0% found this document useful (0 votes)
52 views33 pages

CCA Project Group No 11

Uploaded by

Sachith Kuger kg
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
52 views33 pages

CCA Project Group No 11

Uploaded by

Sachith Kuger kg
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 33

BMS Institute of Technology and Management

(An Autonomous Institution, Af liated to VTU, Belagavi)

Avalahalli, Doddaballapura Main Road, Bengaluru –


560064

DEPARTMENT OF CHEMISTRY

2024-2025/ ODD SEM


REPORT

on

Metal Oxide Gas Sensors for Detecting


NO2 in Industrial Exhaust Gas: Recent
Developments
{GROUP RESEARCH, As a part of Comprehensive
Continuous Assessment}
fi
Submitted by,

R MANAS REDDY
SAMYUKTHA G GOWDA
RAHUL KUMAR MANDAL
SAMRIDH MAHAJAN
SACHITH KUGER

{CSE/CSE-3/2024-25}

Submitted to,

Dr. SWETHA G A
Assistant Professor, Department of Chemistry

Chemo Resistive Gases

A chemo resistive gas sensor (CGS) is a device that detects gases by


measuring the changes in electrical resistance of a material when it
interacts with the gas. CGSs are low-cost, low-power, and highly sensitive.
They can detect a variety of gases and volatile organic compounds
(VOCs).
2.1. Metal oxide (MO)
MOs such as ZnO, NiO, CuO, SnO₂, and In₂O₃ widely used in the detection
of toxic and VOC gases possess high sensitivity, low production cost, and
thermal stability. For example, response of 73.3% for 750 ppb H₂S towards
ZnO sensors prepared with the sol-gel method by room temperature with a
lowest detection limit of 30 ppb is achieved. There are slow gas adsorption-
desorption and poor selectivity at room temperature.
To enhance performance, scientists have engineered MOs. Wang et al.
fabricated ZnO@In₂O₃@ZnO core-shell microspheres that lowered the
working temperature by 100°C and improved sensitivity 30 times with
increased ethanol selectivity. Similarly, RF sputtering-generated CuO/ZnO
composites exhibited a 42.5% response at 200°C to 1 ppm H₂S while pure
ZnO showed none. The Schottky barrier model is one of the common
explanations of the mechanism of gas sensing in MOs. Oxygen from air
adsorbs on MOs surfaces and captures the electrons forming adsorbed
oxygen species O₂⁻, O⁻, O²⁻. As a result, an electron depletion layer is
formed over n-type semiconductors with an increase in surface barriers.
When such reducing gases react with oxygen adsorbed, the backflow of
electrons into MO diminishes EDL and reduces the surface barrier,
therefore lowering resistance.

2.2. Conducting polymers


Conductive polymers include polyaniline (PANI), polypyrrole (PPy),
polythiophene (PTh), among others, which have outstanding electrical
properties due to their π-electron framework, making them promising
candidates for room-temperature gas sensors. The chemical and physical
properties of these conductive polymers can be tuned by various
substituents.

For example, Saaedi et al. deposited PANI on a substrate using spin


coating and found it had a 49% gas sensitivity response to 100 ppm NH₃ at
room temperature with an 8s/13s response/recovery time. NH₃ sensing in
PANI involves protonation (PANI acts as a p-type semiconductor) and
deprotonation (increased resistance).

Similarly, PPy lms deposited through gas-phase polymerization displayed


a 115% response to 350 ppm NH₃ with a 10s/44s response/recovery time.
The PPy sensing is doping and proton transfer dependent. Park et al.
prepared porous PTh lm by plasma jet polymerization, and it revealed
high selectivity and sensitivity toward NO₂ up to 84% for 2 ppm.

2.3. Carbon-based materials


Carbon-based materials, such as carbon nanotubes (CNTs) and graphene,
are widely used in gas sensors due to their unique properties. CNTs,
including single-walled (SWCNT) and multi-walled (MWCNT), are tubular
structures derived from two-dimensional graphene. Their high aspect ratio,
specific surface area, electrical and chemical properties, and hollow
structure enhance sensitivity and lower operating temperatures.

