The Combination of Two-Dimensional Nanomaterials With Metal Oxide Nanoparticles For Gas Sensors: A Review

Download as pdf or txt
Download as pdf or txt
You are on page 1of 40

nanomaterials

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;
[email protected] (T.L.); [email protected] (W.Y.); [email protected] (Y.S.); [email protected] (S.W.)
2 State Key Laboratory, Qingdao University, No. 308 Ningxia Road, Qingdao 266071, China;
[email protected] (S.G.); [email protected] (P.X.)
3 College of Chemistry and Chemical Engineering, Qingdao University, No. 308 Ningxia Road,
Qingdao 266071, China; [email protected] (H.K.); [email protected] (G.Y.)
* Correspondence: [email protected]; 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

Received: 28 February 2022


Accepted: 14 March 2022
1. Introduction
Published: 16 March 2022
In actual production and life, flammable, explosive, toxic, and harmful gases pose a
Publisher’s Note: MDPI stays neutral
serious threat to environmental safety and human health. Therefore, devices with high
with regard to jurisdictional claims in
performance are urgently needed to detect those flammable, explosive, toxic, and harmful
published maps and institutional affil-
gases. The gas sensors play great importance in determining various gases as they can
iations.
convert a certain gas volume fraction into electrical signals. They have the advantages of
low cost, fast response, high sensitivity, and high selectivity. In addition, in some cases, the
gas sensor device can be directly used in electronic interfaces. Therefore, gas sensors have
Copyright: © 2022 by the authors.
been widely used in environmental monitoring, air quality monitoring, vehicle exhaust
Licensee MDPI, Basel, Switzerland. monitoring, medical diagnosis, food/cosmetics monitoring, and many other fields [1–8].
This article is an open access article The fabrication of nanomaterial-based gas sensors has been the focus of research over
distributed under the terms and the past few decades. According to the working principle of the sensors, gas sensors can be
conditions of the Creative Commons divided into several types, such as semiconductor type, polymer type, contact combustion
Attribution (CC BY) license (https:// type, and solid electrolyte [9–12]. Among these gas sensors, the semiconductor gas sensors
creativecommons.org/licenses/by/ have developed into one of the largest and most widely used sensors in the world due to
4.0/). their large types of gases, high sensitivity, low price, and simple fabrication [13]. According

Nanomaterials 2022, 12, 982. https://doi.org/10.3390/nano12060982 https://www.mdpi.com/journal/nanomaterials


Nanomaterials 2022, 12, 982 2 of 40

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.

2. Working Principles of Gas Sensors


2.1. Mechanism of Oxygen Ion Adsorption on the Surface of Metal Oxide Nanoparticles
The sensing mechanism of traditional metal oxide-based gas sensors is related to the
resistance change of the sensing materials caused by the adsorption of oxygen ions on the
material surface [35,36]. When metal oxides are exposed to air, O2 in the air is adsorbed
onto the surface of metal oxides, which acts as electron acceptors to extract electrons from
the conduction band of oxides and dissociates into different forms of negative oxygen ions
(O2 − , O− , O2− ). Since the electrons in the conduction band of the material are captured
by oxygen anions, a hole accumulation layer (also called electron depletion layer) rich in
hole carriers is formed on the surface of the material, thereby increasing the resistance of
the gas-sensing material. Various gases are adsorbed on the surface of metal oxides and
interact with oxygen anions to change the electrical conductivity of metal oxides. Taking
NiO as an example, when exposed to a reducing gas, the adsorbed oxygen undergoes a
redox reaction with the gas, and the captured electrons are released into the conduction
oxygen anions, a hole accumulation layer (also called electron depletion layer) rich in hole
carriers is formed on the surface of the material, thereby increasing the resistance of the
gas-sensing material. Various gases are adsorbed on the surface of metal oxides and inter-
act with oxygen anions to change the electrical conductivity of metal oxides. Taking NiO
as an2022,
Nanomaterials example,
12, 982 when exposed to a reducing gas, the adsorbed oxygen undergoes a redox 3 of 40
reaction with the gas, and the captured electrons are released into the conduction band of
the semiconductor, and these electrons combine with the holes present in the hole accu-
mulation layer of the band of the semiconductor,
sensor material [37].and these electrons
Therefore, combine with
the number the holes is
of carriers present in the hole
reduced,
accumulation layer of the sensor material [37]. Therefore, the number of carriers is reduced,
and the resistance value is increased, so as to achieve the purpose of gas detection, as
and the resistance value is increased, so as to achieve the purpose of gas detection, as
shown in Figure 1. shown in Figure 1.

Figure 1. Schematics of H2 sensing


Figure mechanism
1. Schematics for NiO
of H2 sensing sensor. Hole
mechanism accumulation
for NiO sensor. Holeof the NiO sensor
accumulation of the NiO
exposed in air (a) andsensor
hydrogen (b),
exposed in respectively. Reprinted
air (a) and hydrogen with permission
(b), respectively. Reprintedfrom Ref. [37]. from
with permission Copy-Ref. [37].
right 2018 Elsevier. Copyright 2018 Elsevier.
2.2. Charge Transfer Mechanism of 2D Material-Based Gas Sensors
2.2. Charge Transfer Mechanism of 2D
The sensing Material-Based
mechanism Gas Sensors
of gas sensors based on graphene and related 2D layered
materials is mainly
The sensing mechanism of gasrelated
sensors to the charge
based ontransfer
grapheneprocess,
andinrelated
which the2Dsensing
layeredmaterial
acts as a charge acceptor or donor [38]. When exposed to
materials is mainly related to the charge transfer process, in which the sensing materialdifferent gases, a charge transfer
reaction occurs between the sensing material and the adsorbed gas, and the direction
acts as a charge acceptor or donor [38]. When exposed to different gases, a charge transfer
and amount of charge transfer are different, resulting in different changes in the material
reaction occurs between the sensing
resistance. material
Taking layered MoSand the adsorbed gas, and the direction and
2 as an example, the charge transfer between different gas
amount of charge transfer
moleculesare different,
(including resulting
O2 , H 2 O, NH3 , NO,in different
NO2 , CO)changes in theMoS
and monolayer material re- [39],
2 is different
as shown in Figure 2.
sistance. Taking layered MoS2 as an example, the charge transfer between different gas
Before gas adsorption, some electrons already exist in the conduction band of the
molecules (including O2, H2O, NH3, NO, NO2, CO) and monolayer MoS2 is different [39],
n-type MoS2 monolayer. When MoS2 is exposed to O2 , H2 O, NO, NO2 , and CO gases, the
as shown in Figure electron
2. charge is transferred from MoS2 to the gas atmosphere, resulting in a decrease
Before gas adsorption, some
in the carrier electrons
density of MoSalready exist ininthe
2 and an increase theconduction band
resistance of MoS of the
2 . On n-
the contrary,
When MoS is exposed
type MoS2 monolayer. When2 MoS2 is exposed to NH 3 , the NH molecules adsorbed
to O32, H2O, NO, NO2, and CO on MoS 2 actgases, thedonors
as charge
to transfer electrons to the MoS2 monolayer, increasing the carrier density of the MoS2
electron charge is transferred from MoS2 to the gas atmosphere, resulting in a decrease in
monolayer and reducing its resistance (Figure 2).
the carrier density of MoS2 and an increase in the resistance of MoS2. On the contrary,
When MoS2 is exposed to NH3, the NH3 molecules adsorbed on MoS2 act as charge donors
to transfer electrons to the MoS2 monolayer, increasing the carrier density of the MoS2
monolayer and reducing its resistance (Figure 2).
Nanomaterials 2022, 12, x FOR PEER REVIEW  4  of  40 
  Nanomaterials 2022, 12, 982 4 of 40

 
(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 3. Schematic diagram


Figure of the diagram
3. Schematic sensing of
mechanism
the sensingfor GO/SnOfor
mechanism 2. Reprinted
GO/SnO2 . with permission
Reprinted with permission
from Ref. [40]. Copyright 2019 ACS.
from Ref. [40]. Copyright 2019 ACS.

