You are on page 1of 20

Accepted Manuscript

Full Length Article

Design of superior ethanol gas sensor based on indium oxide/molybdenum di-


sulfide nanocomposite via hydrothermal route

Xiaojing Liu, Li Jiang, Xiumei Jiang, Xueying Tian, Ying Huang, Peiyu Hou,
Shouwei Zhang, Xijin Xu

PII: S0169-4332(18)30790-6
DOI: https://doi.org/10.1016/j.apsusc.2018.03.116
Reference: APSUSC 38864

To appear in: Applied Surface Science

Received Date: 8 February 2018


Revised Date: 9 March 2018
Accepted Date: 14 March 2018

Please cite this article as: X. Liu, L. Jiang, X. Jiang, X. Tian, Y. Huang, P. Hou, S. Zhang, X. Xu, Design of superior
ethanol gas sensor based on indium oxide/molybdenum disulfide nanocomposite via hydrothermal route, Applied
Surface Science (2018), doi: https://doi.org/10.1016/j.apsusc.2018.03.116

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Design of superior ethanol gas sensor based on indium

oxide/molybdenum disulfide nanocomposite via hydrothermal route

Xiaojing Liu*, Li Jiang, Xiumei Jiang, Xueying Tian, Ying Huang, Peiyu Hou,
Shouwei Zhang, Xijin Xu*
School of Physics and Technology, University of Jinan, Jinan 250022, Shandong Province, PR

China

* Corresponding author. Tel.: +86-531-8276-5480; fax: +86-531-8276-5480.

E-mail address: lxj@mail.sdu.edu.cn (X.J. Liu), sps_xuxj@ujn.edu.cn (X.J. Xu).

Abstract

This paper demonstrates an ethanol gas sensor based on indium oxide/molybdenum disulfide

(In2O3/MoS2) nanocomposite via hydrothermal route. The microstructure and micromorphology of

In2O3/MoS2 nanocomposite were fully characterized by various analytical techniques. The

gas-sensing properties of the In2O3/MoS2 composite were investigated upon exposure to different

concentrations of ethanol gas from 1 ppm to 50 ppm at the optimum temperature, and compared

with the pristine In2O3 sensors. Owing to the supporting substrate of specific two-dimensional

MoS2 nanosheets, the sensor based on In2O3/MoS2 composite exhibit superior gas sensing

performance towards ethanol, which outstripped that of pure In2O3 sensor and have potential

applications in the detection of ethanol vapors.

Keywords: Gas sensors; Ultra-sensitive ethanol sensing; Hydrothermal method

1. Introduction
Molybdenum disulfide (MoS2), as a graphene-liked 2D layered semiconductor, is
considered to be a promising candidate due to its extremely large surface-to-volume
ratio, and exceptional electrical properties. Compared with graphene which band gap
is 0, MoS2 layered structure with band-gap varies from 1.2 eV (bulk MoS2) for
indirect-gap to 1.8 eV (monolayer) for direct-band gap, playing a critical role in
improving its sensing performances [1–3]. Therefore, MoS2 has attracted tremendous
interest in various fields, especially gas sensors to detect ethanol, NH3 and NO2 at
room temperature [4–7]. However, intrinsic MoS2 nanosheets gradually degrade upon
exposure to the ambient conditions due to atmospheric oxidation and surface
contamination, with degradation of their electrical properties and sensing abilities [8].
Therefore, the pristine MoS2 nanosheets need to be operated under inert atmosphere
(e.g., N2) for obtaining satisfactory performances, which considerably limits their
application in commercial practice. To solve this problem, the modification of MoS2
with metal oxides becomes an effective route to fabricate high performance MoS2
sensors operated under air environment [9-11].
Indium oxide (In2O3) is an environmentally friendly n-type semiconductor (direct
band gap Eg=3.7 eV and indirect band gap Eg=2.5 eV), and has been widely applied in
various fields such as solar cell, supercapacitor field effects transistors, transparent
thin film transistors, photo-catalyst, flat panel display, light-emitting diodes, and
biological and chemical gas sensors [12-17]. In2O3 nanostructure offers a hopeful
platform for high performance gas sensing devices owing to its low cost, non-toxicity
and high sensitivity to various oxidizing and reducing gases [18-20]. In this paper, we
prepared a highly sensitive ethanol gas sensor based on In2O3/MoS2 (In2O3@MoS2)
nanocomposite. The as-prepared hybrid was fully characterized by many means and
confirmed its successful preparation, morphology, microstructure and composition.
The ethanol-sensing properties of the sensor were investigated, and high response,
swift response-recovery time, good repeatability and selectivity toward ethanol gas
were achieved. At last, the underlying sensing mechanism for In2O3 @MoS2 composite
was also discussed.