For instance, Lee et al. deposited raw CNTs on a microheater to sense H₂.
The localized Joule heating reduced power consumption and facilitated
sensor miniaturization. Although raw CNTs were not initially responsive to
fi
fi
H₂, the local heating enabled reactions that resulted in a 1.82% response to
10% H₂ with response/recovery times of 39s/35s. However, raw CNTs have
low sensitivity, are insensitive to some gases, and agglomerate, which
makes their usage limited.

To address these issues, CNTs are functionalized via covalent or non-


covalent methods. Covalent functionalization involves attaching organic
molecules or metal particles to CNT surfaces. For example, Dong et al.
functionalized SWCNTs with nitric acid to add carboxylic groups, enabling
detection of 10 ppm CO through hydrogen bonding. Non-covalent
functionalization involves combining CNTs with conductive polymers or
metal oxides. Kumar et al improved the gas sensitivity of SWCNT to NO₂ by
doping it with polyethyleneimine (PEI), which increased the response to
37% for 20 ppm NO₂ at room temperature. Functionalization introduces
defects that improve gas adsorption and charge transfer, enhancing sensor
performance. Graphene is characterized as a 2D material consisting of sp²-
hybridized carbon atoms with superior gas adsorption, a large surface area,
a high mobility of electrons, and good mechanical strength, putting it on the
list for use as a gas sensor at RT. However, graphene raw material does
not possess dangling bonds, implying a reduced interaction with gas
molecules that relies only on van der Waals forces of attraction.

To increase sensitivity, researchers modify graphene. For instance, Huang


et al. added flavin mononucleotide sodium salt (FMNS) as a stabilizer and
dopant, which allowed them to achieve a 2.8% response to 10 ppm NH₃ at
RT and detect concentrations as low as 1.6 ppm. FMNS offered active sites
for the adsorption of NH₃, enhancing gas-sensing performance.

More applications for sensors have been in graphene derivatives. For


example, graphene oxide and reduced graphene oxide. Graphene oxide
has oxygen functionalities of carboxyl, hydroxyl, etc., so is much easier to
make, depending on methods like Hummers or Brodie. Narrow band gap
GO synthesized using an improved Hummers method by Park et al.
displayed a 32% response to 5 ppm NO₂ at 150°C with a LOD of 0.21 ppb,
well below the annual EU limit of 21 ppb.

Both CNTs and graphene are considered p-type semiconductors due to


doping by O₂ and H₂O molecules in the air. Their gas-sensing mechanism
involves resistance changes: reducing gases donate electrons, reducing
hole carriers and increasing resistance, while oxidizing gases lower
resistance by accepting electrons.

NO₂ Gas Sensors:


ZnO:-
The most widely used material for NO₂ detection is ZnO, an n-type
semiconductor with a wide band gap (~3.37 eV), because of its chemical
stability, non-toxic nature, and excellent electrical properties.
The morphology of metal oxides greatly affects their gas-sensing
performance. ZnO has been explored in different morphologies, including
nanorods, nanosheets, nanoflowers, and hollow spheres, to improve NO₂
response.

For instance, Godse et al. synthesized hexagonal ZnO nanorods, which


converted into nanotubes at higher temperatures. These nanorods showed
excellent sensitivity toward NO₂, with a 70% response to 5 ppm NO₂ and a
response/recovery time of 16/200 s.

Nanosheets, in a large specific surface area with a thin structure, have


outstanding performance. Nguyet et al. reported nanosheets of ZnO
synthesized via the hydrothermal method (15 nm in thickness, BET 75 m²/
g). That material gave a response in the range 76-0.5 ppm NO at 200°C
and exhibited a detection limit down to 3 ppt.