3. Factors for Affecting the Sensitivity of Gas Sensors


3. Factors for Affecting the Sensitivity of Gas Sensors
3.1. Size, Morphology, and Porosity
3.1. Size, Morphology,Theandgrain
Porositysize, morphology, and porosity are important factors for affecting the
The grain size, morphology,
sensing performance and of porosity are important
semiconductor-based gasfactors
sensors. forInaffecting
gas sensors, thesemiconductor
sens-
ing performancenanoparticles are connected to
of semiconductor-based gasadjacent
sensors. particles
In gasthrough
sensors, grain boundaries tona-
semiconductor form larger
noparticles are connected to adjacent particles through grain boundaries to form largerelectrons
aggregates [41]. On the particle surface, the adsorbed oxygen molecules extract
aggregates [41]. from
On thethe particle
conduction band and
surface, thecapture
adsorbed electrons
oxygen in the form of ions
molecules on theelectrons
extract surface, resulting
in band bending and electron depletion layers, as shown in Figure 4 [42,43]. Since the
from the conduction band and capture electrons in the form of ions on the surface, result-
transport of electrons between grains needs to pass through the electron depletion layer,
ing in band bending and electron depletion layers, as shown in Figure 4 [42,43]. Since the
the grain size has a great influence on the conductivity, which in turn affects the gas sensing
transport of electrons between
performance grains
of the needs to pass
material-based through
gas sensor. the electron
When the particle depletion
size (D) layer,
is much larger
the grain size has a great
than twiceinfluence
the thickness on(L)theofconductivity, which in
the electron depletion turn
layer (Daffects
>> 2L), the
theregasis asens-
wide electron
ing performancechannel
of thebetween
material-based
the grains,gas andsensor. When theofparticle
the gas sensitivity size is(D)
the material is much
mainly controlled by
the surface of the nanoparticles (boundary control). When
larger than twice the thickness (L) of the electron depletion layer (D >> 2L), there is a wide D ≥ 2L, there is a constricted
electron channelconduction
between the channel.
grains, Theandchange in electrical
the gas conductivity
sensitivity dependsisnot
of the material only on
mainly the particle
con-
trolled by the surface of the nanoparticles (boundary control). When D ≥ 2L, there is sensitivity
boundary barrier but also on the cross-sectional area of the channel, and the gas a
of the material is mainly controlled by the contact neck between nanoparticles (neck control).
constricted conduction channel. The change in electrical conductivity depends not only
When D < 2L, the electron depletion layer dominates, the entire nanoparticle is contained
on the particle boundary barrier but also on the cross-sectional area of the channel, and
in the electron depletion layer (grain control), and the sensitivity of the material is very
the gas sensitivity of the material
high. The energy bands is mainly controlled
are nearly by the contact
flat throughout neck between
the interconnected grainnano-
structure, and
particles (neck control).
there is noWhen D < 2L,
significant the electron
impediment to thedepletion
inter-grainlayer
chargedominates,
transport. The thesmall
entire
amount of
nanoparticle is contained
charge gained in the
fromelectron depletion
the surface reaction layer (grain
results in a control),
large change andinthethesensitivity
conductivity of the
of the material is very
entire high. The
structure. energythe
Therefore, bands
smalleraregrain
nearlysizeflat throughout
is beneficial the intercon-
to improving the sensitivity
of the gasand
nected grain structure, sensor.
thereWhenis nothesignificant
grain size is impediment
small enough, to thethecrystal becomes very
inter-grain chargesensitive to
transport. The small amount of charge gained from the surface reaction results in a large material
the surrounding gas molecules. For example, Min et al. used SnO 2 as the sensing
to prepare SnO2 films with different particle sizes (8.4–18.5 nm) and different porosity
change in the conductivity of the entire structure. Therefore, the smaller grain size is ben-
(70.8–99.5%), and found that in the gas sensors with porous structure, high porosity and
eficial to improving the sensitivity of the gas sensor. When the grain size is small enough,
low average particle size exhibited quicker gas sensing response [44].
the crystal becomes very sensitive to the surrounding gas molecules. For example, Min et
al. used SnO2 as the sensing material to prepare SnO2 films with different particle sizes
(8.4–18.5 nm) and different porosity (70.8–99.5%), and found that in the gas sensors with
porous structure, high porosity and low average particle size exhibited quicker gas sens-
ing response [44].
Nanomaterials 2022, 12, 982 6 of 40
Nanomaterials 2022, 12, x FOR PEER REVIEW 6 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.

3.2. Doping of Metals


The doping of metal elements can effectively improve the gas recognition ability of
gas sensing materials, and is an important method to improve gas-sensing performance.
Different dopant species may lead to different types of crystallites, defects and electronic
properties [51,52]. Doping or surface modification by adding metal elements (such as Ag,
Au, Pt, Pd, etc.) on the surface of the gas-sensing materials can increase the number of
active sites, promote the adsorption/desorption reaction on the surface of the gas-sensing
materials, and reduce the reaction activation energy, and reduce the operating temperature,
thereby improving the gas-sensing performance [53].
For instance, Fedorenko et al. [54] prepared Pd-doped SnO2 semiconductor sensors by
a sol–gel method. The effect of Pd additives on methane sensitivity was studied, and it was
found that due to the catalytic activity of Pd, compared with undoped materials, the addi-
tion of Pd to SnO2 significantly improved the sensor response to methane (about 6–7 times).
Barbosa et al. [55] studied the sensing responses of SnO micro-sheets that modified with Ag
and Pd noble metal catalysts towards the gases such as NO2 , H2 , and CO, and found that
the Ag/Pd surface-modified SnO micro-sheets exhibited higher sensitivity to gases such
as H2 and CO. However, the catalyst particles reduced the sensing response to oxidizing
gases such as NO2 . It is clear that the catalytic activity of Pd nanoparticles is related to
chemical sensitization, while the catalytic activity of Ag nanoparticles is related to electronic
sensitization. The Ag-modified samples showed high response to H2 , and Pd-modified
samples showed high response and selectivity to CO. Zhang et al. synthesized Co-doped
sponge-like In2 O3 cubes by a simple and environmentally friendly hydrothermal method
with the help of organic solvents, and studied their acetone gas-sensing properties [56]. It
was found that Co-doped In2 O3 has good gas-sensing performance for acetone gas, and its
porous structure can create more adsorption sites for the adsorption of oxygen molecules
and the diffusion of the target gas, thereby significantly improving the sensing performance.
Compared with the undoped sample, the response value of the doped Co-In2 O3 sample to
acetone was increased by 3.25 times, the response recovery time was fast (1.143 s/37.5 s),
the detection limit was low (5 ppm), the reproducibility was good, and the selectivity was
high. In another case, Ma et al. reported a Pt-modified WO3 mesoporous material with high
sensitivity to CO [57]. Pt acted as a chemical sensitizer, and gas molecules were adsorbed
and flowed into the gas-sensitive material through the spillover effect. In addition, the PtO
formed on the surface of Pt further increased the electron depletion layer, and enhanced
the electron sensitivity. The synergistic effects of both components further improved the
gas-sensing performance.
Nanomaterials 2022, 12, 982 8 of 40

Compound doping is another important method to improve the sensing performance


of gas sensors. When a metal oxide is combined with other metal oxides, a heterojunction
structure is constructed. Since the two materials each have their own Fermi energy levels,
there will be a mutual transfer of carriers between the two materials to form a space charge
layer, so as to achieve the purpose of enhancing the gas-sensing properties of the compound
materials. For example, Ju et al. [58] prepared SnO2 hollow spheres by a template-assisted
hydrothermal method and successfully implanted p-type NiO nanoparticles onto the
surface of SnO2 hollow spheres by the pulsed laser deposition (PLD) to prepare NiO/SnO2
p-n hollow spheres. The gas-sensing performance test indicated that its response to 10 ppm
triethylamine (TEA) gas could reach 48.6, which was much higher than that of the original
SnO2 hollow spheres, and the detection limit was as low as 2 ppm. The optimal operating
temperature dropped to 220 ◦ C, which was 40 ◦ C lower than that of the original SnO2
hollow sphere sensor. Compared with the pristine SnO2 sensor, the enhanced response of
NiO/SnO2 sensor to TEA is mainly attributed to the formation of a depletion layer by the
p-n heterojunction interface, which makes the resistance of hybrid materials in air and TEA
gas change a lot.
In addition, when the gas-sensing materials of different dimensions are compounded,
the stacking of the gas-sensing materials can be prevented, the porosity can be increased,
and the gas-sensing performance can be improved. For example, Kida et al. [59] introduced
monodispersed SnO2 nanoparticles (about 4 nm) into WO3 nanosheet-based films, which
could improve its porosity, prevent the aggregation of flakes, and increase the diffusion
paths and adsorption sites of gas molecules. The response sensitivity of the composites
to NO2 was enhanced when the concentration in air was 20–1000 ppb, indicating the
effectiveness of the microstructure control of the WO3 -based film on high-sensitivity NO2
detection. In another case, Mishra et al. [60] prepared nanocubic In2 O3 @RGO composites by
combining In2 O3 with reduced graphene oxide (RGO), and the sensor based on nanocubic
In2 O3 @RGO heterostructures exhibited high resistance to acetone (~85%) and formaldehyde
(~88%) with good selectivity, long-term stability, and fast response/recovery rates

4. 2D Material-Based Gas Sensors


With the successful preparation of graphene materials, its unique structure and excel-
lent properties have attracted widespread attention, thus setting off a research upsurge
in 2D materials. 2D nanomaterials have a large specific surface area and special electrical
properties due to their nanoscale thin-layer structure. After the gas is adsorbed on the
surface, it will affect the conductivity of the surface, so it can be used as a gas-sensing
material to adsorb and capture certain single species of gas molecules, with excellent gas-
sensing properties. Gas sensors based on 2D nanomaterials exhibit many advantages, such
as high sensitivity, fast response speed, low energy consumption, and the ability to work at
room temperature.