2. Experimental
2.1. Materials fabrication
In2O3 and MoS2 used in this work were prepared by facile hydrothermal route.
Fig.1(a) shows hydrothermal synthesis of MoS2 nanosheets. 2 mmol Sodium
molybdate (Na2MoO4·2H2O) and 9 mmol thiourea (CH4N2S) were dissolved into 70
mL deionized (DI) water, and stirred for 50 min. After that, 2 mmol citric acid
(C6H8O7·H2O) was added to the above solution, and followed with stirring for 30 min.
The obtained solution was transferred to a 100 mL stainless-steel autoclave and
heated oven at 200°C for 21 h. Next, the MoS2 suspension was washed with ethanol
and DI water several times to remove redundant ions, followed by ultrasonicating for
1 h and centrifugating at 3000 rpm for 30 min, respectively. After dried at 60°C for 12
h in air, the black MoS2 precipitate were obtained. Fig.1(b) shows the hydrothermal
synthesis of In2O3@MoS2 sample. 0.5 mmol InCl3·4H2O, and 1.5 mmol NaOH were
added into 50 mL DI water and stirred for 30 min, and then the resulting suspension
was thermally treated at 160°C for 12 h to obtain In2O3 suspension. Secondly, 1 mmol
MoS2 were added into the above solution by ultrasonic treatment for 30 min and
stirring for 2 h, followed by hydrothermally heated in a stainless steel autoclave
at180°C for 18 h. The obtained solution was washed several times with DI water and
ethanol to remove the impurity ions. At last, the In2O3@MoS2 samples were obtained
after drying for 12 h at 60°C.
2.2. Instrument and analysis
The crystal structure of the as-prepared samples were investigated using X-ray
diffractometer (XRD, Bruker D8 Advance) with Cu Kα radiation (λ= 1.5418 Å). The
morphologies of the samples were examined by field-emission scanning electronic
microscope (FESEM, FEI Company, QUANTA FEG 250). The transmission electron
microscope (TEM, Tecnai G2 F20) was operated with an accelerating voltage of 200
kV to further observe the microstructures of composite. X-ray photoelectron
spectroscopy (XPS) analysis using a Thermo Fisher K-Alpha XPS spectrometer was
used to further confirm the surface element composition and chemical state. Gas
sensing properties were investigated by the WS-60A gas sensing measurement system
(WeiSheng Electronics Science and Technology Co., Ltd., Henan Province, China).
After aging for 24 h under laboratory conditions (20%±10% RH, 25±1°C), the sensors
were placed into the transparent testing chamber (V=20 L) in the measuring system.
3. Results and discussion
3.1. Characterization results
In Fig.2, the purity and crystalline phase of hydrothermally synthesized MoS2
nanosheets, In2O3 nanoparticles and In2O3 @MoS2 nanosheets are analyzed by a
powder X-ray diffractogram, which are scanned in the range of 6~60°. As for the pure
MoS2 sample, the detected peaks at 2θ=14.4°, 32.7, 35.9°, 39.5°, 49.8° and 58.3° can
be easily indexed to the (002), (100), (102), (103), (105) and (110) planes in the
hexagonal phase of MoS2 (JCPDS 37-1492) [21]. No other peaks are found in the
