The Combination of Two-Dimensional Nanomaterials With Metal Oxide Nanoparticles For Gas Sensors: A Review
The Combination of Two-Dimensional Nanomaterials With Metal Oxide Nanoparticles For Gas Sensors: A Review
The Combination of Two-Dimensional Nanomaterials With Metal Oxide Nanoparticles For Gas Sensors: A Review
Review
The Combination of Two-Dimensional Nanomaterials with
Metal Oxide Nanoparticles for Gas Sensors: A Review
Tao Li 1 , Wen Yin 1 , Shouwu Gao 2 , Yaning Sun 1 , Peilong Xu 2 , Shaohua Wu 1 , Hao Kong 3 , Guozheng Yang 3
and Gang Wei 3, *
1 College of Textile & Clothing, Qingdao University, No. 308 Ningxia Road, Qingdao 266071, China;
qdlitao2008@126.com (T.L.); yiwin@126.com (W.Y.); qdfzxsyn@163.com (Y.S.); shaohua.wu@qdu.edu.cn (S.W.)
2 State Key Laboratory, Qingdao University, No. 308 Ningxia Road, Qingdao 266071, China;
qdgsw@126.com (S.G.); xpl@qdu.edu.cn (P.X.)
3 College of Chemistry and Chemical Engineering, Qingdao University, No. 308 Ningxia Road,
Qingdao 266071, China; konghao5@outlook.com (H.K.); yangguozheng123@outlook.com (G.Y.)
* Correspondence: weigroup@qdu.edu.cn; Tel.: +86-1506-6242-101
Abstract: Metal oxide nanoparticles have been widely utilized for the fabrication of functional gas
sensors to determine various flammable, explosive, toxic, and harmful gases due to their advantages
of low cost, fast response, and high sensitivity. However, metal oxide-based gas sensors reveal the
shortcomings of high operating temperature, high power requirement, and low selectivity, which
limited their rapid development in the fabrication of high-performance gas sensors. The combination
of metal oxides with two-dimensional (2D) nanomaterials to construct a heterostructure can hybridize
the advantages of each other and overcome their respective shortcomings, thereby improving the
sensing performance of the fabricated gas sensors. In this review, we present recent advances in
the fabrication of metal oxide-, 2D nanomaterials-, as well as 2D material/metal oxide composite-
Citation: Li, T.; Yin, W.; Gao, S.; Sun, based gas sensors with highly sensitive and selective functions. To achieve this aim, we firstly
Y.; Xu, P.; Wu, S.; Kong, H.; Yang, G.; introduce the working principles of various gas sensors, and then discuss the factors that could affect
Wei, G. The Combination of the sensitivity of gas sensors. After that, a lot of cases on the fabrication of gas sensors by using
Two-Dimensional Nanomaterials metal oxides, 2D materials, and 2D material/metal oxide composites are demonstrated. Finally, we
with Metal Oxide Nanoparticles for summarize the current development and discuss potential research directions in this promising topic.
Gas Sensors: A Review. Nanomaterials We believe in this work is helpful for the readers in multidiscipline research fields like materials
2022, 12, 982. https://doi.org/ science, nanotechnology, chemical engineering, environmental science, and other related aspects.
10.3390/nano12060982
Academic Editor: Lyubov Keywords: two-dimensional materials; metal oxide; nanoparticles; composite materials; gas sensors
G. Bulusheva
to the different gas detection methods, semiconductor gas sensors can be divided into two
types: resistive and non-resistive, in which the resistive semiconductor gas sensor detects
gas concentration according to the change of the resistance value of the semiconductor
when it comes into contact with the gas [13]. Currently, gas-sensitive materials such as
semiconductor metal oxides, conductive polymers, and carbon materials have been used
for the fabrication of resistive semiconductor-type gas sensors [14–17].
Among these sensing materials for the fabrication of semiconductor gas sensors,
metal oxides, including zinc oxide (ZnO), indium oxide (In2 O3 ), tin oxide (SnO2 ), and
tungsten oxide (WO3 ) have been proved to be the best candidates for the fabrication for
making resistive gas sensors due to their advantages of simple fabrication, low cost, easy
portability, and high sensitivity [18–21]. However, the high operating temperature, high
power, and low selectivity limited their rapid development [22,23]. Two-dimensional (2D)
nanomaterials, including graphene, transition metal chalcogenides, layered metal oxides,
black phosphorus, and others, have shown great potential in gas sensors due to their
unique single-atom-layer structure, high specific surface area, and many surface-active
sites [24,25].
2D material-based gas sensors have the advantages of high sensitivity, fast response
speed, low energy consumption, and room temperature operation [26–30]. However,
since 2D nanomaterials tend to form a dense stack structure during the formation of the
conductive network, it is not conducive to the full contact between the flakes inside the
conductive network and the gas molecules, making the sensitivity and response recovery
speed relatively low at room temperature. Combining metal oxides with 2D nanomaterials
to construct a heterostructure can combine the advantages of each other and overcome their
respective shortcomings, thereby improving the sensing performance of the fabricated gas
sensors. The combination of the metal oxide and 2D materials can drive transformations
in the design and performance of 2D nanoelectronics devices, such as the graphene/2D
indium oxide/SiC heterostructure [31,32]. Currently, the combination of metal oxides with
graphene, transition metal chalcogenide, and other 2D materials to form heterojunction
nanostructures for gas sensors have been studied widely, which have exhibited significantly
enhanced sensing performance at room temperature [33,34].
In this review, we present the advances in the fabrication and sensing mechanisms of
2D material- and metal oxide nanoparticle-based gas sensors. For this aim, we first intro-
duce the detection mechanism of the resistive semiconductor gas sensors and the factors
that can affect the sensitivity of the gas sensors. Then, various types of gas sensors based
on metal oxides, 2D materials, and 2D materials/metal oxides composites are introduced.
Special emphasis is placed on the recent progress of the combination of metal oxides and
2D nanomaterials for gas sensors. We believe that this review will be helpful for readers
to understand the synthesis of functional 2D material-based composites and promote the
fabrication of 2D material-based sensors for the high-performance determination of gases.
(d) NO (e) NO2 (f) CO
Figure 2. Charge density difference plots for (a) O 2, (b) H2O, (c) NH3, (d) NO, (e) NO2, and (f) CO
Figure 2. Charge density difference plots for (a) O2 , (b) H2 O, (c) NH3 , (d) NO, (e) NO2 , and (f) CO
interacting with monolayer MoS 2. Reprinted with permission from Ref. [39]. Copyright 2013
interacting with monolayer MoS2 . Reprinted with permission from Ref. [39]. Copyright 2013 Springer.
Springer.
2.3. Gas Sensing Mechanism of 2D Material/Metal Oxide Composites
2.3. Gas Sensing Mechanism of 2D Material/Metal Oxide Composites
When one material is composited with another, the bonding between different ma-
When one material is composited with another, the bonding between different mate‐
terials forms p-n, n-n, p-p, and Schottky heterojunctions. Among them, the formation of
rials forms p‐n, n‐n, p‐p, and Schottky heterojunctions. Among them, the formation of p‐
p-n heterojunction is beneficial for adjusting the thickness of the electron depletion layer,
thereby further improving the sensing performance of the fabricated gas sensors. For
n heterojunction is beneficial for adjusting the thickness of the electron depletion layer,
example, when SnO2 is combined with graphene oxide (GO), the p-n heterojunction can
thereby further improving the sensing performance of the fabricated gas sensors. For ex‐
be formed to form a new energy-level structure, as shown in Figure 3 [40]. The electron
ample, when SnO 2 is combined with graphene oxide (GO), the p‐n heterojunction can be
dissipation layer expands at the interface of SnO2 and GO, resulting in increased resistance.
formed to form a new energy‐level structure, as shown in Figure 3 [40]. The electron dis‐
When formaldehyde is introduced, the trapped electrons are released back into the conduc-
sipation layer expands at the interface of SnO 2 and GO, resulting in increased resistance.
tion band, resulting in a reduction in the width of the dissipation layer, which reduces the
When formaldehyde is introduced, the trapped electrons are released back into the con‐
resistance of the sample. The porous and ultrathin structure of the SnO2 /GO composite
duction band, resulting in a reduction in the width of the dissipation layer, which reduces
increases the specific surface area and active sites, facilitates the reaction with HCHO gas,
the resistance of the sample. The porous and ultrathin structure of the SnO
and contributes to the ultrahigh response for gas sensing. It should be noted 2/GO composite
that the
ultrathin nanosheet structure of GO shortens the transport path and greatly improves
increases the specific surface area and active sites, facilitates the reaction with HCHO gas, the
response of the gas sensor. Meanwhile, the abundant pores in SnO are favorable
and contributes to the ultrahigh response for gas sensing. It should be noted that the ul‐
2 for gas
diffusion and help to improve the response/recovery performance. In addition, GO can
trathin nanosheet structure of GO shortens the transport path and greatly improves the
act as a spacer, which reduced the agglomeration of SnO2 nanoparticles, and provided
response of the gas sensor. Meanwhile, the abundant pores in SnO 2 are favorable for gas
abundant adsorption sites for HCHO gas, thereby enhancing the gas sensing response.
diffusion and help to improve the response/recovery performance. In addition, GO can
act as a spacer, which reduced the agglomeration of SnO2 nanoparticles, and provided
abundant adsorption sites for HCHO gas, thereby enhancing the gas sensing response.
2022, 12, x FOR PEER 2022,
Nanomaterials REVIEW
12, 982 5 of 40 5 of 40
Figure 4.
Figure 4. Schematic
Schematicmodel
model of of
thethe
effect of the
effect of crystallite size on
the crystallite theon
size sensitivity of metal-oxide
the sensitivity gas
of metal-oxide gas
sensors: (a) D >> 2L, (b) D ≥ 2L, and (c) D < 2L. Reprinted with permission from Ref. [42]. Copyright
sensors: (a) D >> 2L, (b) D ≥ 2L, and (c) D < 2L. Reprinted with permission from Ref. [42]. Copyright
1991 Elsevier.
1991 Elsevier.
However, when the grain size is excessively reduced, the agglomeration between
However, when the grain size is excessively reduced, the agglomeration between
particles is serious. If the aggregates are relatively dense, only the particles on the surface
particles is serious.
of the aggregates If the
could aggregates
participate in theare
gasrelatively dense, only
sensing reaction, and thetheinternal
particles on the surface
materials
of the aggregates could participate in the gas sensing reaction,
are wasted because they are not in contact with the gas, resulting in a decrease in theand the internal materials
are
utilization rate of the material. In addition, the agglomeration is not conducive to the dif- in the
wasted because they are not in contact with the gas, resulting in a decrease
utilization
fusion of gasrate of the
inside material.
the material andInwill
addition,
reduce the thegasagglomeration
sensing performanceis not [45,46].
conducive
In- to the
creasing the specific surface area not only facilitates the adsorption of
diffusion of gas inside the material and will reduce the gas sensing performance [45,46].oxygen molecules
in the air onthe
Increasing thespecific
surface surface
of the material,
area not but alsofacilitates
only increases more effective active
the adsorption sites and
of oxygen molecules
in the air on the surface of the material, but also increases more effective test
more gas transmission channels to facilitate the diffusion and absorption of the gas.sites and
active
Therefore,
more increasing the channels
gas transmission specific surface area ofthe
to facilitate gas-sensing
diffusionmaterials is an important
and absorption of the test gas.
way to modify the sensing properties of gas sensors. The regulation of both the morphol-
Therefore, increasing the specific surface area of gas-sensing materials is an important way
ogy (flower-like, sea urchin-like, etc.) and porous structure (macropores, mesopores, and
to modify the sensing properties of gas sensors. The regulation of both the morphology
micropores) of materials is an effective way to improve the specific surface area of mate-
(flower-like, sea urchin-like, etc.) and porous structure (macropores, mesopores, and
rials. Nanomaterials with porous structures can increase the effective surface area and
micropores)
active sites ofofthematerials is antoeffective
material due waypore
their special to improve
structure,the specific
so that surfacehas
the material area of materials.
better
Nanomaterials
permeability, making with gasporous structures
molecules can
easier to increase
diffuse the interior
into the effective surface
of the areaand
material, and active
sites of thethe
increasing material due to their
contact between specialand
the material pore
the structure, so that the
gas. It can accelerate the material
diffusion has
of better
permeability,
gas, improve the making
responsegasand
molecules
recovery easier
speed oftothe
diffuse into the
gas sensor, andinterior of theits
thus improve material,
gas and
increasing the contact between the material and the gas. It can accelerate the diffusion of
sensing performance.
For example,
gas, improve the Boudiba
responseetandal. synthesized
recovery speedWO3 materials
of the gas with different
sensor, andmorphologies
thus improve its gas
by directperformance.
sensing precipitation, ion exchange, and hydrothermal methods, and further used the
For example, Boudiba et al. synthesized WO3 materials with different morphologies
by direct precipitation, ion exchange, and hydrothermal methods, and further used the
as-prepared materials to fabricate gas sensors. Their results indicated that the greater the
porosity, the higher the sensitivity to SO2 gas [47]. In another case, Jia et al. prepared
WO3 semiconductor materials with different morphologies by hydrothermal method as
Nanomaterials 2022, 12, 982 7 of 40
sensitive materials. Under the same test conditions, they found that the sensitivity of
WO3 nanorods towards acetone was 19.52, while the sensitivity of WO3 nanospheres
towards acetone was 25.71. In addition, WO3 nanoshpheres exhibited better selectivity
than nanorods [48]. Lü et al. [49] successfully prepared porous materials with extremely
high specific surface area (120.9 m2 ·g−1 ) by simple chemical transformation of Co-based
metal-organic frameworks (Co-MOFs) template and controlling the appropriate calcination
temperature (300 ◦ C). The prepared Co3 O4 concave nanocubes were systematically tested
for their gas-sensing properties to volatile organic compounds (VOCs), including ethanol,
acetone, toluene, and benzene. to the fabricated sensors exhibited excellent performance
in gas sensing, such as high sensitivity, low detection limit (10 ppm), fast response and
recovery (<10 s), and high selectivity for ethanol. Wang et al. synthesized concave Cu2 O
octahedral nanoparticles with a diameter of about 400 nm and performed gas-sensing tests
for benzene (C6 H6 ) and NO2 [50]. It was found that the concave Cu2 O octahedral nanopar-
ticles exhibited better gas sensing properties than Cu2 O nanorods. Unlike conventional
octahedrons, Cu2 O octahedral nanoparticles have a structure similar to icosahedral with
sharp boundaries. Therefore, compared with nanorods, the synthesized Cu2 O octahedral
nanoparticles have a larger specific surface area, which can provide more reactive sites and
thus exhibit better gas sensing properties.
of single gas molecules on the graphene surface [61]. This study opens the door to research
on 2D graphene-based gas sensors.