4.1. 2D Graphene-Based Gas Sensors


Graphene is a honeycomb 2D carbon nanomaterial composed of a single layer of sp2
carbon atoms. It is currently the thinnest 2D material in the world, with a thickness of
only 0.35 nm. Graphene has many excellent properties due to its special structure, such as
good electrical conductivity, high carrier mobility, transparency, and mechanical strength.
As a typical 2D material, every atom in the graphene structure can be considered as a
surface atom, so ideally every atom can interact with the gas, which makes graphene
promising as a kind of gas sensor with ultrahigh sensitivity. In the process of adsorption
and desorption of gas molecules, graphene nanosheets will affect the change of local carrier
concentration in the material, thus showing the transition of electrical signal in the detection
of electrochemical performance, and has a good development prospect in gas adsorption. In
2007, Novoselov’s group first reported a graphene-based gas sensor, which confirmed that
a graphene-based nanoscale gas sensor can be used to detect the adsorption or desorption
Nanomaterials 2022, 12, 982 9 of 40

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].

4.2. 2D Transition Metal Sulfide-Based Gas Sensors


As a typical p-type inorganic 2D material with a hexagonal filled layered structure
of TMDs, it has received extensive attention in energy conversion and storage, especially
in room temperature gas sensors, which have unique advantages and are widely used in
various gas detection. Similar to graphene, MoS2 consists of vertically stacked layers, each
formed by covalently bonded Mo-S atoms, with adjacent layers connected by relatively
weak van der Waals forces. The weak van der Waals interactions allow gas molecules
to permeate and diffuse freely between the layers, so the resistance of MoS2 can change
dramatically with the adsorption and diffusion of gas molecules within the layers. Various
methods for gas sensing using few-layer MoS2 have been reported in the literature, includ-
ing detectors for many kinds of chemical vapors such as H2 , NO2 , and ethanol [74–76].
For example, Li et al. [77] found for the first time that mechanically exfoliated multilayer
MoS2 exhibited high sensitivity to NO gas, while monolayer MoS2 had an unstable re-
sponse to NO gas. They used a mechanical lift-off technique to deposit monolayer and
multilayer MoS2 films on Si/SiO2 surfaces for the fabrication of FETs. The FET acted as a
gas sensor, which realized gas detection by monitoring the change of the conductance of
the FET channel during the adsorption of target gas molecules. Since the mechanically cut
MoS2 sheet is an n-type semiconductor, when the MoS2 channel was exposed to NO gas,
it would result in p-doping of the channel, resulting in an increase in channel resistance
and a decrease in current flow. It was found that although the single-layer MoS2 device
exhibited a fast response after exposure to NO, the current was unstable; the two-layered,
three-layered, and four-layered MoS2 devices all exhibited stable and sensitive responses
to NO at a concentration of 0.8 ppm. Late et al. [78] systematically studied the relationship
between the number of MoS2 layers and gas sensing performance, and found that the
sensitivity and recovery time of 5-layered MoS2 to NH3 and NO2 gases were better than
those of double-layered MoS2 (Figure 6a–d). These findings suggest that a small amount of
layered MoS2 has great potential to detect various polar gas molecules.
three-layered, and four-layered MoS2 devices all exhibited stable and sensitive responses
to NO at a concentration of 0.8 ppm. Late et al. [78] systematically studied the relationship
between the number of MoS2 layers and gas sensing performance, and found that the sen-
sitivity and recovery time of 5-layered MoS2 to NH3 and NO2 gases were better than those
of double-layered MoS2 (Figure 6a–d). These findings suggest that a small amount of lay-
Nanomaterials 2022, 12, 982 11 of 40
ered MoS2 has great potential to detect various polar gas molecules.

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.

4.3. 2D Metal Oxide Based Gas Sensors


2D semiconductor oxide nanosheets are also commonly used 2D materials in the
field of gas sensors. Among them, layered MoO3 , WO3 and SnO2 have attracted much
attention due to their stability in high-temperature air [84–86]. For example, Cho et al. [87]
reported the preparation of MoO3 nanosheets by ultrasonic spray pyrolysis, and studied
their gas-sensing properties, and found that there was still a response even when the gas
concentration of trimethylamine was lower than 45 ppb. The super sensitivity to trimethy-
lamine gas is inseparable from its larger specific surface area, as ultrathin nanosheets with
the larger specific surface area can provide a larger electron depletion layer and a faster gas
diffusion rate across the nanosheets. In addition, MoO3 is an acidic oxide, which is more
likely to react with basic gas preferentially, so it has super selective properties for basic gas
trimethylamine. Wang et al. [88] prepared WO3 porous nanosheet arrays by chemical bath
deposition, and found that WO3 arrays composed of 20 nm ultrathin porous nanosheets
had better low-temperature NO2 gas sensing properties. At an operating temperature of
100 ◦ C, the response to 10 ppm NO2 was as high as 460.
For 2D metal oxide nanomaterials, the difference in exposed crystal planes will affect
their gas sensing properties. For example, Kaneti et al. used a simple and effective
hydrothermal method to prepare ZnO nanosheets. By simulating the adsorption of gas
molecules on the surfaces of different ZnO crystals, it was found that the enhanced gas-
sensing performance of ZnO nanosheets is related to the exposed surface, the (1010) face of
ZnO possesses better adsorption capacity for n-butanol than the (1120) face and (0001) face,
showing higher responsiveness, better selectivity, and higher stability [89]. In addition,
Wang et al. [90] used a two-step method to synthesize ultrathin porous In2 O3 nanosheets
with uniform mesopores and found that they exhibited an ultra-high response to 10 ppb
NOx at a lower operating temperature (120 ◦ C), its sensitivity response value was 213,
and the response time was 4 s. The thickness of the ultrathin nanosheets is about 3.7 nm,
with a large number of active reaction sites, which can enhance the response to NOx, and
the porous structure can shorten the gas transmission path and enhance the gas diffusion
efficiency, thereby improving the gas sensing performance. Wang et al. synthesized
Co3 O4 mesh nanosheet arrays for the detection of NH3 [91]. The porous mesh structure
promoted gas diffusion and provided a larger active reactive surface to react with the
target gas, thereby improving the gas sensing performance. Therefore, even when the
NH3 concentration is 0.2 ppm, the sensor still has obvious response characteristics. The
response/recovery time of Co3 O4 nanosheet arrays to 0.2 ppm NH3 is 9 s/134 s, showing
good reproducibility and long-term room temperature stability.
2D nanostructures have shown great potential in the field of gas sensing due to their
high specific surface area and highly efficient active sites on exposed surfaces. To further en-
Nanomaterials 2022, 12, 982 14 of 40

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.