MoS2 samples, illustrating the MoS2 is well crystallized. The XRD pattern of In2O3
sample presents major peaks at 2θ=21.5°, 30.6, 35.5°, 37.7°, 41.8°, 45.7° and 51.0° ,
which are attributed to the (211), (222), (400), (411), (332), (431) and (440) planes,
which are consistent with the values in the standard card (JCPDS 71-2194) [22].
Moreover, the average crystallite size of In2O3 calculated by Scherrer formula is about
8.34 nm. The XRD patterns for the In2O3/MoS2 samples demonstrate the distinct
peaks corresponding to the presence of individual In2O3 nanoparticles and MoS2
nanosheets, suggesting the successful preparation of In2O3/MoS2 nanocomposites.
The morphologies and nanostructure of the as-prepared samples were characterized
by scanning electron microscopy (SEM) and field emission scanning electron
microscopy (FESEM) observations. As depicted in Fig.3(a-b), it can be seen that the
large amount of uniform MoS2 is hierarchical flower-like nanosphere with a diameter
of about 400-600 nm. Furthermore, the 3D flowerlike architecture is consisted of
many 2D nanosheets, which are tightly aggregated. Fig.3(c) and 3(d) illustrate In2O3
nanoparticles with diameters less than 10 nm tightly dispersed on the edges and inner
space of the MoS2 nanoflower, which leads to the increased specific surface area and
more cusps on the surface of the In2O3@MoS2 composites. Further insight about the
detailed structural information of In2O3@MoS2 nanosheets products was elucidated
by TEM and HRTEM. Fig.3(e) depicts a typical image of MoS2 nanosheets inlayed
with dispersive In2O3 nanoparticles, which further supports that the sample is
constructed by transparent and thin 2D nanosheets. Under a higher magnification
(Fig.3f), it is obviously that many stripes are at the edges of nanosheets, which
indicates that the sample has layered structure. The distance value of lattice fringe
spacing is around 0.62 nm, matching with the spacing of the (002) crystal plane of the
hexagonal MoS2 crystalline structure. Moreover, the lattices of In2O3 nanoparticles
can be also clearly observed with the interplanar distance of 0.41 nm, matching the
(211) plane of cubic In2O3 crystalline structure (JCPDS 71-2194).
X-ray photoelectron spectroscopy (XPS) is a very useful method to determine the
chemical compositions and their chemical states of material surfaces. Fig.4(a) shows
that the main constituent elements are In, O, Mo and S. The doublets of In 3d5/2 and In
3d3/2, located at ~444.08 eV and ~451.58 eV in Fig.4(b), corresponding to trivalent
indium and metal indium [23]. The peak centered at ~532.1 eV belongs to O1s and
from XPS-peak-differentiating analysis, three separate Gaussian function peaks which
corresponding to three kinds of oxygen species on the surface of the material can be
observed (Fig.4c). The first one with the binding energy ~529.9 eV (OL) is lattice
(surface) oxygen O2−, the second one with the binding energy ~531.8 eV (OV) can be
assigned to the oxygen deficient regions O−, and the third one with the binding energy