Single-layer graphene nanosheets, RGO, chemically modified graphene, and GO
have been proven to be good gas sensing materials [62–64]. Since the main advantage of
graphene nanostructures is the low temperature response, this sensor can greatly reduce
the energy consumption of the sensing device. Various graphene-based gas sensors have
been used to detect various harmful gases such as NO2 , NH3 , CO2 , SO2 , and H2 S. For
example, Ricciardella et al. developed a graphene film-based room temperature gas sensor
with a sensitivity of up to 50 ppb (parts-per-billion) to NO2 [65]. Various methods, such as
mechanical exfoliation, chemical vapor deposition (CVD), and epitaxy, have been used to
prepare graphene for gas sensing applications. For example, Balandin et al. [66] prepared
monolayer graphene using a mechanical exfoliation method and reported a monolayer
intrinsic graphene transistor, which can utilize low-frequency noise in combination with
other sensing parameters to realize selective gas sensing of monolithic graphene transistors.
Choi et al. [67] prepared graphene by CVD and transferred it onto flexible substrates, and
demonstrated a gas sensor using graphene as a sensing material on a transparent (Tr > 90%)
flexible substrate. Nomani et al. demonstrated that epitaxial growth of graphene on Si and
C surfaces of semi-insulating 6H-SiC substrates can provide very high NO2 detection sensi-
tivity and selectivity, as well as fast response times [68]. Yang et al. directly grew multilayer
graphene on various substrates through the thermal annealing process of catalytic metal
encapsulation, and tested it as a gas sensor for NO2 and NH3 gas molecules to detect its
response sensitivity to NO2 and NH3 [69]. The schematic diagram of the graphene sensor
is shown in Figure 5a. The NO2 molecules are electron acceptors (p-type dopant), which ex-
tract electrons from graphene, while the NH3 molecules are electron donors (n-type dopant),
which donate electrons to graphene. Therefore, when NO2 molecules are adsorbed, the
Nanomaterials 2022, 12, x FOR PEER REVIEW
conductivity of graphene is enhanced, while when NH3 molecules are adsorbed10 onofthe
40
graphene surface, the conductivity decreases due to the compensation effect (Figure 5b).
Figure 5. Schematic (a) and time-resolved sensitivity (b) of the the graphene
graphene sensor
sensor toward
toward NH NH33 and
gas molecules.
NO22 gas molecules. (c)
(c) Current
Current vs. time
time curves
curves for
for 5–1100
5–1100 ppm
ppm of SO SO22 for
forthe
theoriginal
original GO
GO and
and edge-
edge-
tailored GO nanosheets, and (d) the corresponding
corresponding sensitivities
sensitivities of
of the
the sensors
sensors to SO22 gas. Reprinted
to SO
with permission
with permission from
from Ref.
Ref. [69].
[69]. Copyright
Copyright 2016
2016 ACS.
ACS.
GO is suitable for gas sensors due to its multiple properties, such as easy processing,
high solubility in various solvents, and containing oxygen functional groups or defects.
Since the defects or functional groups in GO can act as reaction sites for gas adsorption,
making the gas easily adsorbed on the surface of GO and improving the selectivity and
sensitivity of the GO-based sensor, the response of the GO-based sensor can be tuned by
Nanomaterials 2022, 12, 982 10 of 40
GO is suitable for gas sensors due to its multiple properties, such as easy processing,
high solubility in various solvents, and containing oxygen functional groups or defects.
Since the defects or functional groups in GO can act as reaction sites for gas adsorption,
making the gas easily adsorbed on the surface of GO and improving the selectivity and
sensitivity of the GO-based sensor, the response of the GO-based sensor can be tuned by
functionalization. Shen et al. [70] prepared edge-trimmed GO nanosheets by periodically
acid-treating GO, and then fabricated field effect transistors (FETs) for gas sensing testing
of SO2 at room temperature (Figure 5c,d). Compared with pristine GO nanosheets, edge-
clipped GO nanosheets were found to have a significant response enhancement effect to
SO2 gas, and the detection concentration range was 5–1100 ppm. Meanwhile, the edge-
trimmed GO device also exhibited a fast response time, which was mainly attributed
to the hygroscopic properties of the GO nanosheets, which can trap water molecules
and react with SO2 to generate sulfuric acid to facilitate their fast protonation process.
By utilizing different reducing agents to remove oxygen from GO and recover aromatic
double-bonded carbons, the selectivity of RGO-based gas sensors could be improved
significantly. For example, Guha et al. [71] developed a gas sensor for NaBH4 reduction
of GO on a ceramic substrate and reported its performance for detecting NH3 at room
temperature. The response to NH3 can be optimized by the reduction time of GO. Through
chemical modification, RGO can introduce some foreign groups or atoms to change its
surface properties, which can enhance its sensing performance. For example, the response
of RGO reduced with p-phenylenediamine (PPD) to dimethyl methylphosphonate (DMMP)
was 4.7 times higher than that of RGO reduced by ordinary methods [72]. The RGO-based
gas sensor reduced by ascorbic acid has high selectivity for corrosive NO2 and Cl2 , and the
detection limit can reach 100 and 500 ppb, respectively [73].
Figure 6. Sensing behavior of atomically thin-layered MoS2 transistors. (a) Schematic of the MoS2
transistor-based NO2 gas-sensing device. (b) Optical photograph of the MoS2 sensing device mounted
on the chip. Comparative two- and five-layer MoS2 cyclic sensing performances with NH3 . (c) and
NO2 (for 100, 200, 500, 1000 ppm) (d). Reprinted with permission from Ref. [78]. Copyright 2013 ACS.
At present, 2D Sn-based sulfide materials (SnS and SnS2 ) are also used in the field of
gas sensors due to their unique performance advantages. For example, Wang et al. [79]
successfully synthesized free-standing large-scale ultrathin SnS crystalline materials by
utilizing the 2D directional attachment growth of colloidal quantum dots in a high-pressure
solvothermal reaction. The SnS ultrathin crystals were rectangular with uniform shape,
the lateral dimension was between 20 and 30 µm, and the thickness was less than 10 nm
(Figure 7a,b). The obtained material was used to fabricate a gas sensor, which exhibited
excellent sensitivity and selectivity for NO2 at room temperature with a detection limit
of 100 ppb (Figure 7c–e). Xiong et al. [80] synthesized 3D flower-like SnS2 nanomaterials
assembled from nanosheets and fabricated them into gas sensors by a simple solvothermal
method, As shown in Figure 7f–h. When 100 ppm NH3 was detected at 200 ◦ C, the
response value was 7.4, the response time was 40.6 s, and the recovery time was 624 s.
The prepared nanoflowers have good selectivity to NH3 with a detection limit of 0.5 ppm.
This study attributes the excellent performance of the SnS2 sensor for NH3 to the unique
thin-layer flower-like nanostructure, which is beneficial to the carrier transfer process and
gas adsorption/desorption process [23].
materials assembled from nanosheets and fabricated them into gas sensors by a simple
solvothermal method, As shown in Figure 7f–h. When 100 ppm NH3 was detected at 200
°C, the response value was 7.4, the response time was 40.6 s, and the recovery time was
624 s. The prepared nanoflowers have good selectivity to NH3 with a detection limit of 0.5
Nanomaterials 2022, 12, 982
ppm. This study attributes the excellent performance of the SnS2 sensor for NH312toof the 40
unique thin-layer flower-like nanostructure, which is beneficial to the carrier transfer pro-
cess and gas adsorption/desorption process [23].
Figure 7. TEM (a) and HRTEM (b) images of 2D thin SnS crystals. Inset in (b) is the corresponding
low-magnification TEM image. (c) Schematic structure of SnS thin-crystal-based gas-sensor device.
The inset shows the optical image of the device. (d) Real-time voltage response after exposure of
the device to NO2 gas with increased concentration. The inset schematically illustrates the electron
transfer process from SnS to NO2 . (e) Selectivity of the sensor to a series of gases of 1 ppm. Inset shows
the Q values of the SnS thin crystal sensor for NO2 as a target gas. Reprinted with permission from
Ref. [79]. Copyright 2016 ACS. (f) Representative FESEM image of the flower-like SnS2 synthesized
by a facile solvothermal technique. (g) Sensor responses of the SnS2 based sensor upon exposure to
six kinds of gases at 200 ◦ C. (h) Typical response-recovery characteristic of the SnS2 based sensor
to different concentrations of NH3 gas at 200 ◦ C (Inset shows the corresponding response curve).
Reprinted with permission from Ref. [80]. Copyright 2018 Elsevier.
In order to further improve the gas sensing performance of TMD materials, people
have improved the gas-sensing performance through external energy strategies (ultraviolet-
assisted irradiation or applying a bias voltage, etc.) or composite strategies with other
materials. For example, Late et al. found that 5-layer MoS2 has a more sensitive response
to both NH3 and NO2 than 2-layer MoS2 with the assistance of a bias voltage (+15 V). The
photoconverted radiation of 4 mW/cm2 can increase the sensitivity of NO2 gas sensing,
while the light intensity of 15 mW/cm2 can reduce the recovery time [78]. Wu et al. [81]
prepared MoTe2 nanosheets by mechanical exfoliation, and the sensitivity to NO2 under
254 nm UV light was significantly improved by one order of magnitude compared with the
Nanomaterials 2022, 12, 982 13 of 40
dark condition, and the detection limit was significantly reduced to 252 ppt. Gu et al. [82]
synthesized 2D SnS2 nanosheets by a solvothermal method, and after irradiation with a
520 nm green LED lamp, realized NO2 detection at room temperature, with good repeata-
bility and selectivity for 8 ppm NO2 in a dry environment and the response value was 10.8.
Cheng et al. [83] combined the excellent sensing performance and gas adsorption capacity
of 2D SnS2 hexagonal nanosheets with the good electrical properties of graphene, and used
the excellent electrical conductivity of graphene to make up for the shortcoming of the poor
conductivity of SnS2 at room temperature and prepared a high-performance RGO/SnS2
heterojunction-based NO2 sensor. Compared with the single SnS2 gas sensor, the graphene-
doped sensor exhibited better selectivity to NO2 , while effectively reducing the optimal
operating temperature of the device, and its response to 5 ppm NO2 gas increased by nearly
one order of magnitude, and the response recovery time was reduced to less than a minute.
Compared with the single SnS2 gas sensor, the graphene-doped sensor exhibited good
selectivity to NO2 , while effectively reducing the optimal operating temperature of the
device, and its response to 5 ppm NO2 gas increased by nearly one order of magnitude,
and the response recovery time was shortened to less than one minute.
hance the gas-sensing properties of 2D metal oxides, ion doping or surface modification on
them is a valuable approach to enhance the response and recovery properties. For example,
Chen et al. [92] prepared 2D Cd-doped porous Co3 O4 nanosheets by microwave-assisted
solvothermal method and in situ annealing process, and investigated their sensing perfor-
mance for NO2 at room temperature. It was found that 5% Cd-doped Co3 O4 nanosheets
significantly improved the response to NO2 at room temperature (3.38), decreased the
recovery time (620 s), and lowered the detection limit to 154 ppb. The reason for the
performance improvement is that Cd doping mainly promotes the adsorption of NO2
through a series of factors such as enhancing the electronic conductivity, increasing the
concentration of oxygen vacancies, and forming Co2+ - O2− , thus promoting its excellent
room temperature sensing performance.
nanomaterials has high response sensitivity and fast response-recovery speed. According
to the semiconductor type, metal oxide semiconductors can be divided into n-type and
p-type. In n-type semiconductors, including SnO2 , ZnO, TiO2 , In2O3 , etc., the carriers are
mainly free electrons. However, in p-type semiconductors, such as CuO, NiO, Co3 O4 , etc.,
the carriers are mainly holes. When n-type semiconductors are exposed to reducing gases
(such as ethanol, NH3 , H2 , etc.), the resistance of the materials will decrease, while when
exposed to oxidizing gases (such as NO3 ), the resistance of the materials will increase. In
contrast to n-type semiconductors, the resistance of p-type semiconductors is higher when
exposed to reducing gas and decreases when exposed to oxidizing gas. At present, the
most studied metal oxide materials for gas sensing are SnO2 , ZnO, TiO2 , CuO, WO3 and so
on [57,104–107].
and H2 gas-sensing properties of undoped and Pd-doped SnO2 nanowires were studied.
It was found that with the increase of Pd doping concentration, the working temperature
decreased and the response of the sensor to H2 increased. Similarly, doping Au into SnO2
thin films can change the morphology of SnO2 thin films, reduce the grain size of SnO2
thin films, decrease the working temperature of the sensor, and improve the sensitivity and
selectivity of SnO2 to reducing gases such as CO [124]. Zhao et al. carried out Cu doping
on SnO2 nanowires. Compared with undoped SnO2 nanoscale arrays, the sensitivity and
selectivity of the sensor to SO2 in a dry environment were improved significantly [125].
In addition, the researchers synthesized composite nanomaterials containing two
different energy band structure materials to form heterostructures to improve the gas
sensing performance of SnO2 -based gas sensors. For example, Chen et al. prepared
Fe2 O3 @SnO2 composite nanorods with multi-stage structure by a two-step hydrothermal
method and found that the composite structure has good selectivity for ethanol [126].
Xue et al., using SnO2 nanorods synthesized by hydrothermal method as carriers, obtained
SnO2 composite nanorods loaded with CuO nanoparticles by ultrasonic and subsequent
calcination in Cu (NO3 )2 solution. The gas sensing properties of the materials for the
detection of H2 S were studied. It is found that the sensitivity of the sensor to 10 ppm H2 S
can reach 9.4 × 106 at 60 ◦ C [127]. The ultra-high sensitivity of the composite is attributed to
the p-n junction formed between CuO and SnO2 . In the air, the formation of heterojunction
increases the height of the energy barrier, hinders the flow of electrons, resulting in an
increase in the resistance of the material. When the material is in contact with H2 S and
reacts, it can form CuS, which is similar to metal conductivity, which greatly enhances the
electrical conductivity of the material. Fu et al. prepared NiO-modified SnO2 nanoparticles,
which increased the thickness of the electron depletion layer on the surface of SnO2 through
the formation of p-n heterojunction in air. While in the SO2 atmosphere, NiO reacted with
SO2 to form NiS, which promoted the release of electrons from the surface adsorbed O−
to SnO2 , thus enhancing the response of the device to SO2 gas and improving the gas
sensitivity of SnO2 materials to SO2 [128].
Figure 8. SEM images of (a) ZnO nanorods, (b) ZnO nanoflowers and (c) ZnO nanospheres.
(d) Dynamic response curves with time of three different ZnO nanostructures; (e) the response
curves with NO2 concentration of three different ZnO nanostructures; (f,g) the response and recovery
time of three different ZnO nanostructures; (h) the repeatability of three different ZnO nanostruc-
tures to 5 ppm NO2 ; (i) the selectivity of three different ZnO nanostructures to other harmful gases.
Reprinted with permission from Ref. [130]. Copyright 2021 Elsevier.
Similar to the SnO2 -based gas sensor, researchers changed the cell parameters of the
original ZnO by element doping, making it produce lattice deformation, cause the surface
defects of the gas sensing materials, and increase the surface active sites, so as to improve
the gas sensing properties of the sensitive materials [131,132]. For example, Chaitra et al.
prepared Al-doped ZnO thin films by the sol–gel method and spin-coating technique [133].