Porous nanosheets enhance the gas sensitivity. Choi et al. synthesized


porous ZnO nanosheets (16% porosity, ~60 nm pore size) via solvothermal
methods. These showed strong NO₂ sensitivity, with a response of 74.68 to
10 ppm NO₂ at 200°C and excellent long-term stability over 10,000
seconds. The porous structure improved gas diffusion and provided a high
surface area (11.51 m²/g), increasing active adsorption sites and boosting
performance. The 3D ZnO nanoflowers comprise porous, flat, net-like
petals with a thickness of ~21 nm and pore sizes of ~10 nm. These sensors
reached a maximum response of 23.3 to 100 ppm NO₂ at 180°C with a
response/recovery time of 4/143 s. Their structure exhibited excellent
mechanical and response stability under ultrasonic oscillation.

Hollow ZnO structures, with double adsorption layers, improve gas


sensitivity due to the availability of more active sites. Minh et al. prepared
hollow ZnO nanoparticles (~30 nm thick) using glucose as a soft template
during hydrothermal processing, followed by calcination to remove the
template. These hollow ZnO sensors responded to 15.3 to 1 ppm NO₂ at
200°C with response/recovery times of 194/173 s and a low detection limit
of 250 ppb.

Preparation: -
Other methods aside from hydrothermal to increase ZnO sensitivity have
been investigated.
1. **Aerosol-Assisted CVD**: Sanchez-Martin et al. used laser interference
lithography and UV laser to structure ZnO with vertically oriented crystal
grains and nano-pattern with a diameter of 800 nm. The sample exhibited
high surface defects that increased carrier mobility, hence boosting the
surface-to-volume ratio, thus displaying a 600% response to 4 ppm NO₂ at
175°C and detect at room temperature.

2. SILAR Method: Ganapathi et al. synthesized ZnO lms using the SILAR
method, which involved repeated immersion of the substrate in cation/anion
precursors and rinsing. This method ensured uniform deposition and
controlled stoichiometry. The ZnO lm sensor showed a 249% response,
short response/recovery time (11/44 s), and a 500 ppb LOD for 20 ppm
NO₂ at 200°C.

3. **In-Situ Thermal Decomposition Deposition (TDD)**: Lou et al. directly


grew ZnO nanoarrays on substrates by heating zinc acetate dihydrate
(C₄H₆O₄Zn·2H₂O) with a glass substrate and Au/Ti electrodes in an electric
furnace. This method eliminates the transfer step, reducing cost and time
while improving material-electrode contact for ef cient sensor fabrication.

Doping:
Doping has heavily improved the gas-sensing properties of ZnO via
modi cation of oxygen vacancies and resistance. Researchers have
heavily investigated doped ZnO for greater gas sensitivity:
fi
fi
fi
fi
1. **Co-Doped ZnO**: Co-doped ZnO samples were prepared by co-
precipitation by Kamble et al. The samples showed superior gas sensitivity
and response times. At 200°C, the sensor showed a 286% response to 100
ppm NO₂ with a response/recovery time of 3/319 s.

2. **Al-Doped ZnO**: Ramgir et al. doped ZnO with Al, introducing defects
that increased carrier concentration, mobility, and oxygen coverage. At
230°C, 1.8 at% Al-ZnO showed a 6.2 response to 4 ppm NO₂, with a
response/recovery time of 18/65 s and a detection limit of 0.5 ppb.

3. **Ni-Doped ZnO**: Kamble et al. synthesized Ni-ZnO and discovered


that 2 at% Ni-ZnO performed the best. At 200°C, it showed a 356%
response to 100 ppm NO₂, with a response/recovery time of 0.67/136.19 s.
The porous spherical morphology increased surface area and chemical
reaction interfaces, enhancing gas sensitivity.
Composites:
Composite materials are extensively researched to improve gas-sensing
performance, particularly ZnO doped with carbon-based materials such as
graphene, rGO, and CNTs.

1. **ZnO/Graphene**: Hassan et al. prepared 3D graphene aerogel hybrid


with ZnO recycled NPs from battery wastes. It offered a high surface area
and fast electronic pathways resulted in a fast response 17.69 to 5 ppm
NO₂ at RT having response/recovery time, 13.15 s/7.7 s.