~532.3 eV (OC) is chemisorbed oxygen species O2 . It has been reported that gas
sensing performance are greatly influenced by the deficient and chemisorbed oxygen
species in material [24]. The increase of OV component can provide more active sites
for the gas reaction and adsorption on the surface of the sensing materials. The
increase of OC component signified that more surface chemisorbed oxygen species
can participate in the oxidation-reduction reaction on the surface of the materials and
thus resulted in a larger change in the conductivity of sensing materials. These results
can be proved to be responsible for the enhancing gas sensing of the gas sensor (see
section 3.2). In Fig.4(d), the two peaks at 229.8 and 232.48 eV can be observed,
corresponding to the doublet Mo 3d5/2 and Mo 3d3/2, respectively, which is attributed
to the Mo4+ of MoS2 [25]. In Fig.4(e), the two peaks at 161.84 eV and 162.85 eV can
be observed, corresponding to the S 2p3/2 and the S 2p1/2, respectively, which is
assigned to the S2- of MoS2.
3.2. Ethanol gas-sensing properties
The responses of gas sensors based on pure In2O3 and In2O3@MoS2 to 50 ppm
ethanol were measured at an operating temperature range of 200-340°C, as shown in
Fig.5(a). Both sensors exhibit peak shaped dependence on the operating temperature,
which may be caused by kinetics and mechanics of gas reaction between chemisorbed
oxygen and reducing gas to n-type semiconductor [26-27]. Obviously, the responses
of In2O3@MoS2 sensor reach the maximum values of R=35.8 at the optimum working
temperature (260°C). Such behavior can also be observed in the case of the sensor
based on pristine In2O3. Its maximum response (R =13.9) appears at 280°C, and the
value is much lower than those of In2O3@MoS2 sensor throughout the working
temperature range. The obtained response and recovery characteristic curves against
ethanol of different concentrations (1-50 ppm) for the two sensors at the optimum
operating temperature are shown in Fig.5(b). With the increasing of the ethanol
concentration, the responses of both sensors increase gradually. The responses of
pristine In2O3 sensor are 1.8, 3.3, 4.2, 5.7, 6.4, 10.4–1, 2, 3, 5, 10, 50 ppm ethanol,
while the responses of In2O3 @MoS2 sensor on are 3.8, 6.5, 9.2, 14.3, 23.6, 35.8–1, 2,
3, 5, 10, 50 ppm ethanol, respectively. In addition, as shown in the inset of Fig.5(b),
we can find that the response of the In2O3@MoS2 sensor doesn’t tend to saturate
gradually when the ethanol concentration is raised to 100 ppm, which indicates that
this gas sensor has a broad test range. The In2O3@MoS2 sensor shows approximately
2-3 times higher response than pristine In2O3 sensor. Fig.5(c) shows a typical
reproducible run of the In2O3@MoS2 sensor after five cycles to 50 ppm of ethanol at
260°C. In conclusion, both relatively high response and excellent reproducibility can
be achieved in detecting low-concentration ethanol using In2O3@MoS2 composites as
gas-sensing material.
Fig.6(a) and (b) compares the response and recovery times of the In2O3@MoS2
sensor and pristine In2O3 sensor to 50 ppm ethanol at 260°C, which is also one of the
important characteristic parameters for gas sensors. The response time is defined as
the time spans to reach a 90% variation in resistance upon exposure to gas, and the
recovery time is fixed as the time necessary for the sensor to recovery 90% of its
initial resistance. The response time of In2O3 @MoS2 sensor is 8 s, which is faster than
those (11 s) of pristine In2O3 sensor. But both sensors take a long time (20-24 s) to
recover its initial resistance state. As it is known, chemical sensing is hardly
reversible because the thermal energy is usually lower than the activation energy for
desorption [28-30], which leads to a long recovery time.
Fig.7(a) exhibits the responses of the two sensors to 50 ppm of various testing
gases at 260°C. The testing gases include ethanol, benzene, acetone, ammonia, and
water. Obviously, both sensors to ethanol show the highest response compared with
those of other gases. The detailed reason for the improved selectivity of the
In2O3@MoS2 sensor toward ethanol is still unclear for us. One possible reason is the
different reaction activity of target gases in terms of bond energy. For instance, the
bond energies of O-H (ethanol), C=C (benzene), and C=O (acetone), are 458.8, 610.3,
and 798.9 kJ/mol, respectively [31]. Owing to the low O-H bond energy, the high
reaction activity of ethanol molecules is beneficial to the improved selectivity.
Another possible reason is optimized loading quantity of In2O3@MoS2 nanoparticles.
The long-term stability of two sensors over 50 days were also measured as depicted in
Fig.7(b). The responses of the two sensors show almost constant value to 50 ppm of
ethanol at 260°C, demonstrating the high stability of these gas sensors.
3.3. Ethanol gas-sensing mechanism
As it is known, In2O3 is a kind of n-type metal oxide semiconductor, in which the
carriers are electrons. Thus, the resistance of sensing material is related to the
concentration of the electrons. The reactions between ethanol and In2O3 are listed as
following [32-34]:
O 2( gas )  O 2( ads )
O 2( ads )  e   O 2 ( ads )
O 2 ( ads )  e   2O  ( ads )
O  ( ads )  e   O 2 ( ads )
C2 H 5OH  6O 2 ( ads )  2CO 2 +3H 2O+12e -