It was found that 2 at.% Al-doped ZnO thin films have the highest sensitivity to 3 ppm SO2
gas at 300 ◦ C, which was lower than the threshold limit. Kolhe et al. prepared Al-doped
ZnO thin films by chemical spray pyrolysis [134]. It was found that the doping of Al
in ZnO led to the fracture of thin nanofilms, resulting in more active sites. Al doping
also leads to the increase of oxygen vacancy-related defects and the change of crystal size
due to the difference of ion radius between Al3+ and Zn2+ ions. The doped sensor has
enhanced sensing characteristics, which also leads to the decrease of the optimal operating
temperature. Xiang et al. used the photochemical method to embed Ag nanoparticles
into ZnO nanorods and studied their gas-sensing properties [135]. It was found that Ag
nanoparticles embedded on the surface of ZnO nanorods could improve the performance
of the sensor. The response of ZnO nanorods to 50 ppm ethanol was almost three times that
of pure ZnO nanorods, and had long-term stability. After 100 days of exposure to ethanol
in 30 ppm, the response of the sensor had no obvious degradation.
Nanomaterials 2022, 12, 982 18 of 40
Figure9.9.Schematic
Figure Schematic of
of the
the reducing
reducinggas
gassensing
sensingmechanism
mechanismin in
thethe
CuO–ZnO C–SC–S
CuO–ZnO NWs. Ec and
NWs. EF EF
Ec and
indicate the conduction band energy and Fermi energy level, respectively, in cases of ZnO shell
indicate the conduction band energy and Fermi energy level, respectively, in cases of ZnO shell layers
layers (a) thinner and (b) thicker than ZnO’s Debye length. Reprinted with permission from Ref.
(a) thinner
[136]. and (b)
Copyright thicker
2016 than ZnO’s Debye length. Reprinted with permission from Ref. [136].
Elsevier.
Copyright 2016 Elsevier.
5.3. CuO-Based Gas Sensors
CuO is a typical p-type semiconductor oxide material with a band gap of 1.2–1.9 eV.
Because of its good electrical properties, chemical stability, catalytic activity and other
physical and chemical properties, CuO has been widely studied in the fields of catalysis,
optoelectronic devices, gas sensors and so on. CuO can respond to reducing gases at lower
operating temperatures, which attracts researchers to prepare different morphologies of
CuO, doped CuO and heterostructure CuO for gas sensors to study their gas sensing
Nanomaterials 2022, 12, 982 19 of 40
Figure
Figure 10.10.
TheThe preparation
preparation of CuO of
NTsCuO NTsNCs
and CuO andandCuO NCs and CO
CO gas-sensing gas-sensing
behaviors behaviors
of CuO NTs and of
andNCs
CuO CuO NCs
at the at the temperature
operation of 175 ◦ C with different
operation temperature of 175 °C COwith different(50–1000
concentrations CO concentrations
ppm).
ppm). Reprinted
Reprinted with from
with permission permission
Ref. [142].from Ref. [142].
Copyright Copyright 2018 Elsevier.
2018 Elsevier.
Although pure CuO as a sensitive material can be used to detect a variety of toxic and
harmful gases. However, it still faces some problems in practical applications, such as low
sensitivity, high working temperature, poor selectivity, long response/recovery time and so
on. For this reason, researchers use doping, recombination and other methods to improve
the gas sensing performance of CuO-based gas sensors, and achieved some remarkable
results. The doping of precious metals or rare earth elements can greatly increase the active
sites of CuO gas-sensing reaction, which is beneficial to the adsorption of gas molecules
on the sensitive material surface, and most of the dopants have strong catalytic activity,
which can further enhance the gas sensing reaction. For example, Hu et al. prepared
CuO nanoflowers with different Pd-doping concentrations by simple water-bath heating
method [143]. Compared with pure CuO, the specific surface area of CuO nanoflowers
with a size of about 400 nm prepared when the mass fraction of Pd was 1.25% increased by
1.8 times, and the response (Rg/Ra) to 50 ppm H2 S at 80 ◦ C was 123.4, which was 7.9 times
that of pure CuO. In addition, the gas sensor has good stability and repeatability. Tang et al.
prepared Pt-doped CuO nanoflowers by the same method, which significantly improved
the gas sensing performance of the sensor to H2 S gas [144]. When the amount of Pt doping
was 1.25 wt.%, the response of the sensor to 10 ppm H2 S at 40 ◦ C was 135.1, which was
13.1 times that of pure CuO. The researchers also selected other metal elements to dope
CuO, and achieved excellent results. For example, Mnethu et al. reported a highly sensitive
and selective Zn-doped CuO nano-chip-based sensor [145]. At 150 ◦ C, the response of 0.1
at.% Zn-doped CuO samples to 100 ppm xylene gas was 53. Bhuvaneshwari et al. reported
a Cr-doped CuO nanoboat, which significantly improves the sensing performance of NH3
in the concentration range of 100–600 ppm at room temperature [146]. The gas sensing test
results show that the sensitivity of CuO nanospheres doped with atomic fraction 6% Cr to
NH3 at room temperature was 2.5 times higher than that of undoped nanospheres. The
enhanced gas sensing performance is attributed to the increase of oxygen vacancy caused
by chromium doping, which makes the nanospheres absorb more surface oxygen, and
chromium doping also reduces the activation energy of the sensor at low temperatures.
Al-doped CuO [147], In-doped CuO [148], and Ag-doped CuO [149] also showed excellent
gas sensing properties for target gases.
The composite gas sensor can integrate the unique properties of the material and
improve the performance of the sensor through complementary enhancement. Researchers
have designed a variety of CuO-based composite gas sensors to improve the selectivity of
target gases, enhance gas sensing properties, shorten response/recovery time and reduce
the optimal operating temperature, especially semiconductor oxides with heterostructures.
For example, Sui et al. used the template-free hydrothermal method to grow multi-layer
heterogeneous CuO/NiO nanowires on ceramic tubes for the detection of H2 S gas [150].
The CuO/NiO-based sensor has a wide linear range in the 50~1000 ppb range and has good
repeatability, selectivity and long-term stability. At 133 ◦ C, the 2.84 at.% CuO modified
NiO showed good sensing properties, and the response to 5 ppm H2 S was 36.9, which
was 5.6 times higher than that of NiO. The detection limit of H2 S is further reduced
from 1 ppb of pure NiO sensor to 0.5 ppb. Park et al. synthesized SnO2 -CuO hollow
nanofibers by electrospinning and thermal processing, which can be used in the field
of H2 S gas sensing [151]. The electrospun nanofiber materials have the advantages of
large surface area, high porosity and permeability to air or moisture, which is conducive
to ionic diffusion and suitable for applications in gas sensors, lithium-ion batteries and
wound healing [152,153]. SnO2 -CuO nanotubes increase the specific surface area, decrease
the working temperature and improve the sensing performance of H2 S. At the working
temperature of 200 ◦ C, the sensitivity of hollow SnO2 -CuO nanotubes to 5 ppm H2 S was
1395 and the response time was 5.27 s. Liang et al. also prepared the heterostructure of
In2 O3 nanofibers supported on CuO by electrospinning and studied the sensing properties
of H2 S [154]. The gas sensor based on the heterostructure had a high sensitivity to 5 ppm
H2 S gas at 150 ◦ C, which was 225 times higher than that based on pure In2 O3 , even at
room temperature. The above research results show that the construction of heterojunction
Nanomaterials 2022, 12, 982 21 of 40
composites can effectively improve the sensitivity and selectivity of the gas sensor, reduce
the working temperature, accelerate the response/recovery speed and prolong the life of
the sensor.
the working temperature of 180 ◦ C, the fabricated sensor showed higher sensitivity, faster
response and recovery speed, as well as better selectivity.
temperature, which is helpful for achieving the ultra-sensitive and high-accuracy detection
of harmful gases [181]. Wang et al. used a one-step hydrothermal method to synthesize
rGO/CuO/ZnO ternary composites to form nanoscale p-n junctions on rGO substrates.
The gas sensors prepared by using the created materials showed excellent response charac-
teristics and good selectivity to acetone, which was almost 1.5 and 2.0 times higher than
those of CuO/ZnO and rGO/ZnO-based gas sensors, respectively [182]. It has become
a research hotspot in the field of gas sensing to improve the performance of gas sensing
materials by combining metal oxide semiconductor materials with 2D graphene materials.
Wang et al. successfully assembled SnO2 onto the surface of GO, and studied the gas-
sensing properties of formaldehyde. It was found that SnO2 can be assembled on the surface
of GO in a large area, and has good gas-sensing response properties to formaldehyde [183].
Wang et al. synthesized SnO2 nanoparticles onto RGO through a hydrothermal reduction to
form SnO2 -RGO composites, which exhibited promising application for room-temperature
gas sensing of NO2 [184]. Yin et al. [185] synthesized SnO2 /rGO nanocomposites with
the SnO2 particle sizes of 3–5 nm uniformly immobilized on rGO nanosheets through a
heteronuclear growth process by a simple redox reaction under microwave irradiation.
The SnO2 /rGO nanostructure on the surface has a sesame cake-like layered structure and
an ultra-high specific surface area of 2110.9 m2 ·g−1 . Compared with SnO2 nanocrystals
(5–10 nm), the designed SnO2 /rGO nanostructures have stronger gas-sensing behavior
due to the unique hierarchical structure, high specific surface area, and synergistic effect of
SnO2 nanoparticles and rGO nanosheets. At the optimal operating temperature of 100 ◦ C,
the SnO2 /rGO-based gas sensor has a sensitivity as high as 78 and a response time as short
as 7 s when exposed to 10 ppm H2 S. In a similar study, Kim et al. [176] also synthesized a
graphene/SnO2 composite material by the microwave-assisted method, and then sprayed
the material onto SiO2 substrate to fabricate a NO2 gas sensor. At the optimal operating
temperature of 150 ◦ C, the response value of 1 ppm NO2 was 24.7. Its excellent gas-sensing
response may be related to the homojunction between SnO2 , the heterojunction between
SnO2 and graphene, and the interstitial defects of Sn atoms in the SnO2 lattice.
Using the novel properties of SnO2 , multi-walled carbon nanotubes (MWCNTs), and
rGO, Tyagi et al. developed a hybrid nanocomposite sensor for efficient detection of SO2
gas [186]. The rGO-SnO2 and MWCNT-SnO2 composites were prepared by physical mixing
and spin-coated onto the surface of Pt interdigital electrodes for SO2 gas detection. The
sensing response of the bare SnO2 sensor to 500 ppm SO2 gas at 220 ◦ C was 1.2. However,
for the same concentration of SO2 gas, the enhanced sensing response of the MWCNT-SnO2
sensor was 5 at 60 ◦ C, while the maximum sensing response of the rGO-SnO2 sensor was
22 at 60 ◦ C. The enhanced SO2 gas sensing performance of these composites is mainly
attributed to the p-n heterojunction formed at the interface between n-type SnO2 and p-type
rGO or MWCNT.
Yu et al. [187] successfully synthesized α-Fe2 O3 @graphene nanocomposites using a
simple low-temperature hydrolysis and calcination process, and fabricated the synthesized
materials into gas sensors to detect different gases. Their prepared α-Fe2 O3 @graphite
nanocomposites consist of porous α-Fe2 O3 nanorods stably and orderly grown on graphitic
nanosheets, as shown in Figure 12a,b. The length of α-Fe2 O3 nanorods is related to the
reaction time. When the reaction time was 12 h, the length reached a maximum value of
about 200–300 nm, and the pore size was about 3.7 nm. Compared with pure α-Fe2 O3 , the
α-Fe2 O3 @graphite nanocomposite-based sensor exhibited higher sensing performance for
acetone. At the optimal temperature of 260 ◦ C, the response of α-Fe2 O3 @graphite nanocom-
posites to 50 ppm acetone reaches a maximum value of 16.9, which was 2.2 times that
of α-Fe2 O3 , as shown in Figure 12c. The high sensing performance was attributed to the
porous structure, high specific surface area, and p-n heterostructure of α-Fe2 O3 @graphite
nanocomposites and the high temperature stability of graphite. When α-Fe2 O3 was re-
combined with graphite, due to the large gradient of the same carrier concentration, the
electrons in α-Fe2 O3 and the holes in graphite diffuse in opposite directions, so that a built-
in electric field is formed between the interfaces, and the electrons in the depletion layer.
Nanomaterials 2022, 12, x FOR PEER REVIEW 26 of 40
porous structure, high specific surface area, and p-n heterostructure of α-Fe2O3@graphite
Nanomaterials 2022, 12, 982 nanocomposites and the high temperature stability of graphite. When α-Fe2O3 was recom- 26 of 40
bined with graphite, due to the large gradient of the same carrier concentration, the elec-
trons in α-Fe2O3 and the holes in graphite diffuse in opposite directions, so that a built-in
electric
The fieldbands
energy is formed
bendbetween
until thethe interfaces,
system andequilibrium
reaches the electronsatinthe
theFermi
depletion
levellayer.
(EF ), The
leading
energy bands bend until the system reaches equilibrium at the Fermi level (E F), leading to
to the formation of a p-n heterojunction. Once the α-Fe2 O3 @graphite heterojunction sensor
isthe formation
exposed of a p-n heterojunction.
to acetone gas, the oxygen Once the α-Fe
anions 2O3@graphite heterojunction sensor is
adsorbed on the sample surface undergo a
redox reaction with acetone molecules and release on
exposed to acetone gas, the oxygen anions adsorbed the sample
electrons backsurface undergo
into α-Fe a re-
2 O3 , resulting
dox reaction with acetone molecules and release electrons back into α-Fe 2O3, resulting in
in a decrease in the resistance of the sensor. At the same time, acetone releases electrons
a decrease in the resistance of the sensor. At the same time, acetone releases electrons to
to combine with holes in p-type graphite, resulting in a decrease in hole concentration.
combine with holes in p-type graphite, resulting in a decrease in hole concentration. The
The reduction of holes in graphite leads to an increase in electrons and reduces the con-
reduction of holes in graphite leads to an increase in electrons and reduces the concentra-
centration gradient of the same carriers on both sides of the p-n heterojunction. Therefore,
tion gradient of the same carriers on both sides of the p-n heterojunction. Therefore, the
the diffusion of carriers was weakened and the barrier height of the depletion layer was
diffusion of carriers was weakened and the barrier height of the depletion layer was re-
reduced, which further reduced the resistance of the α-Fe2 O3 @graphite sensor, as shown in
duced, which further reduced the resistance of the α-Fe2O3@graphite sensor, as shown in
Figure 12e.
Figure 12e.
Co3 O4 , as a direct bandgap p-type metal oxide semiconductor material, has also re-
ceived extensive attention in gas sensors and other fields due to its outstanding advantages
such as strong corrosion resistance and non-toxicity. For instance, Zhou et al. [188] studied
the performance of a Co3 O4 -based gas sensor and found that it can only work at tempera-
tures over 200 ◦ C. Zhang et al. [189] proposed that the rGO/Co3 O4 nanocomposite-based
Nanomaterials 2022, 12, 982 27 of 40
sensor can realize the detection of NO2 gas at room temperature. Srirattanapibul et al. [190]
prepared a Co3 O4 -modified rGO (rGO/Co3 O4 ) nanocomposite-based gas sensor by a
solvothermal method. Co3 O4 nanoparticles were distributed on and between the rGO
flakes, and their dosage changed the bandgap and gas sensing properties of rGO. Deco-
rating rGO with Co3 O4 nanoparticles promoted the formation of the Co-C bridges, which
enable the exchange of charge carriers between Co3 O4 nanoparticles and rGO flakes,
thereby increasing the number of sites for gas reactions to occur and improving gas sensing
performance. The as-prepared 25% rGO/Co3 O4 -based gas sensor has a sensitivity of 1.78%
and a response time of 351 s towards 20 ppm NH3 .