4.4. Other 2D Material-Based Gas Sensors


MXene is a new type of 2D material with layered structure discovered in recent years,
which is generally transition metal carbide or carbon-nitrogen compound, and is a MAX
ternary phase material. Its general structural formula is Mn+1 AXn (n = 1, 2 or 3), where
M is one of transition metal elements, A is one of the main group elements (mainly III,
IV group elements), X It is carbon or nitrogen, and there are more than 70 kinds of MAX
materials. MXene materials are 2D materials formed by extracting element A in MAX. The
general formula is Mn+1 Xn (n = 1, 2 or 3) [93,94]. Due to the characteristics of conventional
semiconductor materials and the fact that a large number of functional groups and other
active sites remain on the surface after etching, which facilitates subsequent modification,
such materials have great application potential in the field of sensing [94,95].
Xiao et al. applied MXene nanomaterials to the detection of NH3 gas in 2015 [83]. Since
the successful synthesis of 2D compound MXenes by Gogotsi et al. in 2011, the application
of 2D MXene nanomaterials in gas sensing has been continuously developed [96–98]. For
example, Lee et al. [99] reported a Ti3 C2 Tx -based gas sensor. After studying the room-
temperature gas sensing performance of Ti3 C2 Tx nanosheets on flexible polyimide, it was
found that the Ti3 C2 Tx sensor exhibited p-type sensing behavior for reduced gases, with a
theoretical detection limit of 9.27 ppm for acetone. Based on the charge interaction between
gas molecules and Ti3 C2 Tx surface functional groups -O and -OH, the sensing mechanism
of Ti3 C2 Tx is proposed. Chae et al. [100] investigated the dominant factors affecting the
oxidation rate of Ti3 C2 Tx flakes and their corresponding sensing properties. In order to
improve the sensing performance of MXenes, the gas sensing performance of MXene-based
sensors has been realized by surface chemistry and composite structure. Yang et al. [101]
prepared organic-like Ti3 C2 Tx by HF acid etching, added the prepared powder to NaOH
solution, and used alkali treatment to demonstrate the effect of surface groups on its sensing
performance. It was found that the response of the alkali-treated Ti3 C2 Tx sensor to 100 ppm
NH3 at room temperature was two times higher than that of the untreated one. This is
due to the adsorption of N atoms in NH3 molecules on top of Ti atoms in Ti3 C2 Tx to form
strong N-Ti bonds. Alkaline treatment increased the -O end, increased N-Ti bond, and
promoted the increase of NH3 adsorption. Furthermore, after oxygen functionalization,
Ti3 C2 Tx increased the resistance by transitioning to a semiconductor, thereby increasing
the gas response signal. Besides Ti3 C2 Tx , 2D MXenes such as V2 CTx and Mo2 C have also
been investigated for gas sensors [102,103]. The 2D V2 CTx sensor composed of monolayer
or multilayer 2D V2 CTx on polyimide film fabricated by Lee et al. [102] can measure polar
gases (hydrogen sulfide, ammonia, acetone, and ethanol) and non-polar gases at room
temperature (hydrogen and methane). The V2 CTx sensor shows ultrahigh sensitivity for
non-polar gases, with minimum detection limits of 2 ppm and 25 ppm for hydrogen and
methane, respectively.

5. Metal Oxide Nanomaterials-Based Gas Sensors


Because of its large specific surface area, high surface activity, many active sites
and sensitive to the surrounding environment, the gas sensor prepared by metal oxide
Nanomaterials 2022, 12, 982 15 of 40

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].

5.1. SnO2 -Based Gas Sensors


SnO2 is a kind of direct band gap wide band gap n-type semiconductor (band gap
~ 3.6 eV), whose carriers are free electrons. The interaction with the reducing gas will
increase the electrical conductivity. However, the oxidized gas will consume the sensing
layer of charged electrons, resulting in a decrease in electrical conductivity [108]. SnO2
nanomaterials are widely used in the field of gas sensing because of their simple preparation,
low cost, easy control of morphology, and microstructure, good thermal/chemical stability,
shallow donor energy level (0.03–0.15 eV), potential barrier of oxygen adsorption on
the surface is 0.3–0.6 eV, oxygen vacancy and excellent gas-sensing properties [109–111].
Thanks to its high sensitivity to different gases, SnO2 sensor can detect low concentration
gases, but its selectivity is low.
In order to improve the sensitivity, stability, and selectivity of SnO2 -based gas sensors
and reduce the working temperature, researchers modified SnO2 materials by a variety of
methods. One method is to control the morphology and size of SnO2 materials to prepare
zero-dimensional (0D), one-dimensional (1D), 2D, three-dimensional (3D) and porous
hollow SnO2 nanomaterials for the detection of various gases [112–117]. For example,
Zhang et al. successfully synthesized leaf-like SnO2 hierarchical architectures by using a
simple template-free hydrothermal synthesis method. The sensor based on this unique
leaf-like SnO2 hierarchical structure had a high response and good selectivity to NO2 at
low operating temperature [118]. Feng et al. synthesized mesoporous SnO2 nanomaterials
with different pore sizes (4.1, 6.1, 8.0 nm) by carbon-assisted synthesis. The gas sensing
properties of the three materials showed high sensitivity and ideal response recovery time
to ethanol gas, and the detection limit was as low as ppb [119].
Using doping modification technology, SnO2 nanomaterials are used as the matrix
materials of gas sensors, which are modified by doping precious metals (such as Pt, Pd and
Au) or other metal ions (such as Ni, Fe and Cu). It is another important means to improve
the gas sensing properties of SnO2 to CO, CH4 , NO2 and other gases. For example, Dong
et al. prepared SnO2 nanofibers and Pt-doped SnO2 nanofibers by electrospinning, which
were used to test the sensitivity to H2 S. It was found that the response of Pt-doped SnO2
nanofibers to H2 S gas was significantly improved. The response of 0.08 wt% Pt-doped
SnO2 nanofibers to 4–20 ppm H2 S was 25.9–40.6 times higher than that of pure SnO2
nanofibers [120]. Chen et al. prepared Pd-doped SnO2 nanoparticles by the coprecipitation
method. Compared with pure SnO2 nanoparticles, the response characteristics of SnO2
to CO were significantly improved [121]. Lee et al. used Pd nanoparticles to modify the
surface of SnO2 nanorod thin films, and studied their sensing properties for H2 and ethanol
gas [122]. It was found that compared with the undoped samples, the responsiveness of
Pd-doped SnO2 nanorod thin films to 1000 ppm H2 and ethanol at 300 ◦ C was increased by
6 and 2.5 times, respectively. They assumed that the improved gas sensing properties are
due to the formation of the electron depletion layer and the enhanced catalytic dissociation
of molecular adsorbates on the surface of Pd nanoparticles. Shen et al. also studied the
gas-sensing properties of SnO2 by Pd doping [123]. SnO2 nanowires with a tetragonal
structure were synthesized by thermal evaporation. The morphology, crystal structure,
Nanomaterials 2022, 12, 982 16 of 40

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].

5.2. ZnO-Based Gas Sensors


ZnO is an n-type metal oxide semiconductor material with a wide band gap (3.3 eV).
ZnO is widely used in the field of gas sensors because of its good chemical stability and low
resistivity. Yuliarto et al. successfully synthesized ZnO nanorod thin films on Al2 O3 sub-
strates by chemical bath deposition (CBD) [129]. ZnO thin films with different thicknesses
were prepared by different times of CBD processes. By optimizing the thickness of ZnO
thin films, the response performance of ZnO-based gas sensors to SO2 was improved. The
gas sensing response of the ZnO film of two CBD to 70 ppm SO2 at 300 ◦ C is 93%, which is
15% higher than that of the ZnO film of one CBD. At different operating temperatures, the
response of ZnO nanorods prepared by two CBD deposition is 20–40% higher than that
of ZnO nanorods deposited by one CBD deposition. Wang et al. successfully prepared
three kinds of ZnO nanostructures (nanorods, flowers, and spheres) with different mor-
phologies by a simple hydrothermal and water-bath method, and studied their sensing
properties of NO2 at room temperature under UV (365 nm LED) excitation, as shown in
Figure 8 [130]. It was found that ZnO nanospheres have the highest response (29.4) to
5 ppm NO2 (Figure 8d,e), which was mainly due to the largest specific surface area and the
largest number of oxygen ions adsorbed on the surface of ZnO nanospheres. However, due
to the high crystallinity, few surface defects and unidirectional electron transfer path, the
response speed and recovery speed of ZnO nanorods are the fastest (9 s and 18 s, respec-
tively) (Figure 8f,g). For ZnO nanoflowers, the gas sensing response, response and recovery
rate are between ZnO nanorods and ZnO nanospheres. All three kinds of ZnO have good
selectivity and repeatability for NO2 (Figure 8h,i). The good selectivity of ZnO to NO2 is
attributed to the following two points: (1) NO2 molecule has an unpaired electron, which
is beneficial to its chemisorption on ZnO surface; (2) NO2 molecule has the smallest bond
However, due to the high crystallinity, few surface defects and unidirectional electron
transfer path, the response speed and recovery speed of ZnO nanorods are the fastest (9 s
and 18 s, respectively) (Figure 8f,g). For ZnO nanoflowers, the gas sensing response, re-
sponse and recovery rate are between ZnO nanorods and ZnO nanospheres. All three
kinds of ZnO have good selectivity and repeatability for NO2 (Figure 8h,i). The good se-
Nanomaterials 2022, 12, 982 17 of 40
lectivity of ZnO to NO2 is attributed to the following two points: (1) NO2 molecule has an
unpaired electron, which is beneficial to its chemisorption on ZnO surface; (2) NO2 mole-
cule has the smallest bond energy, which is about 312.7 kJ/mol. The smaller the bond en-
energy, which is about 312.7 kJ/mol. The smaller the bond energy is, the more favorable
ergy is, the more favorable the sensing reaction is, especially for the sensors working at
the sensing reaction is, especially for the sensors working at room temperature.
room temperature.