2. ZnO@rGO: Gao et al. successfully wrapped ZnO nanorods with rGO,


which signi cantly improved gas sensing at room temperature. The
composite presented a response of 6.77 to 10 ppm NO₂, with response/
recovery times of 150/113 s and an LOD of 100 ppb. It demonstrated
excellent stability over 60 days, with minimal performance deviation, and
maintained good sensitivity (5.40) even at 90% humidity.
fi
Nio:-
NiO is a p-type semiconductor with a band gap of 3.6–4.0 eV. It is valued
for its high chemical stability and excellent electrical properties, which make
it suitable for high-performance sensors. Wilson et al. [131] have
synthesized ultra-thin NiO lms using CVD and studied the effect of lm
thickness on NO₂ gas response. The optimum working temperature was
125°C, improving low-temperature NO₂ detection and the lifetime of the
device. Films that are relatively thinner, especially near the Debye length,
exhibited better gas sensitivity. New studies on NiO towards NO₂ sensing
majorly relate to its application as composite with other materials.

Preparation:-
The research on detecting low gas concentrations with high sensitivity and
selectivity has greatly improved. Urso et al. [132] prepared nanoporous NiO
lms through chemical bath deposition and thermal annealing, which gave
a high response to 140 ppb NO₂ with a detection limit of 20 ppb. The self-
assembled nanoporous network improved gas adsorption and diffusion. At
fi
fi
fi
sub-ppm concentrations, the response of the sensor at room temperature
was greater than that at 150°C, and with better selectivity and stability.

Gomaa et al. [133] synthesized NiO lms through chemical spray pyrolysis
to create a honeycomb-like porous structure, very suitable for gas
adsorption. 5 min of deposition led to the optimal response, 57.3%, to 20
ppm NO₂ at 200°C. The sensor showed very good reversibility, linearity,
and a response of 40% at 10 ppm NO₂. This approach allows fast
production of NiO sensors, which can be used practically.

Doping:-
Doping is one of the main methods to improve metal oxide performance.
Kampitakis et al. [134] used radio frequency sputtering to dope Al into NiO,
forming Al-NiO lms (52–167 nm). The 5.0 at% Al-NiO sensor exhibited a
high response of 271% to 200 ppb NO₂ at 200°C, but the response time
was slow, at 26 minutes, due to low mobility and strong NO₂
electrochemical bonding. Al³⁺ doping enhances the concentration of p-type
carriers and conductivity. Despite the gas response improvement, its poor
response time will require additional study on sensors with fast response/
recovery speeds to match sensitivity.

Composites: -
Combining conductive polymers with metal oxides enhances gas sensor
sensitivity and selectivity. Zhang et al. [135] created a PPy-NiO sensor for
room-temperature NO₂ detection. At a 100:1 molar ratio, PPy-NiO showed
a 45% response to 60 ppm NO₂, 30× higher than pure NiO, with a low LOD
of 49 ppb. The improved performance was due to charge transfer from PPy
fi
fi
to NiO, creating depletion and accumulation layers that ampli ed resistance
changes upon NO₂ adsorption.

Liu et al. [136] fabricated a ZnSe/NiO sensor that exhibited a 96.47%


response to 8 ppm NO₂ at 140°C, against 19.65% for pure NiO, with LOD
of 8.91 ppb. This improvement in sensitivity was attributed to a p-p
homojunction and hole accumulation layer in ZnSe, reducing resistance
and boosting performance.

CuO: -
CuO, a p-type semiconductor with a narrow band gap (1.35 eV), is low-
cost, stable, and highly catalytic, making it suitable for detecting gases like
NO₂, H₂, ethanol, and acetone [137–140]. Oosthuizen et al. [141] developed
CuO nanosheets that demonstrated excellent NO₂ sensing at room
temperature with a gas response of 14.5 to 20 ppm NO₂. However, CuO's
room-temperature sensitivity and response/recovery time require
improvement [142]. Researchers are investigating advanced preparation
fi
techniques [47,138,143,144], doping [50,145], and composites [146–148]
to improve the gas-sensing property of CuO.