Many studies on the synthesis of the composites by combing graphene with other
metal oxides have also been carried out. For example, Hao et al. [191] synthesized
WO3 /rGO porous nanocomposites using a simple hydrothermal and annealing process.
The material-based gas sensor showed good sensitivity to NO2 and some volatile organic
compounds. In another study, Ye et al. [192] fabricated uniform TiO2 /rGO membranes
with enhanced NH3 responsiveness by stepwise deposition of GO and TiO2 layers followed
by simple thermal treatment. To make it more clear, here we summarize the gas sensing
properties of the above graphene/metal oxide nanocomposite-based gas sensors, as shown
in Table 1.
Table 1. Sensors based on MOS modified with graphene/GO/rGO gas sensing performances.
Working
Sensor Materials Analyte Response Refs.
Temperature
ZnO/rGO NO2 17.4% (100 ppm) RT [179]
ZnO/rGO NO2 25.6% (5 ppm) RT [180]
ZnO/rGO NH3 7.2% (1 ppm) RT [181]
SnO2 /GO HCHO 32 (100 ppm) 120 ◦ C [183]
SnO2 /rGO H2 S 78 (10 ppm) 100 ◦ C [185]
Graphite/SnO2 NO2 24.7 (1 ppm) 150 ◦ C [176]
rGO/SnO2 SO2 22 (500 ppm) 60 ◦ C [186]
α-Fe2 O3 @graphite C3 H6 O 16.9 (50 ppm) 260 ◦ C [187]
rGO/Co3 O4 NO2 26.8% (5 ppm) RT [189]
rGO/Co3 O4 NH3 1.78% (20 ppm) RT [190]
WO3 /rGO NO2 4.3 (10 ppm) 90 ◦ C [191]
TiO2 /rGO NH3 0.62 (10 ppm) RT [192]
13.(a)(a)
Figure13.
Figure Schematic
Schematic illustration
illustration of theof the preparation
preparation process process for MoS
for MoS2/SnO 2 /SnO2 nanohybrids.
2 nanohybrids. The in- The
set photographs show the MoS suspension in water before and after adding the SnCl solution.
inset photographs show the MoS2 suspension in water before and after adding the SnCl4 solution.
2 4 (b)
TEM images, (c) SAED pattern and (d) HRTEM images of the MoS2/SnO2 nanohybrids. The inset of
(b) TEM images, (c) SAED pattern and (d) HRTEM images of the MoS2 /SnO2 nanohybrids. The inset
of (d) shows a typical SnO2 nanocrystal on the MoS2 surface. (e) The room temperature dynamic
sensing response of MoS2 nanosheets with and without SnO2 NC decoration against 10 ppm NO2
in a dry air environment, indicating the SnO2 NCs significantly enhanced the stability of MoS2 in
the dry air. (f) Band diagram of the MoS2 /SnO2 nanohybrid. The EF, SnO2 and EF, MoS2 are Fermi
levels of SnO2 and MoS2 , respectively. The CB and VB are the conductance and valance band edges
of MoS2 , respectively. d is the thickness of the electron depletion zone, and ΦB is the Schottky barrier
height. Reprinted with permission from Ref. [194]. Copyright 2015 John Wiley & Sons, Inc.
In addition to SnO2 , ZnO is another wide band gap n-type semiconductor for the
fabrication of high-performance gas sensors. Yan et al. synthesized the ZnO/MoS2 com-
posite structure by coating ZnO nanoparticles onto MoS2 nanosheets through a two-step
Nanomaterials 2022, 12, 982 29 of 40
hydrothermal method [195]. Among the composites, MoS2 is a multi-stage structure com-
posed of nanosheets with a thickness of 5~10 nm, and the size of ZnO particles was about
8 nm. The response value of the composite-based gas sensor to 50 ppm ethanol reached 42.8
at the operating temperature of 260 ◦ C Han et al. designed a MoS2 /ZnO heterostructure on
the MoS2 nanosheets obtained by liquid phase exfoliation by a wet chemical method, and
achieved efficient detection of NO2 gas at room temperature [196]. After surface modifica-
tion, ZnO nanoparticles had a good response to 5 ppm NO2 , and the response value reached
3050%, which was 11 times higher than that of pure phase MoS2 nanoparticles. In addition,
the recoverability of the heterostructure was improved to more than 90% without auxiliary
means, and the sensor also had the characteristics of fast response speed (40 s), reliable
long-term stability within 10 weeks, good selectivity, and a low detection concentration of
50 ppb. The enhanced sensing performance of MoS2 /ZnO heterostructure can be attributed
to the unique 2D/0D heterostructure, synergistic effect and the p-n heterojunction between
ZnO nanoparticles and MoS2 nanosheets.
Besides, Zhao et al. reported the hydrothermal synthesis of MoS2 -modified TiO2
nanotube composites and studied their gas sensing properties [76]. TiO2 nanotubes are
filled and covered by 1–3 layers of flake MoS2 nanosheets. The formed MoS2 -TiO2 com-
posites revealed excellent sensing properties and high sensitivity to ethanol vapor at low
operating temperatures, and their sensitivity was almost 11 times that of TiO2 nanotubes.
The response to 100 ppm ethanol gas was ~14.2 and the optimum working temperature was
as low as 150 ◦ C. Zhang et al. prepared CuO/MoS2 heterostructure sensing films on the
substrate by layer-by-layer self-assembly technique [197]. Compared with the pure phase
CuO and MoS2 , the formed CuO/MoS2 composite structure exhibited higher response,
shorter response/recovery time, better repeatability, higher selectivity, and longer-term
stable H2 S detection performance. The excellent H2 S sensing properties are mainly due to
the existence of a large number of oxygen and sulfur vacancies in the composite structure
of CuO nanorods and MoS2 nanosheets, which brings a large number of active sites for gas
adsorption. In addition, the synergistic effect of binary nanostructures and the modulation
of electron transfer by the formation of p-n heterojunction at the material interface between
p-type CuO semiconductors and n-type MoS2 semiconductors promoted the performance
of the composite structure. Ikram et al. prepared MoO2 /MoS2 nanonetworks by control-
lable vulcanization and successfully applied them to the efficient detection of NO2 gas at
room temperature [198]. The response value of the composite structure to 100 ppm NO2
gas was 19.4, and it had ultra-fast response speed and recovery speed, and the response and
recovery time were 1.06 s and 22.9 s, respectively. The excellent gas-sensing performance of
the sensor can be attributed to the synergistic effect between MoS2 nanosheets and MoO2
nanoparticles. The defects in the synthesis process provided more active sites for NO2 gas
molecules, and the formation of p-n heterojunction accelerated the charge transfer between
NO2 and gas molecules.
In addition to MoS2 , other transition metal dichalcogenides and metal oxide com-
posites have been also often used as sensitive materials for gas detection. For example,
Qin et al. prepared 2D WS2 nanosheets/TiO2 quantum dots composites by chemical strip-
ping method, which have been successfully used in room temperature NH3 sensing. The
fabricated gas sensors exhibited a quicker sensing response to 250 ppm NH3 , which was
almost 17 times that of the original WS2 [199]. Gu et al. prepared SnO2 /SnS2 heterojunc-
tion nanocomposites by the in situ high-temperature oxidizer SnS2 , which significantly
improved the response to NO2 and decreased the working temperature [200]. To make it
more clear, the gas sensing properties of the transition metal dichalcogenides/metal oxide
composite-based gas sensors are shown in Table 2.
Nanomaterials 2022, 12, 982 30 of 40
Table 2. Sensors based on MOS modified with TMDs gas sensing performances.
Sensor Working
Analyte Response Refs.
Materials Temperature
SnO2 /MoS2 C3 H9 N 106.3 (200 ppm) 230 ◦ C [193]
SnO2 /MoS2 NO2 28% (10 ppm) RT [194]
ZnO/MoS2 C2 H6 O 42.8 (50 ppm) 260 ◦ C [195]
ZnO/MoS2 NO2 3050% (5 ppm) RT [196]
MoS2 /TiO2 C2 H6 O 14.2 (100 ppm) 150 ◦ C [76]
CuO/MoS2 H2 S 61 (30 ppm) RT [197]
MoO2 /MoS2 NO2 19.4 (100 ppm) RT [198]
TiO2 QDs/WS2 NH3 43.7% (250 ppm) RT [199]
SnO2 /SnS2 NO2 5.3 (8 ppm) 80 ◦ C [200]
Yin et al. [204] synthesized hierarchical Fe2 O3 /WO3 nanocomposites with ultra-high
specific surface area composed of Fe2 O3 nanoparticles and single-crystal WO3 nanosheets
through microwave heating and in situ growth. The BET specific surface area of the sample
that prepared with 5 wt.% Fe2 O3 /WO3 by this process was as high as 1207 m2 ·g−1 , which
was 5.9 times that of the corresponding WO3 nanosheets (203 m2 ·g−1 ). The significant
enhancement of the specific surface area of the Fe2 O3 @WO3 samples was attributed to the
hierarchical structure of the prepared composite materials, in which the monolayer and
unconnected Fe2 O3 nanoparticles are tightly anchored to the surface of the WO3 nanosheets,
so that the inner surface or interface of the aggregated polycrystal is entirely the outer
surface. The gas-sensing performance tests indicated that Fe2 O3 @WO3 nanocomposites
exhibited high response and selectivity towards H2 S at low operating temperatures due to
the synergistic effect of the components of Fe2 O3 @WO3 nanocomposites and the hierarchi-
cal microstructure with ultra-high specific surface area. At 150 ◦ C, the fabricated gas sensor
showed a response to 10 ppm H2 S of as high as 192, which was four times that of the WO3
nanosheet-based gas sensor.
Author Contributions: Conceptualization, T.L., W.Y. and G.W.; reference analysis, T.L., Y.S., S.G.,
P.X. and S.W.; resources, all authors; writing—original draft preparation, T.L.; writing—review and
editing, T.L., H.K., G.Y. and G.W.; supervision, G.W.; project administration, T.L. and G.W.; funding
acquisition, G.W. All authors have read and agreed to the published version of the manuscript.
Funding: Taishan Scholars Program of Shandong Province (No. tsqn201909104) and the High-Grade
Talents Plan of Qingdao University.
Data Availability Statement: The data presented in this study are available in insert article.
Acknowledgments: The authors thank the financial support from Taishan Scholars Program of
Shandong Province (No. tsqn201909104) and the High-Grade Talents Plan of Qingdao University.
Nanomaterials 2022, 12, 982 33 of 40
References
1. Chatterjee, S.G.; Chatterjee, S.; Ray, A.K.; Chakraborty, A.K. Graphene-metal oxide nanohybrids for toxic gas sensor: A review.
Sens. Actuators B Chem. 2015, 221, 1170–1181. [CrossRef]
2. Ab Kadir, R.; Li, Z.Y.; Sadek, A.; Rani, R.A.; Zoolfakar, A.S.; Field, M.R.; Ou, J.Z.; Chrimes, A.F.; Kalantar-zadeh, K. Electrospun
Granular Hollow SnO2 Nanofibers Hydrogen Gas Sensors Operating at Low Temperatures. J. Phys. Chem. C 2014, 118, 3129–3139.
[CrossRef]
3. Zhang, D.Z.; Wu, J.F.; Li, P.; Cao, Y.H. Room-temperature SO2 gas-sensing properties based on a metal-doped MoS2 nanoflower:
An experimental and density functional theory investigation. J. Mater. Chem. A 2017, 5, 20666–20677. [CrossRef]
4. Zhang, Y.F.; Thorburn, P.J. Handling missing data in near real-time environmental monitoring: A system and a review of selected
methods. Future Gener. Comp. Syst. 2022, 128, 63–72. [CrossRef]
5. Di Natale, C.; Paolesse, R.; Martinelli, E.; Capuano, R. Solid-state gas sensors for breath analysis: A review. Anal. Chim. Acta 2014,
824, 1–17. [CrossRef] [PubMed]
6. Guntner, A.T.; Pineau, N.J.; Mochalski, P.; Wiesenhofer, H.; Agapiou, A.; Mayhew, C.A.; Pratsinis, S.E. Sniffing Entrapped Humans
with Sensor Arrays. Anal. Chem. 2018, 90, 4940–4945.
7. Righettoni, M.; Amann, A.; Pratsinis, S.E. Breath analysis by nanostructured metal oxides as chemo-resistive gas sensors. Mater.
Today 2015, 18, 163–171. [CrossRef]
8. Ponzoni, A.; Comini, E.; Concina, I.; Ferroni, M.; Falasconi, M.; Gobbi, E.; Sberveglieri, V.; Sberveglieri, G. Nanostructured Metal
Oxide Gas Sensors, a Survey of Applications Carried out at SENSOR Lab, Brescia (Italy) in the Security and Food Quality Fields.
Sensors 2012, 12, 17023–17045. [CrossRef]
9. Kim, W.S.; Kim, H.C.; Hong, S.H. Gas sensing properties of MoO3 nanoparticles synthesized by solvothermal method. J. Nanopart.
Res. 2010, 12, 1889–1896. [CrossRef]
10. Chen, Y.J.; Xiao, G.; Wang, T.S.; Zhang, F.; Ma, Y.; Gao, P.; Zhu, C.L.; Zhang, E.D.; Xu, Z.; Li, Q.H. α-MoO3 /TiO2 core/shell
nanorods: Controlled-synthesis and low-temperature gas sensing properties. Sens. Actuators B Chem. 2011, 155, 270–277.
[CrossRef]
11. Lim, S.K.; Hwang, S.H.; Kim, S.; Park, H. Preparation of ZnO nanorods by microemulsion synthesis and their application as a CO
gas sensor. Sens. Actuators B Chem. 2011, 160, 94–98. [CrossRef]
12. Yang, B.; Wang, C.; Xiao, R.; Yu, H.Y.; Wang, J.X.; Xu, J.L.; Liu, H.M.; Xia, F.; Xiao, J.Z. CO Response Characteristics of NiFe2 O4
Sensing Material at Elevated Temperature. J. Electrochem. Soc. 2019, 166, B956–B960. [CrossRef]
13. Kim, H.J.; Lee, J.H. Highly sensitive and selective gas sensors using p-type oxide semiconductors: Overview. Sens. Actuators B
Chem. 2014, 192, 607–627. [CrossRef]
14. Volanti, D.P.; Felix, A.A.; Orlandi, M.O.; Whitfield, G.; Yang, D.J.; Longo, E.; Tuller, H.L.; Varela, J.A. The Role of Hierarchical
Morphologies in the Superior Gas Sensing Performance of CuO-Based Chemiresistors. Adv. Funct. Mater. 2013, 23, 1759–1766.
[CrossRef]
15. Virji, S.; Huang, J.X.; Kaner, R.B.; Weiller, B.H. Polyaniline nanofiber gas sensors: Examination of response mechanisms. Nano Lett.