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

The heterostructure is an important means to improve the gas sensing properties of


semiconductor oxides, which usually includes two kinds of semiconductor oxides with
different Fermi levels. When two kinds of semiconductor oxides come into contact with
each other, the free electrons will change from the oxidation stream with a higher Fermi
level to the oxide with a lower Fermi level. Compared with the single semiconductor
oxide, the electron transfer efficiency of the two semiconductor oxides is higher, and a
thicker electron depletion layer and higher resistance can be formed at the contact interface.
Therefore, the introduction of heterojunction can effectively improve the performance
of the semiconductor gas sensor. For example, Kim et al. synthesized p-n CuO/ZnO
core–shell nanowires by thermal oxidation and atomic layer deposition, and studied their
sensing properties to reduce gas by controlling the thickness of the ZnO shell [136]. When
the shell thickness is less than or equal to Debye wavelength (λD ), a complete electron
depletion layer will be formed. When exposed to the reducing gas (CO), the desorption
of surface oxygen releases electrons back into the conduction band of the shell, returning
the conduction band to its original state and significantly improving the conductivity
(Figure 9a). When the thickness of the shell is higher than λD , only part of the electron loss
will be caused. When reducing gases are introduced, they are adsorbed on the partially
depleted shell, and the resistance changes only slightly, as shown in Figure 9b. Therefore,
for p-n heterostructure nanowires, controlling the shell thickness plays an important role
in improving the performance of gas sensors. Zhou et al. prepared NiO/ZnO nanowires
by one-step hydrothermal method and tested their gas sensing properties to SO2 [137]. It
was found that at the optimum operating temperature of 240 ◦ C, the response of NiO/ZnO
nanowires to 50 ppm SO2 was 28.57, and the gas detection range was 5–800 ppm. The
Nanomaterials 2022, 12, x FOR PEER response
REVIEW time, response time and recovery time of the prepared NiO-ZnO nanowires 19 of 40 gas
sensor to 20 ppm SO2 gas were 16.25, 52, and 41 s, respectively.

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

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
properties. For example, Li et al. prepared porous CuO nanosheets on alumina tubes by
hydrothermal method, which were used to make gas sensors to detect H2 S [138]. It was
found that when the concentration of H2 S is as low as 10 ppb, the response sensitivity of
the sensor was 1.25, and the response and recovery time were 234 and 76 s, respectively.
Navale et al. synthesized CuO thin films that composed of CuO nanocubes on quartz
substrates by simple and catalyst-free thermal evaporation technique, and studied their
gas sensing properties [139]. It was found that CuO thin films have strong selectivity for
NO2 gas, and the response speed and recovery time are fast. At 150 ◦ C, the maximum
response value of CuO sensor film to NO2 of 100 ppm was 76, the detection limit was
1 ppm, and the response time was only 6 s, but the recovery time was 1200 s. Huang et al.
prepared CuO hollow microspheres by precipitation annealing at 270 ◦ C using CuSO4 ,
Na2 CO3 , and cetyltrimethyl ammonium bromide (CTAB) as raw materials [140]. The CuO
hollow microspheres showed good gas sensitivity to ethanol. The response sensitivity
to ethanol at 250 ◦ C was 5.6 and the response and recovery times were 17.0 and 11.9 s,
respectively. Hu et al. fabricated CuO nanoneedle arrays directly on commercial ceramic
Nanomaterials 2022, 12, x FOR PEER tubes
REVIEW by magnetron sputtering, wet chemical etching and annealing, which have good
selectivity, reproducibility and long-term stability for low concentration H2 S (10 ppm) [141].
For metal oxide semiconductors, high specific surface area and exposed crystal plane are
two key factors that determine their gas sensing properties. In order to study the effect
ofsurfaces
surface structure
and CuO on gas sensing properties,
nanocubes on (110)Huo et al. obtained
exposed CuO
surfaces innanotubes on
Cu nanowires an
(111) exposed surfaces and CuO nanocubes on (110) exposed surfaces in Cu nanowires and
nanocubes, respectively, which were used to detect the gas sensing properties of
Cu2 O nanocubes, respectively, which were used to detect the gas sensing properties of CO
as shown
gas, as shownin in Figure
Figure 1010 [142].
[142]. The The results
results indicated
indicated that with
that compared compared with CuO nan
CuO nanocubes,
CuOnanotubes
CuO nanotubes havehave
lowerlower optimal
optimal operating operating temperatures
temperatures and higher
and higher sensitivity sensitivit
for CO
gas detection.
gas detection.

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


and harmful gases. However, it still faces some problems in practical applicatio
as low sensitivity, high working temperature, poor selectivity, long response/
time and so on. For this reason, researchers use doping, recombination and other
to improve the gas sensing performance of CuO-based gas sensors, and achiev
remarkable results. The doping of precious metals or rare earth elements can gr
crease the active sites of CuO gas-sensing reaction, which is beneficial to the ad
Nanomaterials 2022, 12, 982 20 of 40

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.

5.4. Other Metal Oxide-Based Gas Sensors


WO3 is a kind of n-type wide band gap semiconductor oxide, which has the advantages
of photoelectric conversion, electrochromism, photocatalysis, and gas sensitivity, so it is
used in a variety of optoelectronic devices. WO3 is more likely to form oxygen defects and
unsaturated coordination bonds. When WO3 is heated in the air, it is easy for O2 to seize
e− to form O− . The formed O− is chemically adsorbed on the surface of WO3 and forms an
electron depletion layer. When operating at a low temperature, it is easy to form O2− . when
it is at a higher operating temperature, it is easy to form O− and O2− [155,156]. Therefore,
WO3 is an effective gas sensing material, especially more sensitive to reducing gas. For
example, Hu et al. synthesized WO3 nanorods with needle-shape via a hydrothermal
mehod and subsequent calcination, which showed high performance for triethylamine gas
sensing [157]. By comparison, it was found that WO3 nanosheet devices showed the highest
response and the shortest response time to 1–10 ppm SO2 . Li et al. proposed a method
for the synthesis of WO3 particles assisted by ionic liquids. The hollow sphere structure
composed of WO3 nanorods, nanoparticles and nanosheets was synthesized. Their gas
sensing properties for various organic compounds (methanol, ethanol, isopropanol, ethyl
acetate and toluene) were studied. It was found that it has remarkable sensitivity, low
detection limit and fast response/recovery time [158]. Li et al. prepared SnO2 -WO3
hollow nanospheres with a diameter of about 550 nm and a thickness of about 30 nm
by hydrothermal method, and studied the temperature dependence of humidity sensors
prepared at different relative humidity and temperature. It is found that compared with
the original WO3 nanoparticles and SnO2 nanoparticles, SnO2 -WO3 hollow nanospheres
have excellent sensing properties [159].
α-Fe2 O3 is also a typical n-type semiconductor with a narrow band gap (2.2 eV), low
cost, high stability, high corrosion resistance, and non-toxicity, so it has attracted great
attention as a gas sensing material [160,161]. Liang et al. successfully synthesized ultrafine
and highly monodisperse α-Fe2 O3 nanoparticles with an average particle size of 3 nm by a
simple reverse microemulsion method, which showed high sensitivity, high selectivity and
good stability to acetone [162]. Shoorangiz et al. synthesized α-Fe2 O3 nanoparticles by a
sol–gel method and evaluated their gas-sensing properties for ethanol and other gases [163].
At the optimal sensing temperature of 150 ◦ C, it has good selectivity to ethanol gas, and
the response to 100 ppm ethanol gas was 14.5%. Qu et al. reported high-performance
gas sensors based on MoO3 nanoribbons that were modified by Fe2 O3 nanoparticles [164].
Compared with the original MoO3 nanoribbons, the reaction of p-xylene in the Fe2 O3
nanoribbons modified by Fe2 O3 nanoparticles increased by 2–4 times due to the formation
of heterojunction between Fe2 O3 and MoO3 .
As a p-type single metal oxide semiconductor, Co3 O4 is also used to test gas sensitivity.
It has been found to have a good gas sensing response to H2 S in some studies [91,165].
Among different types of metal oxides, p-type Co3 O4 is also considered to be the best
candidate for ethanol gas sensing. For example, Li et al. reported that Co3 O4 nanotubes
are sensitive to ethanol gas at room temperature, and Co3 O4 sensors show excellent re-
peatability after more than 50 tests [166]. Sun et al. obtained monodisperse porous Co3 O4
microspheres by solvothermal method and thermal decomposition [167]. The gas sensing
properties of these Co3 O4 microspheres were compared with those of commercial Co3 O4
nanoparticles. These Co3 O4 microspheres showed higher ethanol sensitivity and selectivity
at relatively low temperatures. In addition, Zhang et al. synthesized three kinds of Co3 O4
with different morphologies (cube, rod, and sheet) by a hydrothermal method, and studied
their sensing properties to toluene [168]. It was found that the sensor based on the Co3 O4
flake structure had better sensing performance for toluene than the other two sensors. At
Nanomaterials 2022, 12, 982 22 of 40

the working temperature of 180 ◦ C, the fabricated sensor showed higher sensitivity, faster
response and recovery speed, as well as better selectivity.