Preparation: -
Researchers use several methods for preparing CuO-based gas sensors,
including hydrothermal, chemical vapor deposition (CVD), pulsed laser
ablation, spray pyrolysis, and plasma spraying.
Nanda et al. [47] prepared CuO sensors using CVD and showed that
surface roughness, grain size, and dangling bond density greatly affect gas
response. The higher surface roughness and smaller grain size lead to
better responses but might cause longer recovery time due to more grain
boundary traps. The response of 220.92% for 4.5 ppm NO₂ at 200°C in
their CuO sensor was reported with a response/recovery time of 50.9/124.1
s and a LOD of 300 ppb.

Al-Jumaili et al. [143] synthesized CuO nanoparticles through PLA and


investigated the impact of annealing time. The optimal response (90% to
NO₂ at RT) was obtained after 2 h of annealing due to the increase in
porosity and a wide depletion layer with a response/recovery time of
3.78/1.04 s. More annealing time reduced the porosity and performance.

Khot et al. [144] utilized spray pyrolysis for the deposition of CuO lms.
They reported a best gas response of 56% to 100 ppm NO₂ at 200°C with a
rough surface 0.15 M precursor concentration. This sensor proved high
selectivity for NO₂ with fast response (20.57 s), however recovery time was
longer (3.92 min).

Doping: -
Doping and loading of noble metals are very common strategies to improve
CuO's performance.
fi
Goyal et al. [50] prepared Zn-doped CuO by the co-precipitation method.
Higher concentrations of Zn doping changed the morphology of CuO from
octahedral to spherical and decreased the photoluminescence intensity,
thus suppressing the recombination of electrons and holes and increasing
oxygen adsorption. Zn-CuO demonstrated higher NO₂ response with
increased work function and conductive carriers. At 150°C, Zn-CuO
demonstrated better gas response than pure CuO, especially for 1 ppm
NO₂.

Chen et al. [145] synthesized CuO nanorods with Au nanoparticle loading


using a one-pot hydrothermal route. Though the structure was maintained
by Au-CuO nanorods, its band gap was reduced from 1.32 eV to 1.22 eV
because of oxygen vacancy generated upon Au loading, which further
acted as active sites for NO₂ adsorption leading to increased sensitivity. It
can give 3% response toward 20 ppm NO with a response/recovery time of
8/176 s at room temperature but 2 mol% Au-CuO in terms of high cost to
industry, thus requiring some cheap alternatives.

Composites: -
During the past few years, CuO has been combined with reduced graphene
oxide (rGO) and black phosphorus (BP) to enhance gas sensing
performance.

Jyoti et al. [146] synthesized CuO nano owers, nanobricks, and


nanochains combined with rGO. CuO nano ower/rGO showed maximum
sensitivity towards NO₂ at room temperature. The response of NO₂ 20 ppm
showed a value of 58.1%. It even surpasses other morphologies.
Enhanced performance has been achieved by nding out the optimal rGO
content at 25 wt %. This even resulted in prolonging recovery time due to
intense NO₂-rGO interaction. CuO/rGO shows excellent selectivity of NO₂.

Bai et al. [147] reported ultra-thin CuO nanosheets combined with rGO.
The composite sensor showed a gas response of 20.6% to 50 ppb NO₂,
much greater than pure CuO. CuO/rGO achieved a 400.8% response to 5
ppm NO₂ at room temperature with a response time of 6.8 s and a
fl
fl
fi
detection limit of 50 ppb because of its mesoporous structure, high surface
area, and fast carrier transfer.

Wang et al. [148] developed a CuO/BP composite gas sensor using


functionalized BP. The CuO/BP sensor had high sensitivity, with a 20.7%
response to 100 ppm NO₂, a 1.1% response to 10 ppb, and fast response/
recovery times (4/56 s). However, its performance declined under high
humidity due to reduced NO₂ adsorption. The enhanced performance was
attributed to oxygen vacancies, improved adsorption, and ef cient electron
transport via heterojunctions.