2004, 4, 491–496. [CrossRef]
16. Gilbertson, L.M.; Busnaina, A.A.; Isaacs, J.A.; Zimmerman, J.B.; Eckelman, M.J. Life Cycle Impacts and Benefits of a Carbon
Nanotube-Enabled Chemical Gas Sensor. Environ. Sci. Technol. 2014, 48, 11360–11368. [CrossRef]
17. Ma, H.Y.; Li, Y.W.; Yang, S.X.; Cao, F.; Gong, J.; Deng, Y.L. Effects of Solvent and Doping Acid on the Morphology of Polyaniline
Prepared with the Ice-Templating Method. J. Phys. Chem. C 2010, 114, 9264–9269. [CrossRef]
18. Patil, V.L.; Vanalakar, S.A.; Patil, P.S.; Kim, J.H. Fabrication of nanostructured ZnO thin films based NO2 gas sensor via SILAR
technique. Sens. Actuators B Chem. 2017, 239, 1185–1193. [CrossRef]
19. Meng, F.L.; Zheng, H.X.; Chang, Y.L.; Zhao, Y.; Li, M.Q.; Wang, C.; Sun, Y.F.; Liu, J.H. One-Step Synthesis of Au/SnO2 /RGO
Nanocomposites and Their VOC Sensing Properties. IEEE Trans. Nanotechnol. 2018, 17, 212–219. [CrossRef]
20. Choi, S.J.; Lee, I.; Jang, B.H.; Youn, D.Y.; Ryu, W.H.; Park, C.O.; Kim, I.D. Selective Diagnosis of Diabetes Using Pt-Functionalized
WO3 Hemitube Networks As a Sensing Layer of Acetone in Exhaled Breath. Anal. Chem. 2013, 85, 1792–1796. [CrossRef]
21. Navale, S.T.; Liu, C.; Yang, Z.; Patil, V.B.; Cao, P.; Du, B.; Mane, R.S.; Stadler, F.J. Low-temperature wet chemical synthesis strategy
of In2 O3 for selective detection of NO2 down to ppb levels. J. Alloys Compd. 2018, 735, 2102–2110. [CrossRef]
22. Fine, G.F.; Cavanagh, L.M.; Afonja, A.; Binions, R. Metal Oxide Semi-Conductor Gas Sensors in Environmental Monitoring.
Sensors 2010, 10, 5469–5502. [CrossRef] [PubMed]
23. Suematsu, K.; Shin, Y.; Ma, N.; Oyama, T.; Sasaki, M.; Yuasa, M.; Kida, T.; Shimanoe, K. Pulse-Driven Micro Gas Sensor Fitted
with Clustered Pd/SnO2 Nanoparticles. Anal. Chem. 2015, 87, 8407–8415. [CrossRef] [PubMed]
24. Chen, Y.; Yang, G.Z.; Liu, B.; Kong, H.; Xiong, Z.; Guo, L.; Wei, G. Biomineralization of ZrO2 nanoparticles on graphene
oxide-supported peptide/cellulose binary nanofibrous membranes for high-performance removal of fluoride ions. Chem. Eng. J.
2022, 430, 132721. [CrossRef]
25. Liu, B.; Jiang, M.; Zhu, D.Z.; Zhang, J.M.; Wei, G. Metal-organic frameworks functionalized with nucleic acids and amino acids
for structure- and function-specific applications: A tutorial review. Chem. Eng. J. 2022, 428, 131118. [CrossRef]
Nanomaterials 2022, 12, 982 34 of 40
26. Liu, X.H.; Ma, T.T.; Pinna, N.; Zhang, J. Two-Dimensional Nanostructured Materials for Gas Sensing. Adv. Funct. Mater. 2017,
27, 1702168. [CrossRef]
27. Neri, G. Thin 2D: The New Dimensionality in Gas Sensing. Chemosensors 2017, 5, 21. [CrossRef]
28. Anichini, C.; Czepa, W.; Pakulski, D.; Aliprandi, A.; Ciesielski, A.; Samori, P. Chemical sensing with 2D materials. Chem. Soc. Rev.
2018, 47, 4860–4908. [CrossRef]
29. Mao, S.; Chang, J.B.; Pu, H.H.; Lu, G.H.; He, Q.Y.; Zhang, H.; Chen, J.H. Two-dimensional nanomaterial-based field-effect
transistors for chemical and biological sensing. Chem. Soc. Rev. 2017, 46, 6872–6904. [CrossRef]
30. Kim, T.H.; Kim, Y.H.; Park, S.Y.; Kim, S.Y.; Jang, H.W. Two-Dimensional Transition Metal Disulfides for Chemoresistive Gas
Sensing: Perspective and Challenges. Chemosensors 2017, 5, 15. [CrossRef]
31. Kakanakova-Georgieva, A.; Giannazzo, F.; Nicotra, G.; Cora, I.; Gueorguiev, G.K.; Persson, P.O.A.; Pecz, B. Material proposal for
2D indium oxide. Appl. Surf. Sci. 2021, 548, 149275. [CrossRef]
32. Dos Santos, R.B.; Rivelino, R.; Gueorguiev, G.K.; Kakanakova-Georgieva, A. Exploring 2D structures of indium oxide of different
stoichiometry. CrystEngComm 2021, 23, 6661–6667. [CrossRef]
33. Wu, J.; Yang, Y.; Yu, H.; Dong, X.T.; Wang, T.T. Ultra-efficient room-temperature H2 S gas sensor based on NiCo2 O4 /r-GO
nanocomposites. New J. Chem. 2019, 43, 10501–10508. [CrossRef]
34. Zhang, D.Z.; Jiang, C.X.; Sun, Y.E. Room-temperature high-performance ammonia gas sensor based on layer-by-layer self-
assembled molybdenum disulfide/zinc oxide nanocomposite film. J. Alloys Compd. 2017, 698, 476–483. [CrossRef]
35. Nakate, U.T.; Ahmad, R.; Patil, P.; Bhat, K.S.; Wang, Y.S.; Mahmoudi, T.; Yu, Y.T.; Suh, E.K.; Hahn, Y.B. High response and
low concentration hydrogen gas sensing properties using hollow ZnO particles transformed from polystyrene@ZnO core-shell
structures. Int. J. Hydrogen Energy 2019, 44, 15677–15688. [CrossRef]
36. Nakate, U.T.; Lee, G.H.; Ahmad, R.; Patil, P.; Hahn, Y.B.; Yu, Y.T.; Suh, E.K. Nano-bitter gourd like structured CuO for enhanced
hydrogen gas sensor application. Int. J. Hydrogen Energy 2018, 43, 22705–22714. [CrossRef]
37. Nakate, U.T.; Lee, G.H.; Ahmad, R.; Patil, P.; Bhopate, D.P.; Hahn, Y.B.; Yu, Y.T.; Suh, E.K. Hydrothermal synthesis of p-type
nanocrystalline NiO nanoplates for high response and low concentration hydrogen gas sensor application. Ceram. Int. 2018, 44,
15721–15729. [CrossRef]
38. Yang, S.X.; Jiang, C.B.; Wei, S.H. Gas sensing in 2D materials. Appl. Phys. Rev. 2017, 4, 021304. [CrossRef]
39. Yue, Q.; Shao, Z.Z.; Chang, S.L.; Li, J.B. Adsorption of gas molecules on monolayer MoS2 and effect of applied electric field.
Nanoscale Res. Lett. 2013, 8, 425. [CrossRef]
40. Wang, D.; Tian, L.; Li, H.; Wan, K.; Yu, X.; Wang, P.; Chen, A.; Wang, X.; Yang, J. Mesoporous Ultrathin SnO2 Nanosheets in Situ
Modified by Graphene Oxide for Extraordinary Formaldehyde Detection at Low Temperatures. ACS Appl. Mater. Interfaces 2019,
11, 12808–12818. [CrossRef]
41. Rothschild, A.; Komem, Y. The effect of grain size on the sensitivity of nanocrystalline metal-oxide gas sensors. J. Appl. Phys.
2004, 95, 6374–6380. [CrossRef]
42. Xu, C.N.; Tamaki, J.; Miura, N.; Yamazoe, N. Grain-size effects on gas sensitivity of porous SnO2 -based elements. Sens. Actuators
B Chem. 1991, 3, 147–155. [CrossRef]
43. Sun, Y.F.; Liu, S.B.; Meng, F.L.; Liu, J.Y.; Jin, Z.; Kong, L.T.; Liu, J.H. Metal oxide nanostructures and their gas sensing properties:
A review. Sensors 2012, 12, 2610–2631. [CrossRef] [PubMed]
44. Han, M.A.; Kim, H.J.; Lee, H.C.; Park, J.S.; Lee, H.N. Effects of porosity and particle size on the gas sensing properties of SnO2
films. Appl. Surf. Sci. 2019, 481, 133–137. [CrossRef]
45. Min, B.K.; Choi, S.D. SnO2 thin film gas sensor fabricated by ion beam deposition. Sens. Actuators B Chem. 2004, 98, 239–246.
[CrossRef]
46. Shoyama, M.; Hashimoto, N. Effect of poly ethylene glycol addition on the microstructure and sensor characteristics of SnO2 thin
films prepared by sol-gel method. Sens. Actuators B Chem. 2003, 93, 585–589. [CrossRef]
47. Boudiba, A.; Zhang, C.; Bittencourt, C.; Umek, P.; Olivier, M.G.; Snyders, R.; Debliquy, M. SO2 gas sensors based on WO3
nanostructures with different morphologies. Procedia Eng. 2012, 47, 1033–1036. [CrossRef]
48. Jia, Q.Q.; Ji, H.M.; Wang, D.H.; Bai, X.; Sun, X.H.; Jin, Z.G. Exposed facets induced enhanced acetone selective sensing property of
nanostructured tungsten oxide. J. Mater. Chem. A 2014, 2, 13602–13611. [CrossRef]
49. Lu, Y.Y.; Zhan, W.W.; He, Y.; Wang, Y.T.; Kong, X.J.; Kuang, Q.; Xie, Z.X.; Zheng, L.S. MOF-Templated Synthesis of Porous Co3 O4
Concave Nanocubes with High Specific Surface Area and Their Gas Sensing Properties. ACS Appl. Mater. Interfaces 2014, 6,
4186–4195. [CrossRef]
50. Wang, L.L.; Zhang, R.; Zhou, T.T.; Lou, Z.; Deng, J.N.; Zhang, T. Concave Cu2 O octahedral nanoparticles as an advanced sensing
material for benzene (C6 H6 ) and nitrogen dioxide (NO2 ) detection. Sens. Actuators B Chem. 2016, 223, 311–317. [CrossRef]
51. Dos Santos, R.B.; Rivelino, R.; Mota, F.D.; Gueorguiev, G.K.; Kakanakova-Georgieva, A. Dopant species with Al-Si and N-Si
bonding in the MOCVD of AlN implementing trimethylaluminum, ammonia and silane. J. Phys. D Appl. Phys. 2015, 48, 295104.
[CrossRef]
52. Kakanakova-Georgieva, A.; Gueorguiev, G.K.; Yakimova, R.; Janzen, E. Effect of impurity incorporation on crystallization in AlN
sublimation epitaxy. J. Appl. Phys. 2004, 96, 5293–5297. [CrossRef]
53. Muller, S.A.; Degler, D.; Feldmann, C.; Turk, M.; Moos, R.; Fink, K.; Studt, F.; Gerthsen, D.; Barsan, N.; Grunwaldt, J.D. Exploiting
Synergies in Catalysis and Gas Sensing using Noble Metal-Loaded Oxide Composites. ChemCatChem 2018, 10, 864–880. [CrossRef]
Nanomaterials 2022, 12, 982 35 of 40
54. Fedorenko, G.; Oleksenko, L.; Maksymovych, N.; Skolyar, G.; Ripko, O. Semiconductor Gas Sensors Based on Pd/SnO2
Nanomaterials for Methane Detection in Air. Nanoscale Res. Lett. 2017, 12, 329. [CrossRef]
55. Barbosa, M.S.; Suman, P.H.; Kim, J.J.; Tuller, H.L.; Varela, J.A.; Orlandi, M.O. Gas sensor properties of Ag− and Pd-decorated SnO
micro-disks to NO2 , H2 and CO: Catalyst enhanced sensor response and selectivity. Sens. Actuators B Chem. 2017, 239, 253–261.
[CrossRef]
56. Zhang, X.L.; Song, D.L.; Liu, Q.; Chen, R.R.; Liu, J.Y.; Zhang, H.S.; Yu, J.; Liu, P.L.; Wang, J. Designed synthesis of Co-doped
sponge-like In2 O3 for highly sensitive detection of acetone gas. CrystEngComm 2019, 21, 1876–1885. [CrossRef]
57. Ma, J.H.; Ren, Y.; Zhou, X.R.; Liu, L.L.; Zhu, Y.H.; Cheng, X.W.; Xu, P.C.; Li, X.X.; Deng, Y.H.; Zhao, D.Y. Pt Nanoparticles
Sensitized Ordered Mesoporous WO3 Semiconductor: Gas Sensing Performance and Mechanism Study. Adv. Funct. Mater. 2018,
28, 1705268. [CrossRef]
58. Ju, D.X.; Xu, H.Y.; Xu, Q.; Gong, H.B.; Qiu, Z.W.; Guo, J.; Zhang, J.; Cao, B.Q. High triethylamine-sensing properties of NiO/SnO2
hollow sphere P–N heterojunction sensors. Sens. Actuators B Chem. 2015, 215, 39–44. [CrossRef]
59. Kida, T.; Nishiyama, A.; Hua, Z.Q.; Suematsu, K.; Yuasa, M.; Shimanoe, K. WO3 Nano lamella Gas Sensor: Porosity Control Using
SnO2 Nanoparticles for Enhanced NO2 Sensing. Langmuir 2014, 30, 2571–2579. [CrossRef]
60. Mishra, R.K.; Murali, G.; Kim, T.H.; Kim, J.H.; Lim, Y.J.; Kim, B.S.; Sahay, P.P.; Lee, S.H. Nanocube In2 O3 @RGO heterostructure
based gas sensor for acetone and formaldehyde detection. RSC Adv. 2017, 7, 38714–38724. [CrossRef]
61. Schedin, F.; Geim, A.K.; Morozov, S.V.; Hill, E.W.; Blake, P.; Katsnelson, M.I.; Novoselov, K.S. Detection of individual gas molecules
adsorbed on graphene. Nat. Mater. 2007, 6, 652–655. [CrossRef] [PubMed]
62. Singh, E.; Meyyappan, M.; Nalwa, H.S. Flexible Graphene-Based Wearable Gas and Chemical Sensors. ACS Appl. Mater. Interfaces
2017, 9, 34544–34586. [CrossRef] [PubMed]
63. Basu, S.; Bhattacharyya, P. Recent developments on graphene and graphene oxide based solid state gas sensors. Sens. Actuators B
Chem. 2012, 173, 1–21. [CrossRef]
64. Varghese, S.; Lonkar, S.; Singh, K.; Swaminathan, S.; Abdala, A. Recent advances in graphene based gas sensors. Sens. Actuators B
Chem. 2015, 218, 160–183. [CrossRef]
65. Ricciardella, F.; Massera, E.; Polichetti, T.; Miglietta, M.L.; Di Francia, G. Calibrated Graphene-Based Chemi-Sensor for Sub
Parts-Per-Million NO2 Detection Operating at Room Temperature. Appl. Phys. Lett. 2014, 104, 183502. [CrossRef]
66. Rumyantsev, S.; Liu, G.X.; Shur, M.S.; Potyrailo, R.A.; Balandin, A.A. Selective Gas Sensing with a Single Pristine Graphene
Transistor. Nano Lett. 2012, 12, 2294–2298. [CrossRef]
67. Choi, H.; Choi, J.S.; Kim, J.S.; Choe, J.H.; Chung, K.H.; Shin, J.W.; Kim, J.T.; Youn, D.H.; Kim, K.C.; Lee, J.I.; et al. Flexible and
Transparent Gas Molecule Sensor Integrated with Sensing and Heating Graphene Layers. Small 2014, 10, 3685–3691. [CrossRef]
68. Nomani, M.W.K.; Shishir, R.; Qazi, M.; Diwan, D.; Shields, V.B.; Spencer, M.G.; Tompa, G.S.; Sbrockey, N.M.; Koley, G. Highly
sensitive and selective detection of NO2 using epitaxial graphene on 6H-SiC. Sens. Actuators B Chem. 2010, 150, 301–307.