6. 2D Materials/Metal Oxide-Based Gas Sensors


As a class of important materials for various sensors, semiconducting metal oxides
have been widely used in various redox gas sensing due to their high sensitivity, simple
preparation, and low price. However, disadvantages such as high temperature and easy
agglomeration can also be found. Especially the disadvantage of high working temperature
makes it extremely demanding on the working conditions of the environment, which greatly
affects the service life. 2D materials such as graphene have unique physical and chemical
properties, which can generate gas-sensitive responses at room temperature and have good
selectivity. 2D materials can not only provide active sites for catalysts and sensors, but also
serve as flat building blocks for forming complex nanostructures. Using 2D materials as
a matrix to support semiconducting metal oxides can reduce the agglomeration of metal
oxides and expose more adsorption and reaction sites. Due to the large specific surface
area and high porosity of 2D materials, the response sensitivity and selectivity of metal
oxides to specific gases can be further improved, and the working temperature can be
reduced. Meanwhile, the synergistic effect and heterostructure of layered 2D materials
and metal oxides can not only bring out the greatest advantages of both components but
also overcome their respective defects, thereby improving the comprehensiveness of gas
sensing performance. Therefore, the combination of 2D materials and metal oxides has
become an important research direction in the field of gas sensors.

6.1. Synthesis of 2D Materials/Metal Oxide Composites


The properties of a material have a great relationship with its morphology, structure,
and its composition. Single-component nanomaterials are far from meeting the develop-
ment and application needs of modern nanotechnology, while multi-component composite
materials combine the characteristics of different materials show better performance than
their single components. It is of great significance in developing new materials, studying
novel properties of materials, and constructing functional gas sensors. The preparation
methods of 2D materials/metal oxide composites mainly include hydrothermal synthesis,
microwave-assisted synthesis, self-assembly, chemical reduction method, and others. In
this section, we would like to present a brief introduction to these potential synthesis
methods for 2D materials/metal oxide composites.
Hydrothermal synthesis is a common method for the preparation of 2D materi-
als/metal oxide nanocomposites, which has the advantages of simple operation, mild
conditions, and low cost. Usually, the precursor solution is put into a high-pressure reactor,
hydrothermally reacted under high temperature and high pressure, and then the composite
material is prepared by post-processing methods such as separation, washing, and drying.
Many 2D materials/metal oxide nanocomposites have been synthesized and used for the
fabrication of gas sensors. For example, Chen et al. [169] prepared a core–shell structure
composed of TiO2 nanoribbons and Sn3 O4 nanosheets by a two-step hydrothermal reaction.
Sn3 O4 nanosheets were uniformly immobilized onto the surface of porous TiO2 nanorib-
bons. It was found that the structural morphology of the products in the hydrothermal
process is affected by the reaction time. Chen et al. [170] investigated the effect of hy-
drothermal reaction temperature on the morphology of oxide materials. At a relatively high
temperature, they obtained SnO2 -decorated TiO2 nanoribbons. Wang et al. [171] obtained a
composite structure based on SnO2 nanoparticles and TiO2 nanoribbons by controlling the
precursor solution. Precise control of the hydrothermal synthesis conditions is a key factor
for the preparation of high-quality and diversely shaped metal oxide nanostructures. In
addition to metal oxide-based 2D composites, the composites of metal oxides and other 2D
materials such as graphene have also been prepared by hydrothermal methods. For exam-
ple, Chen et al. [172] prepared Co3 O4 /rGO composites by the hydrothermal method, and
studied their gas-sensing properties to NO2 and methanol at room temperature. Chen and
Nanomaterials 2022, 12, 982 23 of 40

co-workers [173] synthesized SnO2 nanorods/rGO composite nanostructures by hydrother-


mal method and investigated their NH3 sensing properties. Liu et al. [174] fabricated a
layered flower-like In2 O3 /rGO composites by a one-step hydrothermal method. The syn-
thesized materials were further utilized for the fabrication of gas sensors, which exhibited
improved sensing performance for 1 ppm NO2 at room temperature compared with pure
In2 O3 -based gas sensors. The hydrothermal synthesis usually reacts at high temperatures
(>150 ◦ C). When the metal oxide is coupled with 2D materials for gas sensing, the operation
temperature will be decreased, which is lower than the materials preparation temperature.
Thus, the thermal stability and applicability of the heterostructures for gas sensors can
be improved.
The microwave-assisted synthesis of materials uses microwaves to provide energy for
the reaction, and it is different from the traditional heating method. It uses microwaves
to make the reactants generate heat by themselves to promote the reaction. It has the
characteristics of uniform heating and high heating efficiency. For example, Pienutsa et al.
synthesized SnO2 decorated RGO and further used the created SnO2 -RGO composite for
the real-time monitoring of ethanol vapor [175]. In another similar study, Kim et al. [176]
obtained SnO2/graphene heterostructured composites by microwave-treating graphene/SnO2
nanocomposites, in which graphene-enhanced efficient transmission of microwave energy and
facilitated the evaporation and redeposition of SnOx nanoparticles.
Chemical reduction has been often used to synthesize composite nanostructures of
RGO and metal oxides. Usually, a metal salt solution is used as a precursor to be mixed with
a graphene oxide dispersion solution, and a chemical reducing agent is used to reduce it in
one step to obtain RGO. This method usually relies on a microwave, hydrothermal reaction,
and other sources to provide energy. For example, Russo et al. [177] synthesized SnO2 /rGO
composites using SnCl4 and GO as precursors. Under the irradiation of microwave, GO
and SnCl4 were reduced to form SnO2 /rGO composites, which were further reacted
with H2 PtCl6 to form Pt-SnO2 /rGO composites. The created composites exhibited high
sensitivity to hydrogen at room temperature.
Besides the above-introduced methods, other techniques such as self-assembly can
also be utilized for the synthesis of 2D materials/metal oxide composites. For instance,
Zhang et al. fabricated rGO/TiO2 multilayer composite films using a layer-by-layer self-
assembly process, and the fabrication process is shown in Figure 11 [178]. The rGO/TiO2
multilayer composite films were fabricated by alternately depositing TiO2 nanospheres
and GO via the layer-by-layer self-assembly technique to form nanostructures, and then
thermally reducing GO to rGO. Since p-type rGO and n-type TiO2 form a p-n heterojunction
at the interface, the depletion layer generated by the built-in electric field will be beneficial
to control the carrier transport process inside the material. It is clear that SO2 acts as
an electron donor, which increases the electron concentration of the composite material,
resulting in a decrease in device resistance. The device could be used to detect SO2 gas at
as low as 1 ppb at room temperature with good selectivity and stability. The response of
the fabricated gas sensor to 1 ppm SO2 was 10.08%, and the response and recovery times
were 95 and 128 s, respectively.
eficial to control the carrier transport process inside the material. It is clear that SO2 acts
as an electron donor, which increases the electron concentration of the composite material,
resulting in a decrease in device resistance. The device could be used to detect SO2 gas at
as low as 1 ppb at room temperature with good selectivity and stability. The response of
Nanomaterials 2022, 12, 982 24 of 40
the fabricated gas sensor to 1 ppm SO2 was 10.08%, and the response and recovery times
were 95 and 128 s, respectively.

Figure 11. Fabrication


Figureillustration ofillustration
11. Fabrication TiO2/rGOof multilayer hybrid hybrid
TiO2 /rGO multilayer film by layer-by-layer
film by layer-by-layerself-assem-
self-assembly.
bly. Reprinted with permission from Ref. [178]. Copyright 2017 Elsevier.
Reprinted with permission from Ref. [178]. Copyright 2017 Elsevier.