The sensor property of the In2O3 @MoS2 composites towards ethanol may be
interpreted in the following way (Fig.8).With the modification of MoS2 nanosheets,
the In2O3@MoS2 sensor working at 260°C achieves higher response than that of pure
In2O3 sensor at 280°C. This indicates that the modification of MoS2 into In2O3 can
lower the working temperature and decrease the power consumption, which is likely
due to the large specific surface and good electricity of MoS2, as well as the
synergistic effect between In2O3 and MoS2. It is well known that MoS2 nanosheet can
provides a platform for attaching In2O3 nanoparticles, preventing their interparticle
aggregation and creating more active sites for the adsorption of ethanol molecules.
Such kind of the nanostructure facilitates molecular absorption, gas diffusion and
mass transport and thus enhancing the gas sensing performance. Besides, due to the
high charge carrier mobility, MoS2 provide direct conduction paths for carriers to
transport from the junction to the external electrode, and thus the electrical signals
link closely and propagate rapidly. As a result, the gas-sensing performance can thus
be improved eventually in the composites of dispersing In2O3 nanoparticles on the
surfaces of MoS2 nanosheets.

4. Conclusions
In conclusion, we prepared a highly sensitive ethanol gas sensor based on
In2O3@MoS2 nanocomposite via surfactant assisted hydrothermal method. The
as-prepared samples were characterized by XRD, SEM, TEM and XPS, confirming its
successfully preparation and rationality. A series of experiments of the presented
sensor were performed upon exposure to various ethanol gas concentrations. Our
experimental findings would make a meaningful contribution towards the fabrication
of In2O3@MoS2 based gas sensor to recognize ethanol with excellent gas sensing
performance such as high response, excellent selectivity with good long-term stability
and quick response/recovery.

Acknowledgements
This work was supported by the National Natural Science Foundation of China
(Grant Nos. 61575081, 61205175, 51672109), the Encouragement Foundation for
Excellent Middle-aged and Young Scientist of Shandong Province, China (Grant No.
BS2011CL008) and the Doctoral Foundation of University of Jinan, China (Grant No.
XBS0920).
References

[1] Y.-J. Hsiao, T.-H. Fang, L.-W. Ji, B.-Y. Yang, Red-shift effect and sensitive responsivity of

MoS2/ZnO flexible photodetectors, Nanoscale Res. Lett. 10 (2015) 1-5.

[2] Q.L. Feng, K.Y. Duan, X.L. Ye, D.B. Lu, Y.L. Du, C.M. Wang, A novel way for detection of

eugenol via poly (diallyldimethylammonium chloride) functionalized graphene-MoS2 nano-flower

fabricated electrochemical sensor, Sens. Actuators B 192 (2014) 1-8.

[3] G. Eda, H. Yamaguchi, D. Voiry, T. Fujita, M. Chen, M. Chhowalla, Photoluminescence from

chemically exfoliated MoS2, Nano Lett. 11 (2011) 5111-5116.