[CrossRef]
69. Yang, G.; Kim, H.Y.; Jang, S.; Kim, J. Transfer-Free Growth of Multilayer Graphene Using Self-Assembled Monolayers. ACS Appl.
Mater. Interfaces 2016, 8, 27115–27121. [CrossRef]
70. Shen, F.P.; Wang, D.; Liu, R.; Pei, X.F.; Zhang, T.; Jin, J. Edge-tailored graphene oxide nanosheet-based field effect transistors for
fast and reversible electronic detection of sulfur dioxide. Nanoscale 2013, 5, 537–540. [CrossRef]
71. Ghosh, R.; Midya, A.; Santra, S.; Ray, S.K.; Guha, P.K. Chemically Reduced Graphene Oxide for Ammonia Detection at Room
Temperature. ACS Appl. Mater. Interfaces 2013, 5, 7599–7603. [CrossRef] [PubMed]
72. Hu, N.T.; Wang, Y.Y.; Chai, J.; Gao, R.G.; Yang, Z.; Kong, E.S.W.; Zhang, Y.F. Gas sensor based on p-phenylenediamine reduced
graphene oxide. Sens. Actuators B Chem. 2012, 163, 107–114. [CrossRef]
73. Dua, V.; Surwade, S.P.; Ammu, S.; Agnihotra, S.R.; Jain, S.; Roberts, K.E.; Park, S.; Ruoff, R.S.; Manohar, S.K. All-organic vapor
sensor using inkjet-printed reduced graphene oxide. Angew. Chem. Int. Ed. 2010, 49, 2154–2157. [CrossRef]
74. Baek, D.H.; Kim, J. MoS2 gas sensor functionalized by Pd for the detection of hydrogen. Sens. Actuators B Chem. 2017, 250,
686–691. [CrossRef]
75. Donarelli, M.; Prezioso, S.; Perrozzi, F.; Bisti, F.; Nardone, M.; Giancaterini, L.; Cantalini, C.; Ottaviano, L. Response to NO2 and
other gases of resistive chemically exfoliated MoS2 -based gas sensors. Sens. Actuators B Chem. 2015, 207, 602–613. [CrossRef]
76. Zhao, P.X.; Tang, Y.; Mao, J.; Chen, Y.X.; Song, H.; Wang, J.W.; Song, Y.; Liang, Y.Q.; Zhang, X.M. One-Dimensional MoS2 -Decorated
TiO2 nanotube gas sensors for efficient alcohol sensing. J. Alloys Compd. 2016, 674, 252–258. [CrossRef]
77. Li, H.; Yin, Z.Y.; He, Q.Y.; Li, H.; Huang, X.; Lu, G.; Fam, D.W.H.; Tok, A.I.Y.; Zhang, Q.; Zhang, H. Fabrication of Single- and
Multilayer MoS2 Film-Based Field-Effect Transistors for Sensing NO at Room Temperature. Small 2012, 8, 63–67. [CrossRef]
78. Late, D.J.; Huang, Y.K.; Liu, B.; Acharya, J.; Shirodkar, S.; Luo, J.; Yan, A.; Carles, D.; Waghmare, U.V.; Dravid, V.P.; et al. Sensing
Behavior of Atomically Thin-Layered MoS2 Transistors. ACS Nano 2013, 7, 4879–4891. [CrossRef]
79. Wang, J.; Lian, G.; Xu, Z.; Fu, C.; Lin, Z.; Li, L.; Wang, Q.; Cui, D.; Wong, C.P. Growth of Large-Size SnS Thin Crystals Driven
by Oriented Attachment and Applications to Gas Sensors and Photodetectors. ACS Appl. Mater. Interfaces 2016, 8, 9545–9551.
[CrossRef]
80. Xiong, Y.; Xu, W.; Ding, D.; Lu, W.; Zhu, L.; Zhu, Z.; Wang, Y.; Xue, Q. Ultra-sensitive NH3 sensor based on flower-shaped SnS2
nanostructures with sub-ppm detection ability. J. Hazard. Mater. 2018, 341, 159–167. [CrossRef]
Nanomaterials 2022, 12, 982 36 of 40
81. Wu, E.X.; Xie, Y.; Yuan, B.; Zhang, H.; Hu, X.D.; Liu, J.; Zhang, D.H. Ultrasensitive and Fully Reversible NO2 Gas Sensing Based
on p-Type MoTe2 under Ultraviolet Illumination. ACS Sens. 2018, 3, 1719–1726. [CrossRef] [PubMed]
82. Gu, D.; Wang, X.Y.; Liu, W.; Li, X.G.; Lin, S.W.; Wang, J.; Rumyantseva, M.N.; Gaskov, A.M.; Akbar, S.A. Visible-light activated
room temperature NO2 sensing of SnS2 nanosheets based chemiresistive sensors. Sens. Actuators B Chem. 2020, 305, 127455.
[CrossRef]
83. Cheng, M.; Wu, Z.P.; Liu, G.N.; Zhao, L.J.; Gao, Y.; Zhang, B.; Liu, F.M.; Yan, X.; Liang, X.S.; Sun, P.; et al. Highly sensitive sensors
based on quasi-2D rGO/SnS2 hybrid for rapid detection of NO2 gas. Sens. Actuators B 2019, 291, 216–225. [CrossRef]
84. Xu, S.P.; Sun, F.Q.; Gu, F.L.; Zuo, Y.B.; Zhang, L.H.; Fan, C.F.; Yang, S.M.; Li, W.S. Photochemistry-Based Method for the Fabrication
of SnO2 Monolayer Ordered Porous Films with Size-Tunable Surface Pores for Direct Application in Resistive-Type Gas Sensor.
ACS Appl. Mater. Interfaces 2014, 6, 1251–1257. [CrossRef]
85. Ji, F.X.; Ren, X.P.; Zheng, X.Y.; Liu, Y.C.; Pang, L.Q.; Jiang, J.X.; Liu, S.Z. 2D-MoO3 nanosheets for superior gas sensors. Nanoscale
2016, 8, 8696–8703. [CrossRef]
86. Boudiba, A.; Zhang, C.; Bittencourt, C.; Umek, P.; Olivier, M.G.; Snyders, R.; Debliquy, M. Hydrothermal Synthesis of Two
Dimensional WO3 Nanostructures for NO2 Detection in the ppb-level. Proc. Eng. 2012, 47, 228–231. [CrossRef]
87. Cho, Y.H.; Ko, Y.N.; Kang, Y.C.; Kim, I.D.; Lee, J.H. Ultraselective and ultrasensitive detection of trimethylamine using MoO3
nanoplates prepared by ultrasonic spray pyrolysis. Sens. Actuators B Chem. 2014, 195, 189–196. [CrossRef]
88. Wang, M.S.; Wang, Y.W.; Li, X.J.; Ge, C.X.; Hussain, S.; Liu, G.W.; Qiao, G.J. WO3 porous nanosheet arrays with enhanced low
temperature NO2 gas sensing performance. Sens. Actuators B Chem. 2020, 316, 128050. [CrossRef]
89. Kaneti, Y.V.; Yue, J.; Jiang, X.C.; Yu, A.B. Controllable Synthesis of ZnO Nanoflakes with Exposed (10(1)over-bar0) for Enhanced
Gas Sensing Performance. J. Phys. Chem. C 2013, 117, 13153–13162. [CrossRef]
90. Wang, X.; Su, J.; Chen, H.; Li, G.D.; Shi, Z.F.; Zou, H.F.; Zou, X.X. Ultrathin In2 O3 Nanosheets with Uniform Mesopores for Highly
Sensitive Nitric Oxide Detection. ACS Appl. Mater. Interfaces 2017, 9, 16335–16342. [CrossRef]
91. Li, Z.J.; Lin, Z.J.; Wang, N.N.; Wang, J.Q.; Liu, W.; Sun, K.; Fu, Y.Q.; Wang, Z.G. High precision NH3 sensing using network
nano-sheet Co3 O4 arrays based sensor at room temperature. Sens. Actuators B Chem. 2016, 235, 222–231. [CrossRef]
92. Chen, X.W.; Wang, S.; Su, C.; Han, Y.T.; Zou, C.; Zeng, M.; Hu, N.T.; Su, Y.J.; Zhou, Z.H.; Yang, Z. Two-dimensional Cd-doped
porous Co3 O4 nanosheets for enhanced roomtemperature NO2 sensing performance. Sens. Actuators B. Chem. 2020, 305, 127393.
[CrossRef]
93. Lin, H.; Gao, S.S.; Dai, C.; Chen, Y.; Shi, J.L. A Two-Dimensional Biodegradable Niobium Carbide (MXene) for Photothermal
Tumor Eradication in NIR-I and NIR-II Biowindows. J. Am. Chem. Soc. 2017, 139, 16235–16247. [CrossRef] [PubMed]
94. He, T.T.; Liu, W.; Lv, T.; Ma, M.S.; Liu, Z.F.; Vasiliev, A.; Li, X.G. MXene/SnO2 heterojunction based chemical gas sensors. Sens.
Actuators B Chem. 2021, 329, 129275. [CrossRef]
95. Hermawan, A.; Zhang, B.; Taufik, A.; Asakura, Y.; Hasegawa, T.; Zhu, J.F.; Shi, P.; Yin, S. CuO Nanoparticles/Ti3 C2 Tx MXene
Hybrid Nanocomposites for Detection of Toluene Gas. ACS Appl. Nano Mater. 2020, 3, 4755–4766. [CrossRef]
96. Lee, E.; Kim, D.J. Review-Recent Exploration of Two-Dimensional MXenes for Gas Sensing: From a Theoretical to an Experimental
View. J. Electrochem. Soc. 2020, 167, 037515. [CrossRef]
97. Aghaei, S.M.; Aasi, A.; Panchapakesan, B. Experimental and Theoretical Advances in MXene-Based Gas Sensors. ACS Omega
2021, 6, 2450–2461. [CrossRef]
98. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-dimensional
nanocrystals produced by exfoliation of Ti3 AlC2 . Adv. Mater. 2011, 23, 4248–4253. [CrossRef]
99. Lee, E.; Mohammadi, A.V.; Prorok, B.C.; Yoon, Y.S.; Beidaghi, M.; Kim, D.J. Room Temperature Gas Sensing of Two-Dimensional
Titanium Carbide (MXene). ACS Appl. Mater. Interfaces 2017, 9, 37184–37190. [CrossRef]
100. Chae, Y.; Kim, S.J.; Cho, S.Y.; Choi, J.; Maleski, K.; Lee, B.J.; Jung, H.T.; Gogotsi, Y.; Lee, Y.; Ahn, C.W. An investigation into the
factors governing the oxidation of two-dimensional Ti3 C2 MXene. Nanoscale 2019, 11, 8387–8393. [CrossRef]
101. Yang, Z.J.; Liu, A.; Wang, C.L.; Liu, F.M.; He, J.M.; Li, S.Q.; Wang, J.; You, R.; Yan, X.; Sun, P.; et al. Improvement of Gas and
Humidity Sensing Properties of Organ-like MXene by Alkaline Treatment. ACS Sens. 2019, 4, 1261–1269. [CrossRef] [PubMed]
102. Lee, E.; VahidMohammadi, A.; Yoon, Y.S.; Beidaghi, M.; Kim, D.J. Two-Dimensional Vanadium Carbide MXene for Gas Sensors
with Ultrahigh Sensitivity Toward Nonpolar Gases. ACS Sens. 2019, 4, 1603–1611. [CrossRef] [PubMed]
103. Cho, S.Y.; Kim, J.Y.; Kwon, O.; Kim, J.; Jung, H.T. Molybdenum carbide chemical sensors with ultrahigh signal-to-noise ratios and
ambient stability. J. Mater. Chem. A 2018, 6, 23408–23416. [CrossRef]
104. Wu, T.; Wang, Z.B.; Tian, M.H.; Miao, J.Y.; Zhang, H.X.; Sun, J.B. UV Excitation NO2 Gas Sensor Sensitized by ZnO Quantum
Dots at Room Temperature. Sens. Actuators B Chem. 2018, 259, 526–531. [CrossRef]
105. Lee, J.; Jung, Y.; Sung, S.H.; Lee, G.; Kim, J.; Seong, J.; Shim, Y.S.; Jun, S.C.; Jeon, S. High-performance gas sensor array for indoor
air quality monitoring: The role of Au nanoparticles on WO3 , SnO2, and NiO-based gas sensors. J. Mater. Chem. A 2021, 9,
1159–1167. [CrossRef]
106. Pan, F.J.; Lin, H.; Zhai, H.Z.; Miao, Z.; Zhang, Y.; Xu, K.L.; Guan, B.; Huang, H.; Zhang, H. Pd-doped TiO2 Film Sensors Prepared
by Premixed Stagnation Flames for CO and NH3 Gas Sensing. Sens. Actuators B Chem. 2018, 261, 451–459. [CrossRef]
107. Zhao, Y.; Zhang, J.; Wang, Y.; Chen, Z. A highly sensitive and room temperature CNTs/SnO2 /CuO sensor for H2 S gas sensing
applications. Nanoscale Res. Lett. 2020, 15, 40. [CrossRef]
Nanomaterials 2022, 12, 982 37 of 40
108. Schipani, F.; Aldao, C.M.; Ponce, M.A. Schottky barriers measurements through Arrhenius plots in gas sensors based on
semiconductor films. AIP Adv. 2012, 2, 032138. [CrossRef]
109. Zhou, Q.; Chen, W.G.; Xu, L.N.; Kumar, R.; Gui, Y.G.; Zhao, Z.Y.; Tang, C.; Zhu, S.P. Highly sensitive carbon monoxide (CO) gas
sensors based on Ni and Zn doped SnO2 nanomaterials. Ceram. Int. 2018, 44, 4392–4399. [CrossRef]
110. Li, J.; Xian, J.B.; Wang, W.J.; Cheng, K.; Zeng, M.; Zhang, A.H.; Wu, S.J.; Gao, X.S.; Lu, X.B.; Liu, J.M. Ultrafast response and
high-sensitivity acetone gas sensor based on porous hollow Ru-doped SnO2 nanotubes. Sens. Actuators B Chem. 2022, 352, 131061.