6.2. Graphene/Metal Oxide Composite-Based Gas Sensors


6.2. Graphene/Metal Oxide Composite-Based Gas Sensors
Metal oxide-based gas sensors have the advantages of low production cost, good stability,
Metal oxide-based gas sensors
wide application have
range, and easythe advantages
integration of lowdevices,
with portable production
and have cost,
beengood
widelysta-
used
bility, wide application range, and
in the measurement andeasy integration
monitoring of toxic with portable
and harmful devices,
gases and have
[128,130,148,163]. been
However,
widely used in since metal
the oxide-based sensors
measurement and aremonitoring
limited by theof required
toxichighandoperating
harmful temperature,
gases
they often bring additional energy consumption. In order to reduce the temperature for
[128,130,148,163]. However, since metal oxide-based sensors are limited by the required
gas detection, composite materials are used as gas sensing materials. Taking advantage of
high operating temperature,
the excellent gasthey often
sensing bring additional
properties of metal oxidesenergy
and theconsumption.
unique electrical, Inmechanical
order to
reduce the temperature for gas detection,
and thermodynamic properties of composite
graphene, thematerials are used asoxide
formed graphene/metal gas composites
sensing
materials. Takingrevealed high potential
advantage of the inexcellent
the field ofgasgassensing
sensing. properties of metal oxides and
Compared with traditional semiconductor
the unique electrical, mechanical and thermodynamic properties gas sensing of materials,
graphene, graphene/metal
the formed
oxide composite materials combine the advantages of the two materials and can produce
graphene/metalsynergistic
oxide composites revealed
effects for gas sensing,high
oftenpotential
with higherinsensitivity,
the fieldfaster
of gas sensing.
response/recovery
Compared speed,
with traditional semiconductor
and lower noise gas sensing
signal. For example, materials,
Li et al. [179] preparedgraphene/metal
the rGO/ZnO hollow ox-
ide composite materials combine the advantages of the two materials and can produce
spheres by a one-step solvothermal method, in which the ZnO hollow spheres were uni-
formly
synergistic effects for immobilized
gas sensing, on often
the surface
withofhigher
rGO nanosheets.
sensitivity,When used as
faster a material for the gas
response/recovery
sensor to detect NO2 , the composite material exhibited fast response and high sensitivity to
speed, and lower noise signal. For example, Li et al. [179] prepared the rGO/ZnO hollow
NO2 at room temperature. In another case, Liu et al. synthesized rGO/ZnO composites by
spheres by a one-step solvothermal
a redox method, and the method,
prepared gasin sensor
whichhadthea response
ZnO hollow spheres
value of 25.6% towere
5 ppmuni-
NO2
formly immobilized
with a on the surface
response ofsrGO
time of 165 and a nanosheets.
recovery time of When usedSun
499 s [180]. as eta al.
material
used PVPfor the
to assist
the synthesis
gas sensor to detect NO2, theof the composite material
composite material rGO/ZnO
exhibitednanowires. The gas-sensing
fast response and high performance
sensi-
test indicated that the rGO/ZnO composite material can respond to 500 ppb NH3 at room
tivity to NO2 at room temperature. In another case, Liu et al. synthesized rGO/ZnO com-
posites by a redox method, and the prepared gas sensor had a response value of 25.6% to
Nanomaterials 2022, 12, 982 25 of 40

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.

Figure (a) Synthetic


Figure 12. (a) Syntheticdiagram
diagramofofα-Fe
α-Fe 2O
2O 3 @graphite
3@graphite nanocomposites.
nanocomposites. (b) FE-SEM
(b) FE-SEM and and
TEMTEM images
images
of α-Fe
of α-Fe22O33@graphite.
@graphite.(c)(c)The
Theresponses
responses to to
50 ppm acetone
50 ppm of sensor
acetone based
of sensor on α-Fe
based 2O3 and
on α-Fe α- α-
2 3 and
O
Fe 2O3@graphite (with different reaction times) operated at different operating temperatures. (d) The
Fe2 O3 @graphite (with different reaction times) operated at different operating temperatures. (d) The
response of α-Fe2O3 and α-Fe2O3@graphite (with different reaction times) sensors to 50 ppm of dif-
response of α-Fe2 O3 and α-Fe2 O3 @graphite (with different reaction times) sensors to 50 ppm of
ferent gases at 260 °C. ◦(e) Reaction mechanism diagram of α-Fe2O3@graphite based gas sensor (Oδ−
different gases
− 2− at 260− C. (e) Reaction mechanism diagram of α-Fe 2 O 3 @graphite
means O , O and O2 ). Reprinted with permission from Ref. [187]. Copyright 2019 Elsevier. based gas sensor
(Oδ− means O− , O2− and O2 − ). Reprinted with permission from Ref. [187]. Copyright 2019 Elsevier.

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]

6.3. 2D TMD/Metal Oxide Composite-Based Gas Sensors


Compared with graphene with a zero-band gap, the electronic structure of 2D transi-
tion metal dichalcogenides (TMDs) almost spans the whole range of the electronic structure,
showing richer physical properties and a good application potential in gas sensing. How-
ever, TMD nanosheets are easy to form a dense stack structure in the process of forming a
conductive network, which is not conducive to the full contact between the thin sheet and
gas molecules in the conductive network, so its sensitivity and response recovery speed
at room temperature need to be improved. TMDs and other metal oxide semiconductor
materials form p-n heterojunction as a new type of electronic device, which can improve
its gas sensing performance. The unique characteristics of TMDs make the composites an
ideal choice for high-performance sensing materials at low temperatures.
In the research of gas sensing applications, SnO2 has been the most widely used n-type
semiconductor, so the combination of SnO2 and MoS2 is an effective way to improve the
sensing performance of MoS2 . For instance, Qiao et al. loaded MoS2 nanosheets onto SnO2
nanofibers by hydrothermal synthesis. Through the optimal regulation of MoS2 /SnO2
heterostructure, they achieved high-performance detection of trimethylamine at 230 ◦ C,
showing excellent sensing selectivity and long-term stability [193]. In the field of room
temperature gas sensors, Cui et al. reported a new hybrid material by decorating MoS2
nanosheets with SnO2 nanocrystals, which achieved high-performance sensing of NO2
at room temperature, and the loaded SnO2 nanoparticles could significantly enhance
In the research of gas sensing applications, SnO2 has been the most widely used n-
type semiconductor, so the combination of SnO2 and MoS2 is an effective way to improve
the sensing performance of MoS2. For instance, Qiao et al. loaded MoS2 nanosheets onto
Nanomaterials 2022, 12, 982 SnO2 nanofibers by hydrothermal synthesis. Through the optimal regulation of 28 of 40
MoS2/SnO2 heterostructure, they achieved high-performance detection of trimethylamine
at 230 °C, showing excellent sensing selectivity and long-term stability [193]. In the field
of room temperature gas sensors, Cui et al. reported a new hybrid material by decorating
the 2stability
MoS nanosheets of with
MoS2SnO nanosheets in the
2 nanocrystals, air achieved
which (Figure high-performance
13) [194]. In thesensingcomposite,
of SnO2
nanocrystals
NO acted as a strong
2 at room temperature, and thep-type
loadeddopant at the top of
SnO2 nanoparticles MoSsignificantly
could 2 , resultingenhance
in the formation
the stabilitychannels
of p-type of MoS2 nanosheets in the air (Figure
in MoS2 nanosheets. 13) [194].
In terms of In the composite,
sensing mechanism,SnO2 nano-
they believed
crystals
that SnO acted
2
as a strong
nanocrystals p-type
may dopant
be the at the
main top
gas of MoS 2, resulting
adsorption in
center, the formation
while MoS of
2 acted as a
p-type channels
conductive in MoS
channel nanosheets.
at2room In termsThe
temperature. of sensing mechanism,
close electrical they between
contact believed that
two different
SnO 2 nanocrystals may be the main gas adsorption center, while MoS2 acted as a conduc-
semiconductor materials led to the formation of charge transfer and charge depletion
tive channel at room temperature. The close electrical contact between two different sem-
layer. Because the work function of SnO2 (5.7 eV) was larger than that of MoS2 (5.2 eV),
iconductor materials led to the formation of charge transfer and charge depletion layer.
electrons are transferred from MoS to SnO2 , resulting in a depletion layer and a Schottky
Because the work function of SnO2 (5.7 2eV) was larger than that of MoS2 (5.2 eV), electrons
barrier
are (Figurefrom
transferred 13f).MoS
These electron depletion regions can be connected to each other on
2 to SnO2, resulting in a depletion layer and a Schottky barrier
the surface
(Figure of MoS
13f). These 2 nanosheets
electron depletionand act as
regions cana be
passivation
connected layer
to eachtoother
prevent
on thethe interaction
sur-
between oxygen and
face of MoS2 nanosheets and actMoS , thus enhancing the stability of MoS
2 as a passivation layer to prevent the interaction 2 nanosheets
betweenin dry air.
The fabricated gas sensor exhibited high sensitivity, excellent repeatability,
oxygen and MoS2, thus enhancing the stability of MoS2 nanosheets in dry air. The fabri- and excellent
selectivity
cated to NO
gas sensor in the
exhibited
2 actual
high dry air
sensitivity, environment,
excellent and
repeatability,the detection
and excellent limit could reach
selectiv-
ity to NO
0.5 ppm. 2 in the actual dry air environment, and the detection limit could reach 0.5 ppm.