[4] D.Z. Zhang, Yan’e Sun, C.X. Jiang, Y. Zhang, Room temperature hydrogen gas sensor based

on palladium decorated tin oxide/molybdenum disulfide ternary hybrid via hydrothermal route,

Sensors and Actuators B 242 (2017) 15-24.

[5] M. Donarelli, S. Prezioso, F. Perrozzi, F. Bisti, M. Nardone, L. Giancaterini, C. Cantalini, L.

Ottaviano, Response to NO2 and other gases of resistive chemically exfoliated MoS2-based gas

sensors, Sens. Actuators B 207 (2015) 602-613.

[6] S.J. Ray, First-principles study of MoS2, phosphorene and graphene based single electron

transistor for gas sensing applications, Sens. Actuators B 222 (2016) 492-498.

[7] Y.J. Liu, L.Z. Hao, W. Gao, Z.P. Wu, Y.L. Lin, G.X. Li, W.Y. Guo, L.Q. Yu, H.Z. Zeng, J. Zhu,

W.L. Zhang, Hydrogen gas sensing properties of MoS2/Si heterojunction, Sens. Actuators B 211

(2015) 537-543.

[8] W. Park, J. Park, J. Jang, H. Lee, H. Jeong, K. Cho, S. Hong, T. Lee, Oxygen environmental

and passivation effects on molybdenum disulfide field effect transistors, Nanotechnology 24 (2013)

095202.

[9] S. Cui, Z. Wen, X. Huang, J. Chang, J. Chen, Stabilizing MoS 2 nanosheets through SnO2

nanocrystal decoration for high-performance gas sensing in air, Small 11 (2015) 2305-2313.

[10] P. Zhao, Y. Tang, J. Mao, Y. Chen, H. Song, J. Wang, Y. Song, Y. Liang, X. Zhang,

One-dimensional MoS2-decorated TiO2 nanotube gas sensors for efficient alcohol sensing, J.

Alloys Compd. 674 (2016) 252-258.

[11] H. Yan, P. Song, S. Zhang, Z. Yang, Q. Wang, Facile synthesis, characterization and gas

sensing performance of ZnO nanoparticles-coated MoS2nanosheets,J. Alloys Compd. 662 (2016)


118-125.

[12] N. G. Pramod and S. N. Pandey, Influence of Sb doping on the structural, optical, electrical

and acetone sensing properties of In2O3 thin films, Ceram. Int., 2014, 40, 3461-3468.

[13] G. Shen, J. Xu, X. Wang, H. Huang and D. Chen, Growth of Directly Transferable In2O3

Nanowire Mats for Transparent Thin-film Transistor Applications, Adv.Mater., 2011, 23, 771-775.

[14] K.R.R. Gil, Y. Sun, E. R. Garcia and D. Raftery, Characterization of Photoactive Centers in

N-Doped In2O3 Visible Photocatalysts for Water Oxidation, J.Phys.Chem.C, 2009, 113,

12558-12570.

[15] C.Y. Huang, G.C. Lin, Y.J. Wu, T.Y. Lin, Y.J. Yang and Y.F. Chen, Efficient Light Harvesting

by Well-Aligned In2O3 Nanopushpins as Antireflection Layer on Si Solar Cells, J. Phys. Chem. C,

2011, 115, 13083-13087.

[16] L.C. Chen, C.H. Tien and W.C. Liao, A phosphor-free white light-emitting diode using In2O3:

Tb transparent conductive light converter, J. Phys. D: Appl. Phys., 2011, 44, 165101.

[17] P.C. Chen, G. Shen, S. Sukcharoenchoke and C. Zhou, Flexible and transparent

supercapacitor based on nanowire/carbon nanotube heterogeneous films, Appl. Phys. Lett., 2009,

94, 043113.

[18] X. S. Niu, H. X. Zhong, X. J. Wang and K. Jiang, Sensing properties of rare earth oxide

doped In2O3 by a sol–gel method, Sens. Actuators, B, 2006, 115, 434-438.