[CrossRef]
111. Wang, T.T.; Wang, Y.; Sun, Q.; Zheng, S.L.; Liu, L.Z.; Li, J.L.; Hao, J.Y. Boosted interfacial charge transfer in SnO2 /SnSe2
heterostructures: Toward ultrasensitive room-temperature H2 S detection. Inorg. Chem. Front. 2021, 8, 2068–2077. [CrossRef]
112. Zhong, X.; Shen, Y.; Zhao, S.; Li, T.; Lu, R.; Yin, Y.; Han, C.; Wei, D.; Zhang, Y.; Wei, K. Effect of pore structure of the metakaolin-
based porous substrate on the growth of SnO2 nanowires and their H2 S sensing properties. Vacuum 2019, 167, 118–128. [CrossRef]
113. Hu, D.; Han, B.Q.; Han, R.; Deng, S.J.; Wang, Y.; Li, Q.; Wang, Y.D. SnO2 nanorods based sensing material as an isopropanol
vapor sensor. New J. Chem. 2014, 38, 2443–2450. [CrossRef]
114. Xu, R.; Zhang, L.X.; Li, M.W.; Yin, Y.Y.; Yin, J.; Zhu, M.Y.; Chen, J.J.; Wang, Y.; Bie, L.J. Ultrathin SnO2 nanosheets with dominant
high-energy {001} facets for low temperature formaldehyde gas sensor. Sens. Actuators B Chem. 2019, 289, 186–194. [CrossRef]
115. Ren, H.B.; Zhao, W.; Wang, L.Y.; Ryu, S.O.; Gu, C.P. Preparation of porous flower-like SnO2 micro/nano structures and their
enhanced gas sensing property. J. Alloy Compd. 2015, 653, 611–618. [CrossRef]
116. Sun, H.M.; Zhang, C.; Peng, Y.J.; Gao, W. Synthesis of double-shelled SnO2 hollow cubes for superior isopropanol sensing
performance. New J. Chem. 2019, 43, 4721–4726. [CrossRef]
117. Lee, J.H. Gas sensors using hierarchical and hollow oxide nanostructures: Overview. Sens. Actuators B Chem. 2009, 140, 319–336.
[CrossRef]
118. Zhang, Y.Q.; Li, D.; Qin, L.G.; Zhao, P.L.; Liu, F.M.; Chuai, X.H.; Sun, P.; Liang, X.S.; Gao, Y.; Sun, Y.F.; et al. Preparation and gas
sensing properties of hierarchical leaf-like SnO2 materials. Sens. Actuators B Chem. 2018, 255, 2944–2951. [CrossRef]
119. Feng, X.Y.; Jiang, J.; Ding, H.; Ding, R.M.; Luo, D.; Zhu, J.H.; Feng, Y.M.; Huang, X.T. Carbon-assisted synthesis of mesoporous
SnO2 nanomaterial as highly sensitive ethanol gas sensor. Sens. Actuators B Chem. 2013, 183, 526–534. [CrossRef]
120. Dong, K.Y.; Choi, J.K.; Hwang, I.S.; Lee, J.W.; Kang, B.H.; Ham, D.J.; Lee, J.H.; Ju, B.K. Enhanced H2 S sensing characteristics of Pt
doped SnO2 nanofibers sensors with micro heater. Sens. Actuators B Chem. 2011, 157, 154–161. [CrossRef]
121. Chen, Y.P.; Qin, H.W.; Hu, J.F. CO sensing properties and mechanism of Pd doped SnO2 thick-films. Appl. Surf. Sci. 2018, 428,
207–217. [CrossRef]
122. Lee, Y.C.; Huang, H.; Tan, O.K.; Tse, M.S. Semiconductor gas sensor based on Pd-doped SnO2 nanorod thin films. Sens. Actuators
B Chem. 2008, 135, 239–242. [CrossRef]
123. Shen, Y.B.; Yamazaki, T.; Liu, Z.F.; Meng, D.; Kikuta, T.; Nakatani, N.; Saito, M.; Mori, M. Microstructure and H-2 gas sensing
properties of undoped and Pd-doped SnO2 nanowires. Sens. Actuators B Chem. 2009, 135, 524–529. [CrossRef]
124. Ramgir, N.S.; Hwang, Y.K.; Jhung, S.H.; Kim, H.K.; Hwang, J.S.; Mulla, I.S.; Chang, J.S. CO sensor derived from mesostructured
Au-doped SnO2 thin film. Appl. Surf. Sci. 2006, 252, 4298–4305. [CrossRef]
125. Zhao, C.H.; Gong, H.M.; Niu, G.Q.; Wang, F. Ultrasensitive SO2 sensor for sub-ppm detection using Cu-doped SnO2 nanosheet
arrays directly grown on chip. Sens. Actuators B Chem. 2020, 324, 128745. [CrossRef]
126. Chen, Y.; Zhu, C.L.; Shi, X.L.; Cao, M.S.; Jin, H.B. The synthesis and selective gas sensing characteristics of SnO2 /alpha-Fe2 O3
hierarchical nanostructures. Nanotechnology 2008, 19, 205603. [CrossRef]
127. Xue, X.Y.; Xing, L.L.; Chen, Y.J.; Shi, S.L.; Wang, Y.G.; Wang, T.H. Synthesis and H2 S sensing properties of CuO-SnO2 core/shell
PN-junction nanorods. J. Phys. Chem. C 2008, 112, 12157–12160. [CrossRef]
128. Haoyuan, F.Y.L.J.X. SnO2 recycled fromtin slime for enhanced SO2 sensing properties by NiO surface decoration. Mater. Sci.
Semicond. Process. 2020, 114, 105073.
129. Yuliarto, B.; Ramadhani, M.F.; Nugraha; Septiani, N.L.W.; Hamam, K.A. Enhancement of SO2 gas sensing performance using
ZnO nanorod thin films: The role of deposition time. J. Mater. Sci. 2017, 52, 4543–4554. [CrossRef]
130. Wang, H.T.; Dai, M.; Li, Y.Y.; Bai, J.H.; Liu, Y.Y.; Li, Y.; Wang, C.C.; Liu, F.M.; Lu, G.Y. The influence of different ZnO nanostructures
on NO2 sensing performance. Sens. Actuators B Chem. 2021, 329, 129145. [CrossRef]
131. Liu, Y.X.; Hang, T.; Xie, Y.Z.; Bao, Z.; Song, J.; Zhang, H.L.; Xie, E.Q. Effect of Mg doping on the hydrogen-sensing characteristics
of ZnO thin films. Sens. Actuators B Chem. 2011, 160, 266–270. [CrossRef]
132. Hassan, H.S.; Kashyout, A.B.; Soliman HM, A.; Uosif, M.A.; Afify, N. Afify Effect of reaction time and Sb doping ratios on the
architecturing of ZnO nanomaterials for gas sensor applications. Appl. Surf. Sci. 2013, 277, 73–82. [CrossRef]
133. Chaitra, U.; Ali, A.V.M.; Viegas, A.E.; Kekud, D.; Rao, K.M. Growth and characterization of undoped and aluminium doped zinc
oxide thin films for SO2 gas sensing below threshold value limit. Appl. Surf. Sci. 2019, 496, 143724. [CrossRef]
134. Kolhe, P.S.; Shinde, A.B.; Kulkarni, S.G.; Maiti, N.; Koinkar, P.M.; Sonawane, K.M. Gas sensing performance of Al doped ZnO thin
film for H2 S detection. J. Alloys Compd. 2018, 748, 6–11. [CrossRef]
135. Xiang, Q.; Meng, G.F.; Zhang, Y.; Xu, J.Q.; Xu, P.C.; Pan, Q.Y.; Yu, W.J. Ag nanoparticle embedded-ZnO nanorods synthesized via
a photochemical method and its gas-sensing properties. Sens. Actuators B Chem. 2010, 143, 635–640. [CrossRef]
136. Kim, J.H.; Katoch, A.; Kim, S.S. Optimum shell thickness and underlying sensing mechanism in p-n CuO-ZnO core-shell
nanowires. Sens. Actuators B Chem. 2016, 222, 249–256. [CrossRef]
Nanomaterials 2022, 12, 982 38 of 40
137. Zhou, Q.; Zeng, W.; Chen, W.G.; Xu, L.N.; Kumar, R.; Umar, A. High sensitive and low concentration sulfur dioxide (SO2 ) gas
sensor application of heterostructure NiO-ZnO nanodisks. Sens. Actuators B Chem. 2019, 298, 126870. [CrossRef]
138. Li, Z.J.; Wang, N.N.; Lin, Z.J.; Wang, J.Q.; Liu, W.; Sun, K.; Fu, Y.Q.; Wang, Z.G. Room-Temperature High-Performance H2 S
Sensor Based on Porous CuO Nanosheets Prepared by Hydrothermal Method. ACS Appl. Mater. Interfaces 2016, 8, 20962–20968.
[CrossRef]
139. Navale, Y.H.; Navale, S.T.; Galluzzi, M.; Stadler, F.J.; Debnath, A.K.; Ramgir, N.S.; Gadkari, S.C.; Gupta, S.K.; Aswal, D.K.;
Patil, V.B. Rapid synthesis strategy of CuO nanocubes for sensitive and selective detection of NO2 . J. Alloys Compd. 2017, 708,
456–463. [CrossRef]
140. Huang, X.; Ren, Z.B.; Zheng, X.H.; Tang, D.P.; Wu, X.; Lin, C. A facile route to batch synthesis CuO hollow microspheres with
excellent gas sensing properties. J. Mater. Sci. Mater. Electron. 2018, 29, 5969–5974. [CrossRef]
141. Hu, Q.; Zhang, W.J.; Wang, X.Y.; Wang, Q.; Huang, B.Y.; Li, Y.; Hua, X.H.; Liu, G.; Li, B.S.; Zhou, J.Y.; et al. Binder-free CuO
nanoneedle arrays based tube-type sensor for H2 S gas sensing. Sens. Actuators B Chem. 2021, 326, 128993. [CrossRef]
142. Hou, L.; Zhang, C.M.; Li, L.; Du, C.; Li, X.K.; Kang, X.F.; Chen, W. CO gas sensors based on p-type CuO nanotubes and CuO
nanocubes: Morphology and surface structure effects on the sensing performance. Talanta 2018, 188, 41–49. [CrossRef] [PubMed]
143. Hu, X.B.; Zhu, Z.G.; Chen, C.; Wen, T.Y.; Zhao, X.L.; Xie, L.L. Highly sensitive H2 S gas sensors based on Pd-doped CuO
nanoflowers with low operating temperature. Sens. Actuators B Chem. 2017, 253, 809–817. [CrossRef]
144. Tang, Q.; Hu, X.-B.; He, M.; Xie, L.-L.; Zhu, Z.-G.; Wu, J.-Q. Effect of Platinum Doping on the Morphology and Sensing
Performance for CuO-Based Gas Sensor. Appl. Sci. 2018, 8, 1091. [CrossRef]
145. Mnethu, O.; Nkosi, S.S.; Kortidis, I.; Motaung, D.E.; Kroon, R.E.; Swart, H.C.; Ntsasa, N.G.; Tshilongo, J.; Moyo, T. Ultra-sensitive
and selective p-xylene gas sensor at low operating temperature utilizing Zn doped CuO nanoplatelets: Insignificant vestiges of
oxygen vacancies. J. Colloid Interface Sci. 2020, 576, 364–375. [CrossRef]
146. Bhuvaneshwari, S.; Gopalakrishnan, N. Enhanced ammonia sensing characteristics of Cr doped CuO nanoboats. J. Alloys Compd.
2016, 654, 202–208. [CrossRef]
147. Molavi, R.; Sheikhi, M.H. Facile wet chemical synthesis of Al doped CuO nanoleaves for carbon monoxide gas sensor applications.
Mater. Sci. Semicond. Process. 2020, 106, 104767. [CrossRef]
148. Zhang, H.; Li, H.R.; Cai, L.N.; Lei, Q.; Wang, J.N.; Fan, W.H.; Shi, K.; Han, G.L. Performances of In-doped CuO-based
heterojunction gas sensor. J. Mater. Sci. Mater. Electron. 2020, 31, 910–919. [CrossRef]
149. Wang, Z.F.; Li, F.; Wang, H.T.; Wang, A.; Wu, S.M. An enhanced ultra-fast responding ethanol gas sensor based on Ag functional-
ized CuO nanoribbons at room-temperature. J. Mater. Sci. Mater. Electron. 2018, 29, 16654–16659. [CrossRef]
150. Sui, L.L.; Yu, T.T.; Zhao, D.; Cheng, X.L.; Zhang, X.F.; Wang, P.; Xu, Y.M.; Gao, S.; Zhao, H.; Gao, Y.; et al. In situ deposited
hierarchical CuO/NiO nanowall arrays film sensor with enhanced gas sensing performance to H2 S. J. Hazard. Mater. 2020,
385, 121570. [CrossRef]
151. Park, K.R.; Cho, H.B.; Lee, J.; Song, Y.; Kim, W.B.; Choa, Y.H. Design of highly porous SnO2 -CuO nanotubes for enhancing H2 S
gas sensor performance. Sens. Actuators B Chem. 2020, 302, 127179. [CrossRef]
152. Li, T.; Sun, M.; Wu, S. State-of-the-Art Review of Electrospun Gelatin-Based Nanofiber Dressings for Wound Healing Applications.
Nanomaterials 2022, 12, 784. [CrossRef] [PubMed]
153. Sheng, X.; Li, T.; Sun, M.; Liu, G.; Zhang, Q.; Ling, Z.; Gao, S.; Diao, F.; Zhang, J.; Rosei, F.; et al. Flexible electrospun iron
compounds/carbon fibers: Phase transformation and electrochemical properties. Electrochim. Acta 2022, 407, 139892. [CrossRef]
154. Liang, X.; Kim, T.H.; Yoon, J.W.; Kwak, C.H.; Lee, J.H. Ultrasensitive and ultraselective detection of H2 S using electrospun
CuO-loaded In2 O3 nanofiber sensors assisted by pulse heating. Sens. Actuators B Chem. 2015, 209, 934–942. [CrossRef]
155. Barsan, N.; Koziej, D.; Weimar, U. Metal oxide-based gas sensor research: How to? Sens. Actuators B Chem. 2007, 121, 18–35.
[CrossRef]
156. Upadhyay, S.B.; Mishra, R.K.; Sahay, P.P. Enhanced acetone response in co-precipitated WO3 nanostructures upon indium doping.
Sens. Actuators B Chem. 2015, 209, 368–376. [CrossRef]
157. Hu, Q.; Chang, J.; Gao, J.; Huang, J.; Feng, L. Needle-shaped WO3 nanorods for triethylamine gas sensing. ACS Appl. Nano Mater.