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]

6.4. Other 2D Material/Metal Oxide Composites-Based Gas Sensors


As an important wide band gap and semiconductor gas sensing material with a special
layered structure, MoO3 is easy to form metal phase MoS2 in the process of reacting with
H2 S, which makes MoO3 have unique response characteristics to H2 S gas. However, single-
phase MoO3 has some shortcomings, such as high working temperature and poor limit
detection ability [201,202], so it needs to be compounded with other materials to improve
its gas-sensing performance. For example, Gao et al. used graphene as a sacrificial template
and prepared porous MoO3 /SnO2 composite nanosheets with n-n heterostructure by hy-
drothermal method, and studied their gas sensing properties, as shown in Figure 14 [203].
The TEM characterizations indicated that the composite is a porous structure composed
of MoO3 and SnO2 nanoparticles, and the lattice distortion at the interface indicates the
existence of MoO3 -SnO2 n-n heterojunction (Figure 14a,b). The gas-sensing performance
test showed that compared with SnO2 nanosheets, the introduction of MoO3 and the forma-
tion of heterojunctions lead to the change of energy band structure and carrier separation
transfer rate, and the optimum operating temperature of MoO3 /SnO2 nanosheet sensor is
lower and the sensitivity is higher. The detection limit of MoO3 /SnO2 nanoparticles for
H2 S was as low as 100 ppb at 115 ◦ C, and the response and recovery times were 22 s and
10 s, respectively (Figure 14c–e). The excellent gas-sensing performance of the MoO3 /SnO2
composite nanosheet gas sensor was attributed to the following aspects.
Firstly, in the composites, the n-n heterojunction formed at the interface between MoO3
and SnO2 , and the charge accumulation layer and charge consumption layer are formed on
both sides of the heterojunction, respectively, thus forming an internal electric field and
hindering the free transport of free electrons in the semiconductor. The sensing mechanism
of the composite material is shown in Figure 14f,g. When the material is exposed to air, the
electrons in the semiconductor MoO3 and SnO2 conduction bands adsorb the oxygen in the
air to the surface of the semiconductor gas sensing material to form the adsorbed oxygen.
This leads to the loss of electrons in the conduction bands of semiconductors MoO3 and
SnO2 (Figure 14f). On the energy band diagram, the energy bands of semiconductors MoO3
and SnO2 bend upward, and the barrier height increases. In the atmosphere of H2 S, when
H2 S reacts with the adsorbed oxygen on the surface of semiconductor sensitive materials,
the electrons captured by adsorbed oxygen are released back into the semiconductor MoO3
and SnO2 conduction bands, which leads to a significant decrease in the barrier height at
the n-n heterojunction and a significant decrease in the resistance of the sensor (Figure 14g).
The electrical conductivity of composite semiconductor materials is inversely proportional
to the barrier height, and the barrier height can effectively control the gas sensing response
characteristics of sensitive materials with heterojunctions. Therefore, a large change of
barrier height has a significant contribution to the H2 S performance of the MoO3 /SnO2
gas sensor. Secondly, the large specific surface area of MoO3 /SnO2 composite nanosheets
means that when exposed to H2 S gas, the MoO3 /SnO2 sensor with a large specific surface
area can contact more H2 S gas molecules and react with them, resulting in greater resistance
change and higher sensitivity response. Thirdly, the composite has high porosity, and the
erojunctions. Therefore, a large change of barrier height has a significant contribution to
the H2S performance of the MoO3/SnO2 gas sensor. Secondly, the large specific surface
area of MoO3/SnO2 composite nanosheets means that when exposed to H2S gas, the
MoO3/SnO2 sensor with a large specific surface area can contact more H2S gas molecules
Nanomaterials 2022, 12, 982 31 of 40
and react with them, resulting in greater resistance change and higher sensitivity re-
sponse. Thirdly, the composite has high porosity, and the larger pore volume can not only
promote the adsorption of H2S gas molecules to the sensitive material surface more
larger pore volume can not only promote the adsorption of H2 S gas molecules to the
quickly, but also provide
sensitivea material
shorter surface
gas conduction path
more quickly, butand
alsopromote
provide a the gasgas
shorter molecules topath
conduction
break away from the andsensitive
promote the material surface
gas molecules andaway
to break thenfrom
achieve the ideal
the sensitive statesurface
material of rapid
and then
achieve
response and rapid recovery. the ideal state of rapid response and rapid recovery.

Figure 14. (a,b) Low-magnification TEM images and


Figure 14. (a,b) Low-magnification TEMHRTEM
images andimages
HRTEM of images
MoSn-S nanoflakes.
of 2MoSn-S Red Red
2 nanoflakes.
circles in (b) shows the lattice distortions at the interfaces between
circles in (b) shows the lattice distortions at the interfaces between MoO3 and SnO32. (c) Sensor MoO and SnO 2 . (c)
re-Sensor
sponses of as-preparedresponses
samples of as-prepared
as a functionsamples as a function
of different of different
operating operating temperatures
temperatures to 10 ppm to of10Hppm2S
of
H S concentration. (d) Typical sensor responses of SnO , MoSn-S1,
concentration. (d) Typical sensor responses of SnO2, MoSn-S1, MoSn-S2, and MoSn-S3 toward 100
2 2 MoSn-S 2 , and MoSn-S 3 toward
100 and 500 ppb of H2 S gas at optimal working temperature, and (e) response and recovery times
and 500 ppb of H2S gas at optimal working temperature, and (e) response and recovery times curve
curve of MoSn-S2 NFs to 10 ppm of H2 S at 115 ◦ C. (f,g) Diagram of energy band structure of
of MoSn-S2 NFs to 10 ppm of H2S at 115 °C. (f,g) Diagram of energy band structure of MoSn−nano-
MoSn−nanocomposites in (f) air and (g) H2 S. Reprinted with permission from Ref. [203]. Copyright
composites in (f) air and (g) H2S. Reprinted with permission from Ref. [203]. Copyright 2019 ACS.
2019 ACS.
Nanomaterials 2022, 12, 982 32 of 40

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.

7. Conclusions and Future Perspectives


In this paper, we summarize the research progress of gas sensors using 2D materials,
metal oxides, and their composites as sensitive materials. The gas sensing mechanism,
main factors affecting sensing performance, and the applications of various gas sensors are
presented and discussed in detail. It can be concluded that the 2D material/metal oxide-
based gas sensor can efficiently identify and detect toxic and harmful gases. Compared
with pure metal oxide semiconductors, composite materials have higher carrier rates, larger
high mechanical strength, and large specific surface area, and the synergistic effect of the
two components can further enhance the gas sensing performance. The -OH, -O, and
other functional groups on the surface of 2D materials not only provide chemical bonds
to form composite metal oxide materials during the composite process of the material,
but also give more active sites for the gas sensing process, thereby further improving the
gas sensing performance. In addition, the composite materials can effectively reduce the
working temperature.
Although nanomaterial-based gas sensors have made great progress in the past few
decades, the operating temperature of metal oxides is too high, and the selectivity of
2D materials is still unsatisfactory. In the future, effective strategies such as building
composite structures are highly needed to improve the selectivity, reduce the operating
temperature, and improve the sensitivity and other properties. In addition, the research on
the combination of metal oxides with 2D materials is still at an early stage, and its sensing
mechanisms of the composite-based gas sensors should be further studied. Only when the
mechanism and process are clear, the preparation and assembly of nanomaterial-based gas
sensors can be achieved purposefully. In addition, it is necessary for researchers to develop
new design strategies to further optimize metal oxide nanomaterials with 2D nanomaterials
to make them more suitable for gas sensing. Finally, it is expected that facile assembly and
fabrication processes will be developed to enable batch fabrication of gas sensors with high
stability, selectivity, sensitivity, reproducibility, and quick response in the future.

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

Conflicts of Interest: The authors declare no conflict of interest.

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]

You might also like