[19] X.J. Liu, L. Jiang, X.M. Jiang, X.Y. Tian, X. Sun, Y.L. Wang, W.D. He, P.Y. Hou, X.L. Deng,

X.J. Xu, Synthesis of Ce-doped In2O3 nanostructure for gas sensor applications, Applied Surface

Science 428 (2018) 478-484.

[20] M. Curreli, C. Li, Y. Sun, B. Lei, M. A. Gundersen, M. E. Thompson and C. Zhou, Selective

Functionalization of In2O3 Nanowire Mat Devices for Biosensing Applications, J. Am. Chem. Soc.,

2005, 127, 6922-6923.

[21] X.H. Zhang, X.H. Huang, M.Q. Xue, X. Ye, W.N. Lei, H. Tang, C.S. Li, Hydrothermal

synthesis and characterization of 3D flower-like MoS2 microspheres, Mater. Lett. 148 (2015)

67-70.

[22] X.J. Liu, L. Jiang, X.M. Jiang, X.Y. Tian, X. Sun, Y.L. Wang, W.D. He, P.Y. Hou, X.L.

Deng, X.J. Xu, Synthesis of Ce-doped In2O3 nanostructure for gas sensor applications, Applied

Surface Science 428 (2018) 478-484.


[23] H. Yang, S. Wang, Y. Yang, Zn-doped In2O3 nanostructures: preparation: structure and

gas-sensing properties, CrystEngComm 14 (2012) 1135-1142.

[24] L. Gao, F. Ren, Z. Cheng, Y. Zhang, Q. Xiang, J. Xu, Porous corundum-type In2O3

nanoflowers: controllable synthesis, enhanced ethanol-sensing properties and response mechanism,

CrystEngComm 17 (2015) 3268.

[25] X.Q. Qiao, F.C. Hu, D.F. Hou, D.S. Li, PEG assisted hydrothermal synthesis of hierarchical

MoS2 microspheres with excellent adsorption behavior, Mater.Lett. 169 (2016) 241-245.

[26] M. Arienzo, L. Armelao, C. M. Mari, S. Polizzi, R. Ruffo, R. Scotti and F. Morazzoni, J. Am.

Chem. Soc., 2011, 133, 5296–5304.

[27] L. X. Zhang, J. H. Zhao, H. Q. Lu, L. Li, J. F. Zheng, H. Li and Z. P. Zhu, Sens. Actuators, B,

2012, 161, 209-215.

[28] Z.Y. Fan, J.G. Lu, Gate-refreshable nanowire chemical sensors, Appl. Phys. Lett. 86 (2015)

123510.

[29] Q.Q. Zhao, D.X. Ju, X.F. Song, X.L. Deng, M. Ding, X.J. Xu, H.B. Zeng, Polyhedral

Zn2SnO4: Synthesis, Enhanced Gas Sensing and Photocatalytic Performance, Sensors and

Actuators B-Chemical, 229 (2016) 627-634.

[30] Q.Q. Zhao, X.L. Deng, M. Ding, L. Gan, T.Y. Zhai, X.J. Xu, One-pot synthesis of Zn-doped

SnO2 nanosheet-based hierarchical architectures as glycol gas sensor and photocatalyst,

CrystEngComm 17 (2015) 4394-4401.

[31] Q. Xu, D.X. Ju, Z.C. Zhang, S. Yuan, J. Zhang, H.Y. Xu, B.Q. Cao, Near room-temperature

triethylamine sensor constructed with CuO/ZnO P-N heterostructural nanorods directly on flat

electrode, Sens. Actuators B 225 (2016) 16–23.

[32] L. Wang, Z. Lou, J. Deng, R. Zhang, T. Zhang, Ethanol gas detection using a yolk-shell

(core-shell) α-Fe2O3 nanospheres as sensing material, ACS Appl.Mater. Interfaces 23 (2015)

13098-13104.