Overall, these studies show the future possibility of CuO-based composites


in the sensitive detection of room-temperature NO₂, although humidity
resistance remains challenging.

SnO2: -
SnO₂ is a wide-bandgap (3.6 eV) n-type semiconductor material, widely
applied in gas sensors, owning advantages such as high stability, low cost,
ease in synthesis, good light transmission, and high electron mobility up to
160 cm²/V·s. Niu et al. [152] have designed patterned SnO₂ nanoarrays on
MEMS chips via a one-step precipitation method. The sensor that resulted
showed good NO₂ sensitivity, responding 150.9 to 10 ppm NO₂ at 150°C,
and with a detection limit as low as 30 ppb, which opens up the possibility
of highly sensitive NO₂ sensors.

Morphology: -
fi
Morphology control has a signi cant impact on material properties,
especially on the performance of gas sensors. Larger speci c surface area
structures offer more active sites for adsorption and reaction, while
conductive channels enhance electron mobility and response/recovery
speeds.

Recent developments in SnO₂ gas sensors include:


1. Mesoporous SnO₂ with Biological Morphology:
Li et al. [154] made use of the templates to form 3D mesoporous SnO2 via
the cannabis stems. Its sensor showed high sensitivity of NO2 at room
temperature: 35.83% response to 100 ppm NO2, with 2.67 s response time
and LOD = 10 ppb with stability over 220 days. Such performance was
described for this sensor due to porous channel structure, exposed (110)
crystal planes, high surface area, and residual carbon p-n heterojunction.

2. **SnO₂ Nanotoast Structure**:


Li et al. [46] prepared a porous "nanotoast" structure at the particle size of
7 nm and surface area 79.94 m²/g. Room-temperature sensor showed
excellent sensitivity of 94 % response to 10 ppm NO₂ with high selectivity.
At 50 °C, it showed shortening of response time into 37.2 s. Its oxygen
vacancies, porosity, and particle arrangement enhance the adsorption,
reaction, and carrier transport abilities and therefore indicate potential for
application in monitoring NO₂ levels at school entrances and parking
garages. 3. **Porous Hollow SnO₂ Microspheres**

Zhou et al. [155] used electrospray deposition to create SnO₂ quantum wire
shells forming porous hollow microspheres (3 ± 0.5 µm diameter). At room
temperature, the sensor showed a 10% response to 10 ppm NO₂, with a
response/recovery time of 17/65 s and a theoretical LOD of 26 ppb. This
fi
fi
innovation highlights the potential for low-cost, high-performance room
temperature gas sensors.
These studies underscore the role of morphology in optimizing gas sensor
sensitivity, response time, and stability.

Preparation: -
To improve the sensitivity of SnO₂ gas sensors, different preparation
methods have been studied. Aqeel et al. [48] prepared SnO₂ by a soft
template method, which resulted in a mesoporous polycrystalline structure.
At a calcination temperature of 400°C, the sensor response to 1 ppm NO₂
was found to be 407%. Other methods include femtosecond laser
irradiation and pulsed ultraviolet radiation for further sensitivity
improvements.

Devabharathi et al. [49] synthesized a mesoporous SnO₂ ink and inkjet


printed it onto a silicon substrate. The SnO₂ lm obtained by printing
showed well-connected pores, and the gas response was 11,507% to 5
ppm NO₂ at 175°C. The response to 20 ppb NO₂ reached 81% at 125°C.
This was attributed to the mesoporous structure, which provided excellent
electron transport properties, high surface area, and structural exibility.

Doping: -
Doping remains a highlighted area in gas sensor study. Noble metals such
as Pd and rare earth metals such as Y, Er, S, N, Ni, and Sb can be doped
into SnO₂ to improve its performance in detection of NO₂. Among the
dopants are better sensitivity, response speed, selectivity, repeatibility, and
lower detection limits.
fi
fl
For example, Choi et al. [158] embedded Pd in SnO₂ using ame chemical
vapor deposition, achieving a higher NO₂ response at room temperature
and shorter response times. Sohal et al. [51] doped SnO₂ with Y, reaching a
1445% response to 1 ppm NO₂ at room temperature. Li et al. [160] added
sulfur to SnO₂, improving the response to 100 ppm NO₂ to 57.38% at room
temperature. Ni-doped SnO₂ has a high 850% response to 3 ppm NO₂.

