrystEngComm synthesis to gas sensing applications, ACS Appl. Mater. Interfaces 4 (2012)
18 (2016) 3655–3666. 928–934.

10
J. Sun, Y. Wang, P. Song et al. Journal of Alloys and Compounds 879 (2021) 160472

[48] D. Bekermann, A. Gasparotto, D. Barreca, A. Devi, R.A. Fischer, p-Co3O4/n-ZnO, [54] K.P. Yuan, L.Y. Zhu, J.H. Yang, C.Z. Hang, J.J. Tao, H.P. Ma, A.Q. Jiang, D.W. Zhang,
obtained by PECVD, analyzed by X-ray photoelectron spectroscopy, Surf. Sci. H.L. Lu, Precise preparation of WO3@SnO2 core shell nanosheets for efficient NH3
Spectra 18 (2011) 36–45. gas sensing, J. Colloid Interface Sci. 568 (2020) 81–88.
[49] J.S. Ri, X.W. Lia, C.L. Shao, Y. Liu, C.H. Han, X.H. Li, Y.C. Liu, Sn-doping induced [55] Y. Xiong, Z.Y. Zhu, T.C. Guo, H. Li, Q.Z. Xue, Synthesis of nanowire bundle-like
oxygen vacancies on the surface of the In2O3 nanofibers and their promoting WO3-W18O49 heterostructures for highly sensitive NH3 sensor application, J.
effect on sensitive NO2 detection at low temperature, Sens. Actuators B Chem. Hazard. Mater. 353 (2018) 290–299.
317 (2020) 128194. [56] Q.X. Zhang, S.Y. Ma, R. Zhang, K.M. Zhu, Y. Tie, S.T. Pei, Optimization NH3 sensing
[50] F.B. Gu, C.J. Li, D.M. Han, Z.H. Wang, Manipulating the defect structure (VO) of performance manifested by gas sensor based on Pr-SnS2/ZnS hierarchical na­
In2O3 nanoparticles for enhancement of formaldehyde detection, ACS Appl. noflowers, J. Alloy. Compd. 807 (2019) 151650.
Mater. Interfaces 10 (2018) 933–942. [57] Y. Xiong, W.W. Xu, D.G. Ding, W.B. Lu, L. Zhu, Z.Y. Zhu, Y. Wang, Q.Z. Xue, Ultra-
[51] H. Baqiah, N.B. Ibrahim, M.H. Abdi, S.A. Halim, Electrical transport, micro­ sensitive NH3 sensor based on flower-shaped SnS2 nanostructures with sub-ppm
structure and optical properties of Cr-doped In2O3 thin film prepared by sol–gel detection ability, J. Hazard. Mater. 341 (2018) 159–167.
method, J. Alloy. Compd. 575 (2013) 198–206. [58] L.V. Thong, L.T. Ngoc Loan, N.V. Hieu, Comparative study of gas sensor perfor­
[52] J.N. Deng, R. Zhang, L.L. Wang, Z. Lou, T. Zhang, Enhanced sensing performance of the mance of SnO2 nanowires and their hierarchical nanostructures, Sens. Actuators
Co3O4 hierarchical nanorods to NH3 gas, Sens. Actuators B Chem. 209 (2015) 449–455. B Chem. 150 (2010) 112–119.
[53] L.L. Wang, Z. Lou, R. Zhang, T.T. Zhou, J.N. Deng, T. Zhang, Hybrid Co3O4/SnO2 [59] Y.L. Wang, J. Liu, X.B. Cui, Y. Gao, J. Ma, Y.F. Sun, P. Sun, F.M. Liu, X.S. Liang,
core−shell nanospheres as real-time rapid response sensors for ammonia gas, T. Zhang, G.Y. Lu, NH3 gas sensing performance enhanced by Pt-loaded on me­
ACS Appl. Mater. Interfaces 8 (2016) 6539–6545. soporous WO3, Sens. Actuators B Chem. 238 (2017) 473–481.

11

You might also like