[33] S. Yang, Y.L. Liu, W. Chen, W. Jin, J. Zhou, H. Zhang, G.S. Zakharova, High sensitivity and

good selectivity of ultralong MoO3 nanobelts for trimethylamine gas, Sens. Actuators B 226 (2016)

478-485.

[34] S.H. Park, S.Y. Kim, G.J. Sun, C.M. Lee, Synthesis structure, and ethanol gas sensing

properties of In2O3 nanorods decorated with Bi2O3 nanoparticles, ACS Appl. Mater. Interfaces 7
(2015) 8138-8146.
Figures captions:

Fig.1. Hydrothermal route for the fabrication of In2O3/MoS2 suspension.


Fig.2. XRD patterns of pure MoS2, In2O3, and In2O3@MoS2 composites.
Fig.3. (a)-(b) SEM images of pure MoS2 samples, (c)-(d) SEM images and (e and f)
TEM images of hierarchical In2O3@MoS2 composites.
Fig.4. XPS spectra of In2O3@MoS2 sample: (a) survey spectrum, (b) In 3d core level
XPS spectrum, (c) O 1s core level XPS spectrum, (d) Mo 3d core level XPS spectrum,
(e) S 2p core level XPS spectrum.
Fig.5. (a) Sensing properties of pure In2O3 and In2O3@MoS2 sensor towards 50 ppm
ethanol gas at different operating temperatures; (b) Corresponding relationship
between sensor response and ethanol concentration; (c) Repeatability test of the
In2O3@MoS2 sensor to 50 ppm of ethanol at 260°C.
Fig.6. Response and recovery time of the In2O3 and In2O3@MoS2 sensor.
Fig.7. (a) The selectivity comparison of two sensors for different target gases at
260°C; (b) Long-term stability of the two sensors to 50 ppm of ethanol at 260°C.
Fig.8. Schematic illustration of the gas sensing mechanism for In2O3@MoS2 sensor
exposed to ethanol.
Fig.1. Hydrothermal route for the fabrication of In2O3/MoS2 suspension.

Fig.2. XRD patterns of pure MoS2, In2O3, and In2O3@MoS2 composites.

14
Fig.3. (a)-(b) SEM images of pure MoS2 samples, (c)-(d) SEM images and (e and f) TEM images

of hierarchical In2O3@MoS2 composites.

15
Fig.4. XPS spectra of In2O3@MoS2 sample: (a) survey spectrum, (b) In 3d core level XPS

spectrum, (c) O 1s core level XPS spectrum, (d) Mo 3d core level XPS spectrum, (e) S 2p core

level XPS spectrum.

16
Fig.5. (a) Sensing properties of pure In2O3 and In2O3@MoS2 sensor towards 50 ppm ethanol gas

at different operating temperatures; (b) Corresponding relationship between sensor response and

ethanol concentration; (c) Repeatability test of the In2O3@MoS2 sensor to 50 ppm of ethanol at

260°C.

Fig.6. Response and recovery time of the In2O3 and In2O3@MoS2 sensor.

17
Fig.7. (a) The selectivity comparison of two sensors for different target gases at 260°C; (b)

Long-term stability of the two sensors to 50 ppm of ethanol at 260°C.

Fig.8. Schematic illustration of the gas sensing mechanism for In2O3@MoS2 sensor exposed to

ethanol.

18
 An ethanol gas sensor based on indium oxide/molybdenum disulfide (In2O3/MoS2)

nanocomposite has been prepared via hydrothermal route.

 The microstructure and micromorphology of In2 O3/MoS2 nanocomposite were

fully characterized by various analytical techniques.

 The sensor based on In2O3/MoS2 composite exhibit superior gas sensing

performance towards ethanol, which outstripped that of pristine In2O3 sensor.

 It is expected that hierarchical In2O3/MoS2 composite with excellent gas-sensing

performance have potential applications in ethanol detection.

19

You might also like