These results are important as rare earth elements may be used to further
enhance the performance of the gas sensor. Optimization in terms of
morphology and composition could lead to even more sensitive, high-
performance sensors, especially at room temperature.

Composites:-
fl
In recent years, SnO₂-based composite materials have been developed,
with reduced graphene oxide (rGO) being the most studied. For instance, Li
et al. [179] combined rGO and SnO₂ nanospheres to create a NO₂ sensor,
achieving a 53.57% response to 3 ppm NO₂ at 125°C, with fast response/
recovery times. Yan et al. [181] introduced silver (Ag) into rGO/SnO₂
aerogel, enhancing electron mobility and reducing the sensor's operating
temperature, making it suitable for room temperature NO₂ detection.
Choudhari et al. [178] used ultrasound to ensure uniform SnO₂ deposition
on rGO, achieving a 99.9% response to 100 ppm NO₂ at 150°C. Wang et
al. [177] developed a humidity-insensitive rGO/SnO₂ NO₂ sensor by
creating a super-hydrophobic structure, preventing the sensor's
performance from being affected by humidity. These advances show the
potential of rGO/SnO₂ sensors for use in environments with high humidity,
such as mines and petrochemical areas.
WO₃: -

WO₃ is an n-type semiconductor that has stability, low cost, non-toxicity,


and fast electron mobility. It is thus very ef cient for gas sensing. The ability
of WO₃ to change valence states to W⁶⁺ makes it sensitive to both oxidizing
and reducing gases. For instance, Khudadad et al. [187] reported that the
WO₃ heat-treated with a grape-like morphology and high surface area
showed 392% response to 150 ppm NO₂ with rapid responses and low
detection limit.

WO₃ has been doped with elements such as Pt, Au, and Pd for enhancing
its performance. Shen et al. [194] reported that the nanosheets of WO₃
doped with Au showed higher sensitivity and a faster response than the
pristine WO₃. Humidity, however, degrades the performance of WO₃
sensors. Ghosh et al. [196] demonstrated that highly oxidized WO₃ lms
were less sensitive to humidity, and high-temperature aging at 90°C and
90% RH helped reduce the impact of humidity on gas sensitivity, which
may be a potential method for developing humidity-insensitive NO₂
sensors.
fi
fi
Morphology: -
WO₃ nanomaterials of various structures have been designed by
researchers, including 0D (nanoparticles), 1D (nano bers), 2D
(nanosheets), and 3D (nanospheres, nano owers, nanomesh, biomorphic
fl
fi
hierarchical structures).
For instance, WO₃ nanoparticles with

two crystal forms synthesized by Matovic et al. [197] revealed a response


of 450% at 1 ppm NO₂ at room temperature. Zhang et al. [198] synthesized
WO₃ nano bers at a heating rate of 10°C/min, which showed a high
sensitivity of 101.3% to 3 ppm NO₂ at 90°C. Wang et al. [199] synthesized
WO₃ nanosheet arrays that had a high response of 460% to 10 ppm NO₂ at
100°C.

Zhao et al. [200] fabricated hollow WO₃ spheres, showing a response to


sub-ppm NO₂, while Liu et al. [202] made WO₃ nanomesh that detected 50
ppb NO₂ at 160°C. Lv et al. [203] applied hemp stems as a template to
prepare hierarchical WO₃ structure with ultra-high response of 71.07%
toward 100 ppm NO₂. These various morphologies of WO₃ nanomaterials
contribute to improved gas sensitivity, due to their large surface areas and
oxygen vacancies.