2020, 3, 9046–9054. [CrossRef]
158. Li, Z.H.; Li, J.C.; Song, L.L.; Gong, H.Q.; Niu, Q. Ionic liquid-assisted synthesis of WO3 particles with enhanced gas sensing
properties. J. Mater. Chem. A 2013, 1, 15377–15382. [CrossRef]
159. Li, H.; Liu, B.; Cai, D.P.; Wang, Y.R.; Liu, Y.; Mei, L.; Wang, L.L.; Wang, D.D.; Li, Q.H.; Wang, T.H. High-temperature humidity
sensors based on WO3 -SnO2 composite hollow nanospheres. J. Mater. Chem. A 2014, 2, 6854–6862. [CrossRef]
160. Thu, N.T.A.; Cuong, N.D.; Nguyen, L.C.; Khieu, D.Q.; Nam, P.C.; Van Toan, N.; Hung, C.M.; Van Hieu, N. Fe2 O3 nanoporous
network fabricated from Fe3 O4 /reduced graphene oxide for high-performance ethanol gas sensor. Sens. Actuators B Chem. 2018,
255, 3275–3283. [CrossRef]
161. Li, Z.J.; Huang, Y.W.; Zhang, S.C.; Chen, W.M.; Kuang, Z.; Ao, D.Y.; Liu, W.; Fu, Y.Q. A fast response & recovery H2 S gas sensor
based on α-Fe2 O3 nanoparticles with ppb level detection limit. J. Hazard. Mater. 2015, 300, 167–174. [PubMed]
162. Liang, S.; Li, J.P.; Wang, F.; Qin, J.L.; Lai, X.Y.; Jiang, X.M. Highly sensitive acetone gas sensor based on ultrafine α-Fe2 O3
nanoparticles. Sens. Actuators B Chem. 2017, 238, 923–927. [CrossRef]
163. Shoorangiz, M.; Shariatifard, L.; Roshan, H.; Mirzaei, A. Selective ethanol sensor based on alpha-Fe2O3 nanoparticles. Inorg.
Chem. Commun. 2021, 133, 108961. [CrossRef]
Nanomaterials 2022, 12, 982 39 of 40
164. Qu, F.D.; Zhou, X.X.; Zhang, B.X.; Zhang, S.D.; Jiang, C.J.; Ruan, S.P.; Yang, M.H. Fe2 O3 nanoparticles-decorated MoO3 nanobelts
for enhanced chemiresistive gas sensing. J. Alloys Compd. 2019, 782, 672–678. [CrossRef]
165. Navale, S.T.; Liu, C.S.T.; Gaikar, P.S.; Patil, V.B.; Sagar, R.U.R.; Du, B.; Mane, R.S.; Stadler, F.J. Solution-processed rapid synthesis
strategy of Co3 O4 for the sensitive and selective detection of H2 S. Sens. Actuators B Chem. 2017, 245, 524–532. [CrossRef]
166. Li, W.Y.; Xu, L.N.; Chen, J. Co3 O4 nanomaterials in lithium-ion batteries and gassensors. Adv. Funct. Mater. 2005, 15, 851–857.
[CrossRef]
167. Sun, C.W.; Rajasekhara, S.; Chen, Y.J.; Goodenough, J.B. Facile synthesis of monodisperse porous Co3 O4 microspheres with
superior ethanol sensing properties. Chem. Commun. 2011, 47, 12852–12854. [CrossRef]
168. Zhang, R.; Gao, S.; Zhou, T.T.; Tu, J.C.; Zhang, T. Facile preparation of hierarchical structure based on p-type Co3 O4 as toluene
detecting sensor. Appl. Surf. Sci. 2020, 503, 144167. [CrossRef]
169. Chen, X.F.; Huang, Y.; Zhang, K.C.; Feng, X.S.; Wang, M.Y. Porous TiO2 nanobelts coated with mixed transition-metal oxides
Sn3 O4 nanosheets core-shell composites as high-performance anode materials of lithium ion batteries. Electrochim. Acta Mater.
2018, 259, 131–142. [CrossRef]
170. Chen, G.; Ji, S.; Li, H.; Kang, X.; Chang, S.; Wang, Y.; Yu, G.; Lu, J.; Claverie, J.; Sang, Y.; et al. High-energy faceted SnO2 -coated
TiO2 nanobelt heterostructure for near-ambient temperature-responsive ethanol sensor. ACS Appl. Mater. Interfaces 2015, 7,
24950–24956. [CrossRef]
171. Wang, X.H.; Sang, Y.H.; Wang, D.Z.; Ji, S.Z.; Liu, H. Enhanced gas sensing property of SnO2 nanoparticles by constructing the
SnO2 -TiO2 nanobelt heterostructure. J. Alloys Compd. 2015, 639, 571–576. [CrossRef]
172. Chen, N.; Li, X.G.; Wang, X.Y.; Yu, J.; Wang, J.; Tang, Z.A.; Akbar, S.A. Enhanced room temperature sensing of Co3 O4 -intercalated
reduced graphene oxide based gas sensors. Sens. Actuators B Chem. 2013, 188, 902–908. [CrossRef]
173. Chen, Y.; Zhang, W.; Wu, Q.S. A highly sensitive room-temperature sensing material for NH3 : SnO2 -nanorods coupled by rGO.
Sens. Actuators B Chem. 2017, 242, 1216–1226. [CrossRef]
174. Liu, J.; Li, S.; Zhang, B.; Wang, Y.L.; Gao, Y.; Liang, X.S.; Wang, Y.; Lu, G.Y. Flower-like In2 O3 modified by reduced graphene
oxide sheets serving as a highly sensitive gas sensor for trace NO2 detection. J. Colloid Interface Sci. 2017, 504, 206–213. [CrossRef]
[PubMed]
175. Pienutsa, N.; Roongruangsree, P.; Seedokbuab, V.; Yannawibut, K.; Phatoomvijitwong, C.; Srinives, S. SnO2 -graphene composite
gas sensor for a room temperature detection of ethanol. Nanotechnology 2021, 32, 115502. [CrossRef]
176. Kim, H.W.; Na, H.G.; Kwon, Y.J.; Kang, S.Y.; Choi, M.S.; Bang, J.H.; Wu, P.; Kim, S.S. Microwave-Assisted Synthesis of Graphene-
SnO2 Nanocomposites and Their Applications in Gas Sensors. ACS Appl. Mater. Interfaces 2017, 9, 31667–31682. [CrossRef]
177. Russo, P.A.; Donato, N.; Leonardi, S.G.; Baek, S.; Conte, D.E.; Neri, G.; Pinna, N. Room-Temperature Hydrogen Sensing with
Heteronanostructures Based on Reduced Graphene Oxide and Tin Oxide. Angew. Chem. Int. Ed. 2012, 51, 11053–11057. [CrossRef]
178. Zhang, D.Z.; Liu, J.J.; Jiang, C.X.; Li, P.; Sun, Y.E. High-performance sulfur dioxide sensing properties of layer-by-layer self-
assembled titania-modified graphene hybrid nanocomposite. Sens. Actuators B Chem. 2017, 245, 560–567. [CrossRef]
179. Li, J.; Liu, X.; Sun, J.B. One step solvothermal synthesis of urchin-like ZnO nanorods/graphene hollow spheres and their NO2 ,
gas sensing properties. Ceram. lnt. 2016, 42, 2085–2090. [CrossRef]
180. Liu, S.; Yu, B.; Zhang, H.; Fei, T.; Zhang, T. Enhancing NO2 gas sensing performances at room temperature based on reduced
graphene oxide-ZnO nanoparticles hybrids. Sens. Actuators B Chem. 2014, 202, 272–278. [CrossRef]
181. Sun, Z.; Huang, D.; Yang, Z.; Li, X.L.; Hu, N.T.; Yang, C.; Wei, H.; Yin, G.L.; He, D.N.; Zhang, Y.F. ZnO Nanowire-Reduced
Graphene Oxide Hybrid Based Portable NH3 Gas Sensing Electron Device. IEEE Electron Dev. Lett. 2015, 36, 1376–1379. [CrossRef]
182. Wang, C. Reduced graphene oxide decorated with CuO-ZnO hetero-junctions: Towards high selective gas-sensing property to
acetone. J. Mater. Chem. A 2014, 2, 18635–18643. [CrossRef]
183. Wang, D.; Zhang, M.L.; Chen, Z.L.; Li, H.J.; Chen, A.Y.; Wang, X.Y.; Yang, J.H. Enhanced formaldehyde sensing properties of
hollow SnO2 nanofibers by graphene oxide. Sens. Actuators B Chem. 2017, 250, 533–542. [CrossRef]
184. Wang, Z.; Hang, T.; Fei, T.; Liu, S.; Zhang, T. Investigation of microstructure effect on NO2 sensors based on SnO2 nanoparti-
cles/reduced graphene oxide hybrids. ACS Appl. Mater. Interfaces 2018, 10, 41773–41783. [CrossRef] [PubMed]
185. Yin, L.; Chen, D.L.; Cui, X.; Ge, L.F.; Yang, J.; Yu, L.T.; Zhang, B.; Zhang, R.; Shao, G.S. Normal-pressure microwave rapid synthesis
of hierarchical SnO2 @rGO nanostructures with superhigh surface areas as high-quality gas-sensing and electrochemical active
materials. Nanoscale 2014, 6, 13690–13700. [CrossRef]
186. Tyagi, P.; Sharma, A.; Tomar, M.; Gupta, V. A comparative study of RGO-SnO2 and MWCNT-SnO2 nanocomposites based SO2
gas sensors. Sens. Actuators B Chem. 2017, 248, 980–986. [CrossRef]
187. Yu, X.J.; Cheng, C.D.; Feng, S.P.; Jia, X.H.; Song, H.J. Porous α-Fe2 O3 nanorods@graphite nanocomposites with improved high
temperature gas sensitive properties. J. Alloys Compd. 2019, 784, 1261–1269. [CrossRef]
188. Zhou, T.T.; Zhang, T.; Deng, J.A.; Zhang, R.; Lou, Z.; Wang, L.L. P-type Co3 O4 nanomaterials-based gas sensor: Preparation and
acetone sensing performance. Sens. Actuators B Chem. 2017, 242, 369–377. [CrossRef]
189. Zhang, B.; Cheng, M.; Liu, G.N.; Gao, Y.; Zhao, L.J.; Li, S.; Wang, Y.P.; Liu, F.M.; Liang, X.S.; Zhang, T.; et al. Room temperature
NO2 gas sensor based on porous Co3 O4 slices/reduced graphene oxide hybrid. Sens. Actuators B Chem. 2018, 263, 387–399.
[CrossRef]
190. Srirattanapibul, S.; Nakarungsee, P.; Issro, C.; Tang, I.M.; Thongmee, S. Enhanced room temperature NH3 sensing of rGO/Co3 O4
nanocomposites. Mater. Chem. Phys. 2021, 272, 125033. [CrossRef]
Nanomaterials 2022, 12, 982 40 of 40
191. Hao, Q.; Liu, T.; Liu, J.Y.; Liu, Q.; Jing, X.Y.; Zhang, H.Q.; Huang, G.Q.; Wang, J. Controllable synthesis and enhanced gas sensing
properties of a single-crystalline WO3 -rGO porous nanocomposite. RSC Adv. 2017, 7, 14192. [CrossRef]
192. Ye, Z.B.; Tai, H.L.; Xie, T.; Su, Y.J.; Yuan, Z.; Liu, C.H.; Jiang, Y.D. A facile method to develop novel TiO2 /rGO layered film sensor
for detecting ammonia at room temperature. Mater. Lett. 2016, 165, 127. [CrossRef]
193. Qao, X.Q.; Zhang, Z.W.; Hou, D.F.; Li, D.S.; Liu, Y.L.; Lan, Y.Q.; Zhang, J.; Feng, P.Y.; Bu, X.H. Tunable MoS2 /SnO2 P–N
Heterojunctions for an Efficient Trimethylamine Gas Sensor and 4-Nitrophenol Reduction Catalyst. ACS Sustainable Chem. Eng.
2018, 6, 12375–12384. [CrossRef]
194. Cui, S.M.; Wen, Z.H.; Huang, X.K.; Chang, J.B.; Chen, J.H. Stabilizing MoS2 Nanosheets through SnO2 Nanocrystal Decoration for
High-Performance Gas Sensing in Air. Small 2015, 11, 2305–2313. [CrossRef]
195. Yan, H.H.; Song, P.; Zhang, S.; Yang, Z.X.; Wang, Q. Facile synthesis, characterization and gas sensing performance of ZnO
nanoparticles-coated MoS2 nanosheets. J. Alloys Compd. 2016, 662, 118–125. [CrossRef]
196. Han, Y.T.; Huang, D.; Ma, Y.J.; He, G.L.; Hu, J.; Zhang, J.; Hu, N.T.; Su, Y.J.; Zhou, Z.H.; Zhang, Y.F.; et al. Design of Hetero-
Nanostructures on MoS2 Nanosheets To Boost NO2 Room-Temperature Sensing. ACS Appl. Mater. Interfaces 2018, 10, 22640–22649.
[CrossRef]
197. Zhang, D.Z.; Wu, J.F.; Cao, Y.H. Ultrasensitive H2 S gas detection at room temperature based on copper oxide/molybdenum
disulfide nanocomposite with synergistic effect. Sens. Actuators B Chem. 2019, 287, 346–355. [CrossRef]
198. Ikram, M.; Liu, L.J.; Liu, Y.; Ullah, M.; Ma, L.F.; Bakhtiar, S.U.; Wu, H.Y.; Yu, H.T.; Wang, R.H.; Shi, K.Y. Controllable synthesis of
MoS2 @MoO2 nanonetworks for enhanced NO2 room temperature sensing in air. Nanoscale 2019, 11, 8554–8564. [CrossRef]
199. Qin, Z.Y.; Ouyang, C.; Zhang, J.; Wan, L.; Wang, S.M.; Xie, C.S.; Zeng, D.W. 2D WS2 nanosheets with TiO2 quantum dots
decoration for high-performance ammonia gas sensing at room temperature. Sens. Actuators B Chem. 2017, 253, 1034. [CrossRef]
200. Gu, D.; Li, X.; Zhao, Y.; Wang, J. Enhanced NO2 sensing of SnO2 /SnS2 heterojunction based sensor. Sens. Actuators B Chem. 2017,
244, 67. [CrossRef]
201. Zhang, L.; Liu, Z.L.; Jin, L.; Zhang, B.B.; Zhang, H.T.; Zhu, M.H.; Yang, W.Q. Self-assembly gridding alpha-MoO3 nanobelts for
highly toxic H2S gas sensors. Sens. Actuators B Chem. 2016, 237, 350–357. [CrossRef]
202. Bai, S.L.; Chen, C.; Zhang, D.F.; Luo, R.X.; Li, D.Q.; Chen, A.F.; Liu, C.C. Intrinsic characteristic and mechanism in enhancing H2 S
sensing of Cd-doped α-MoO3 nanobelts. Sens. Actuators B Chem. 2014, 204, 754–762. [CrossRef]
203. Gao, X.M.; Ouyang, Q.Y.; Zhu, C.L.; Zhang, X.T.; Chen, Y.J. Porous MoO3 /SnO2 Nanoflakes with n–n Junctions for Sensing H2 S.
ACS Appl. Nano Mater. 2019, 2, 2418–2425. [CrossRef]
204. Yin, L.; Chen, D.L.; Feng, M.J.; Ge, L.F.; Yang, D.W.; Song, Z.H.; Fan, B.B.; Zhang, R.; Shao, G.S. Hierarchical Fe2 O3 @WO3
nanostructures with ultrahigh specific surface areas: Microwave-assisted synthesis and enhanced H2 S-sensing performance. RSC
Adv. 2015, 5, 328–337. [CrossRef]