Modi cation: -
Ionic liquids (ILs) have been incorporated into the modi cation of metal
oxides, for example, WO₃, to enhance gas sensing. Zhang et al. [204]
added 1-butyl-3-methyl-tetra uoroborate ionic liquid to WO₃ preparation,
fi
fi
fl
fi
thus improving the dispersion and doubling the speci c surface area of the
WO₃ prepared. This modi ed WO₃ displayed a 19.5% response at 10 ppm
NO₂ and a low detection limit of 100 ppb. Parellada-Monreal et al. [205]
coated WO₃ lms with 2D DLIP, making a porous structure that reduces the
LOD to 10 ppb, which is half of the untreated sample. The laser-induced
defects improved the gas adsorption capacity of the sensor and its
performance.

Composites:-
Recent studies have focused on composite materials of WO₃ with materials
such as CNTs, rGO, and MXenes to enhance gas sensing. Hung et al.
[207] incorporated multi-walled carbon nanotubes (MWCNTs) into WO₃ and
analyzed their effect on NO₂ sensing. The MWCNTs were well dispersed
within the WO₃ matrix, improving conductivity. Optimal NO₂ sensor
performance was achieved with 0.5 wt% MWCNT, which provided the
highest gas response of 18% and fast response/recovery times of 87/300 s
at 150°C. The sensing properties of this composite material were improved
compared to pure WO₃.
fi
fi
fi
MO/MO Composites:-
Recent studies have investigated composite metal oxide (MO/MO)
materials for improved NO₂ gas sensing. For instance, Navale et al. [9]
synthesized a composite of CuO nanoparticles and ZnO nanorods, which
exhibited high sensitivity to NO₂ with a 96% response at 150°C for 100 ppm
NO₂. The ZnO/CuO composite created a p-n heterojunction that enhanced
conductivity upon NO₂ exposure. Similarly, Guo et al. [35] syn

thesized SnO₂/ZnO composites, which signi cantly increased NO₂


response compared to pure SnO₂, with a high response to sub-ppm NO₂ at
room temperature. The composite also demonstrated good selectivity to
NO₂. Jin et al. [223] developed NiO/ZnO composites, where adjusting the
NiO content allowed selective detection of either oxidizing gases (NO₂) or
reducing gases (CO), depending on the p-n or p-p junction behavior.
fi
Conclusions: -

In summary, this article reviewed NO2 gas sensors based on metal oxides.
Firstly, the hazards brought by industrial waste gas and the different
detection methods of the gas were discussed. In industrial waste gas, NO2
was an oxidizing gas, which was considered to be the most dangerous gas
among toxic gases. At the same time, because of the advantages of high
sensitivity, selectivity, stability and fast response/recovery speed, the
chemical resistance sensor is used as a NO2 gas sensor. Then, different
sensitive materials of chemical resistance sensors were discussed,
including
metal oxides, conductive polymers, and carbon-based materials. Finally,
the NO2 gas sensor based on metal oxide was discussed in the aspects of
morphology, preparation method, modi cation, doping and composite.
The most concerning aspect of metal oxide-based NO2 gas sensors was
the
high operating temperature. Researchers were committed to reducing the
operating temperature of MO, improving gas response and accelerating the
response/recovery speed. Not enough research has been done to increase
insensitivity to humidity. However, humidity played a key role in the gas
sensor and the adsorption of H2O molecules will affect the adsorption and
reaction of the target gas on the sensor surface, thereby reducing the
sensitivity of the gas sensor. In the future, more attention could be paid to
fi
improving the insensitivity to humidity for preparing a NO2 gas sensor with
high performance and high humidity resistance suitable for practical
applications.

We would like to express our sincere gratitude to Dr.


Swetha ma'am, and her invaluable guidance and support
throughout the preparation of this report on “Metal Oxide
Gas sensors for detecting NO2 industrial exhaust gases,
recent developments. We would also like to acknowledge
the insightful article by Qingting Li , Wen Zeng Yanqiong Li
whose research provided signi cant input and inspiration
for the study. Their work has been crucial in shaping the
understanding of the advancements in metal oxide base
sensor for environmental monitoring.
fi

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy