You are on page 1of 27

Sensors & Actuators: B.

Chemical 364 (2022) 131876

Contents lists available at ScienceDirect

Sensors and Actuators: B. Chemical


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

Recent advances in SnO2 nanostructure based gas sensors


Yoshitake Masuda
National Institute of Advanced Industrial Science and Technology (AIST), 2266-98 Anagahora, Shimoshidami, Moriyama-ku, Nagoya 463-8560, Japan

A R T I C L E I N F O A B S T R A C T

Keywords: The detection of low-concentration gases and odors in fields such as healthcare, mobility, and indoor environ­
Gas sensor ment control is attracting tremendous attention. The development of high-sensitivity gas sensors capable of
Odor sensor detecting low concentrations of gases and molecules is strongly desired. This review focuses on tin oxide
SnO2
nanomaterials, which are substances that are being actively researched as semiconductor-type gas sensors. In
Nanostructure
Tin oxide
particular, the development of novel tin oxide nanomaterials over the last decade and their gas sensing appli­
cations are discussed. It is revealed that dimension and morphology affect the sensing performance. Tin oxide
nanomaterials having nano-meter size which is similar to size of a depletion layer, and controlled microstructure
such as a nanosheet structure shows high sensing performance. The dendritic structure in which 2D nanosheets
are connected by crystal growth points the direction for future sensor development. Furthermore, the high gas
adsorption performance and reactivity of the metastable crystal plane will be a guide for future sensor
development.

1. Introduction development of taste sensors and odor sensors must be focused on


because of the difficulties involved in their fabrication compared to that
The five basic human senses—eyesight, hearing, touch, taste, and of other sensors.
smell—are important for collecting information. Electronic devices that A taste sensor containing lipid/polymer receptors is commercially
imitate the functions of these five senses have been developed for de­ available, in which taste is quantified using the output potentials from
cades. Among these senses, eyesight is the most developed. Image sen­ the receptors (Taste sensor TS-5000Z, Irie Corporation.) [1,2]. However,
sors and video cameras have been widely used for capturing moving aspects such as high-speed sensing, continuous monitoring, miniaturi­
images. Telephoto, high-speed, and infrared cameras have more func­ zation, and cost reduction must be accomplished for taste sensors
tions than those of human eyesight. In addition, information processing compared to visual and auditory sensors.
technology is applied as a brain substitute to analyze the image data and The development of odor sensors is crucial because of their possible
to recognize objects and situations. Moreover, image recognition based applications for various purposes. Odor sensors have been extensively
on artificial intelligence (AI) technology has also made remarkable used to detect gas leakage and in fields such as healthcare, mobility, and
progress. Hearing is also being imitated by electronic devices, such as spatial environment control, in which the detection of low-
recording devices, which conveniently record audio data, and acoustic concentration biogas and odor is crucial [3]. However, sensors that
and ultrasonic sensors. Moreover, voice information processing and can detect high-concentration gas leaks cannot analyze
recognition technologies are being developed using AI. Visual and low-concentration odors, which requires the development of sensors
auditory imitation devices are at the technical level of sensing as well as with higher sensitivities. In addition, most odor components have a
information processing and recognition. Judgment technology based on larger molecular weight than that of gas molecules in gas leakage
information recognition has also been developed in combination with detection. Moreover, certain odors are mixtures of multiple molecules,
AI. Force sensors, slip sensors, and proximity sensors have been devel­ and the coexisting molecules and water vapor behave as contaminant or
oped as touch-imitation devices. Force sensors are the basis of touch noise gases during detection of the target molecules. Therefore, the
sensors and can detect pressure, torque, and direction. Slip sensors can identification of target molecules even in the presence of these
detect the speed and degree of slip between an object and a sensor. contaminant gases is critical. Sensing technology is advancing rapidly
Proximity sensors can detect distances to noncontact objects and are with the internet of things (IoT) technology. Constant monitoring using
more functional than the human touch sense. However, the numerous sensors at various locations is a possible reality in the future,

E-mail address: masuda-y@aist.go.jp.

https://doi.org/10.1016/j.snb.2022.131876
Received 9 February 2022; Received in revised form 5 April 2022; Accepted 9 April 2022
Available online 15 April 2022
0925-4005/© 2022 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

which necessitates the development of odor sensors that are compact, respectively. The percentage of searched papers for SnO2, ZnO, TiO2,
cost-effective, vibration-resistant, stable against temperature fluctua­ WO3, In2O3, Fe2O3, or CeO2 is, 28.82%, 26.65%, 11.42%, 9.21%, 6.13%,
tions, durable against ultraviolet rays and drying, and exhibiting 4.22%, or 1.52%, respectively. The percentage of searched papers for
long-term stability and reliability. CuO, NiO, Co3O4, Cr2O3, or Mn3O4 was 4.61%, 4.26%, 2.31%, 0.72%, or
Nikkei Business Publications, Inc has reported that biogas measure­ 0.14%, respectively. There are more papers on gas sensors using n-type
ment is attracting attention as a non-invasive method that does not oxides than those on p-type. In particular, SnO2 has received a great deal
burden the body [4]. Examples that have already been introduced of attention. Gas sensors with SnO2 is focused on in this review. Espe­
include the breath NO (nitric oxide) test for asthma and the urea breath cially, recent advances in SnO2 nanostructure based gas sensors are
test for the helicobacter pylori test that is bad in the stomach and duo­ reviewed.
denum. Their usefulness is medically recognized. They introduced a
bio-sniffer (Tokyo Medical and Dental University) [5], a membrane-type 2. Development of SnO2 nanostructures for gas sensing
surface stress sensor (National Institute for Materials Science, Swiss
Federal Institute of Technology in Lausanne) [6], a breath sensor (Na­ Metal oxides, such as SnO2, ZnO, and TiO2, are used in various ap­
tional Institute of Advanced Industrial Science and Technology, Figaro plications such as gas sensing, molecular sensing, and for fabricating
Engineering Inc), a breath ammonia sensor (FUJITSU LABORATORIES secondary batteries and catalysts. In particular, metal oxide nano­
LTD.), A breath sensor (NTT DOCOMO, INC.) And a breath ammonia materials are expected to exhibit outstanding device characteristics
sensor (NIHON DEMPA KOGYO CO., LTD., The University of Kita­ owing to the effects of their high specific surface area and specific crystal
kyushu) as typical biogas sensors in Japan. planes that exhibit unique physical properties, chemical characteristics,
Various types of gas sensors and odor sensors have been developed. crystal defects, surface defects, and quantum effects. The synthesis of
Semiconductor-type gas sensors are highly reliable sensors that have various SnO2 nanostructures has been attempted because SnO2 exhibits
been used in gas alarms for many years. Various metal oxides have been excellent characteristics as a semiconductor-type gas sensor.
studied for gas sensors. SnO2, In2O3 and WO3 have advantages in sta­ There are active reports on the sensor performance, sensing mech­
bility among them. These materials are therefore used as commercial anism, guidelines for development, etc. for semiconductor gas sensors
sensors. SnO2 was the first material developed as a semiconductor using SnO2 [7–10]. N. Yamazoe and K. Shimanoe have reported valuable
sensor. Research on various SnO2 structures, basic properties and ad­ guidelines for development of metal oxide nanostrures as follows [7].
ditives has been carried out for many years. Research has also been Crystallite sizes of oxide semiconductors should be as small as possible.
conducted on the mechanism of SnO2 type sensors. SnO2 is one of the Sensitizers should be dispersed as finely as possible. Sensing layer
most important materials for semiconductor sensors in terms of both thickness and porosity should also be optimized to improve selectivity
application and basic science. However, SnO2 thick film for gas alarms is and durability.
not sufficient to detect low-concentration biogas and odors. Therefore, In this review, SnO2 nanomaterials and detection limit of low-
the development of SnO2 nanostructures has been attracting attention in concentration gas have been focused. Gas sensing mechanism of
recent years in order to further improve the characteristics of SnO2 type semiconductor-type gas sensor is discussed in Section 2.1. Nano­
gas sensors. materials are categorized to 0D (nanoparticles), 1D (wires, needles,
In this review, semiconductor-type gas sensors are focused on whiskers, belts, etc.) or 2D (nanosheets). The 0D nanomaterials are
because they are highly stable and reliable and have many years of reviewed in Sections 2–2. The 1D nanomaterials are reviewed in Sec­
commercial experience in areas such as gas leak detection [7–10]. In tions 2–3. 2D heterostructure is reviewed in Sections 2–4. 2D nanosheet
particular, the use of SnO2 nanomaterials is highlighted because they are is discussed in detail in Section 3 because the 2D nanosheet is expected
being actively researched and developed as sensing elements for to be developed further for low-concentration gas sensing.
semiconductor-type gas sensors. Moreover, examples of research and
development in this area over the past decade are presented. 2.1. Gas sensing mechanism of semiconductor-type gas sensor
The number of research papers has been investigated on the Web of
Science on 22 March, 2022 (Fig. 1). Key words of the chemical formula Gas sensing mechanism of semiconductor-type gas sensor has been
of the sensor material and “gas sensor” are used for search. For instance, discussed for both of n-type metal oxides [7] or p-type metal oxides [11].
the keywords of “SnO2” and “gas sensor” are used to search for SnO2 gas Metal oxide particles in n-type semiconductor gas sensor act as receptors
sensors. The number of searched articles for n-type SnO2, ZnO, TiO2, and transducers. The role of a transducer is played by each contact be­
WO3, In2O3, Fe2O3, or CeO2 is 7516, 6952, 2978, 2402, 1600, 1100, or tween the particles. Interface between particles is the most resistant part
396, respectively. The searched articles for p-type CuO, NiO, Co3O4, of the gas sensor. It determines the resistance of the gas sensor. The
Cr2O3, or Mn3O4 is 1202, 1110, 603, 187, or 37, respectively. The per­ surface space charge layer model or the double Schottky barrier model
centage of searched papers for n-type or p-type is 88.0% or 12.0%, are shown in Fig. 2a or Fig. 2b [7], respectively. The thickness of the
depletion layer (w) increases as oxygen adsorption as anionic species
(typically O− ) increases. It decreases as the adsorbed oxygen is
consumed with an inflammable gas such as H2. The double Schottky
barrier is formed across the contact between particles. Height of the
barrier changes to induce the contact resistance change and the resis­
tance change in the gas sensor (Fig. 2b) [7]. The receptor model in
Fig. 2a [7] assumes that the semiconductor particles are sufficiently
large. The mechanism has been further discussed in details to give
quantitative information regarding gas response. Size of the metal oxide
nanoparticles is very tiny (typically about 10 nm in diameter) in recent
advanced gas sensors. The space charge layer can easily extend over the
entire area of grains; that is, w grows to grain radius (r), at PO2 signif­
icantly below that in air, PO2 (Fig. 2c) [7]. A new process of electron
depletion has to take place afterward until the grains reach electrostatic
equilibrium with oxygen adsorption at PO2 (a). A mechanism is pro­
Fig. 1. Studies on n- and p-type oxide semiconductor gas sensors (internet posed in which electron depletion is achieved by shifting the Fermi level
search of Web of Science on 22 March, 2022). downward by pkT. Electrons move near surface of the particles as shown

2
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 2. Diagrams of electron depletion for oxide grains and the resistance of contact between grains. (a) Space charge layer model, (b) double Schottky barrier model,
(c) regional and volume depletion model, (d) surface conductive grains contact model, (e) volume depletion model for the dendric growth of thin nanosheet.

in Fig. 2d [7]. 2.1.1. Hierarchically structured SnO2 microspheres


Regional and volume depletion model [7] can be applied to the [1-ppm-formaldehyde (HCHO)].
nanosheets. Separated nanosheets have been synthesized with exfolia­ Li et al. developed SnO2 microspheres with a hierarchical structure
tion method [12–16]. Grain boundary is formed between each separated for formaldehyde (HCHO) gas sensing [18], which were synthesized
nanosheets. On the other hand, a dendritic structured film of SnO2 using a hydrothermal method without the use of polymer templates or
nanosheet has been developed with facet controlled growth method surfactants (Fig. 3a). The as-synthesized SnO2 microspheres were
[17]. Dendritic growth governed the crystal growth of the SnO2 (101) composed of numerous small spheres with an average diameter of
nanosheet assembled film. The branch angle of the connected nano­ approximately 250 nm. Each small sphere consisted of numerous pri­
sheets is in two types of connections, which are defined by the branch mary nanocrystallites with an average size of approximately 8 nm
angle of either 90◦ (type I) or 46.48◦ (type II). (Fig. 3b− c). The SnO2 microspheres exhibited gas sensing responses
Volume depletion model can be applied to the dendric growth of thin (Ra/Rg) of 5.72, 7.70, 10.85, 21.12, 38.26, 66.25, and 144.90 for 1, 3, 5,
nanosheet (Fig. 2e). The SnO2 nanosheets in the dendritic structure has a 50, 100, 200, and 500 ppm formaldehyde, respectively, at an operating
thickness of 5–10 nm and an in-plane size of 100–1600 nm [17]. Half of temperature of 200 ◦ C. Additionally, the response to 100 ppm formal­
nanosheet thickness (t) is smaller than depletion layer thickness (w). dehyde was higher than that of acetone (6.42), ammonia (1.74), xylene
Electrons move without influence of grain boundary and Schottky bar­ (1.99), methanol (19.38), and toluene (1.55) (Fig. 3d). The ~8-nm-sized
rier. Conductance is influenced by electron concentration. nanocrystallites exhibited superior selectivity for formaldehyde gas
Half of the sheet thickness (t) in Fig. 2e is illustrated with the same sensing. Moreover, the reaction mechanism of the SnO2-microspher­
size as the grain radius (r) in Fig. 2c volume depletion and Fig. 2d. Both e-based gas sensor for formaldehyde sensing was elucidated (Fig. 3e− f).
of the nanoparticles with w=r (Fig. 2c, volume depletion) and the
dendric growth of thin nanosheet with w=t/2 (Fig. 2e) are the maximum 2.1.2. Single-atom Pt-functionalized 9-nm-thick SnO2 thin film
size to realize volume depletion model. The dendric-structured nano­ [10-ppb-triethylamine, LOD: 7 ppb].
sheet film has effective conductive path compared with particulate film Xu et al. fabricated single-atom Pt-functionalized SnO2 (Pt/SnO2)
having many grain boundaries. Therefore, the dendric-structured thin films with thicknesses of 4–18 nm [19]. The 9-nm-thick Pt/SnO2
nanosheet film has the advantage to form the ultra thin high response films showed an optimal response to 10 ppb triethylamine (TEA) owing
membrane for a microelectromechanical system (MEMS)-type to their thickness being comparable to the Debye length. A decrease in
microsensor. the film thickness to 4 nm led to a carrier concentration that was two
Volume depletion model can be also applied to the grain boundary of orders of magnitude lower. The LOD of the Pt/SnO2 sensor was esti­
thin nanosheets. Separated thin nanosheets can be synthesized with the mated to be 7 ppb. Additionally, the effects of the Pt catalyst and the
solution process [17]. Electrons move without influence of grain SnO2 thickness were discussed.
boundary and Schottky barrier as shown in volume depletion model for SnO2 thin films were deposited on SnO2/Si wafers by thermal atomic
nanoparticle in Fig. 2c-d. Conductance is influenced by electron layer deposition (ALD) for application as sensing layers in resistive gas
concentration. sensors. Tetrakis(dimethylamino)tin(IV) (TDMASn) and H2O were used
On the other hand, an oxidizing gas such as NO2 increases the as the precursors. Single-atom Pt catalysts deposited on the 9-nm-thick
resistance of the semiconductor gas sensor for the following reasons. SnO2 film led to a considerable enhancement in the sensor response
Similar to the case where oxygen is negatively charged on the surface of and a decrease in its operating temperature. The response of Pt/SnO2 to
an n-type oxide semiconductor such as SnO2 particles, when an 10 ppm TEA was as high as 136.2 at 200 ◦ C, which was almost seven
oxidizing gas such as NO2 is adsorbed on the surface of SnO2 particles, times higher than that of SnO2 at 260 ◦ C (20.3). The response time of the
electrons are trapped and NO2 becomes NO-2. Therefore, the gas sensor Pt/SnO2 sensor was 3 s, which was only half that of SnO2. The Pt/SnO2
shows a higher resistance value in the oxidizing gas than in the air. sensor recovered within 6 s, which was two orders of magnitude faster
than that of pure SnO2 (300 s). The extremely rapid response–recovery

3
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 3. (a) A schematic illustration of the for­


mation process of SnO2 microsphere. (b)
HRTEM image of as-synthesized SnO2 micro­
spheres and (c) the corresponding HRTEM
image with obvious SnO2 lattice fringe. (d)
Response of the SnO2 microspheres gas sensor
to 100 ppm of different gases at 200 ◦ C. A
schematic diagram of the proposed reaction
mechanism of the SnO2 microspheres gas sensor
to formaldehyde: (e) in air, (f) in formaldehyde
(HCHO) [18].
Reprinted with permission from ref. [18]
Copyright 2021 Elsevier. .

dynamics of the Pt/SnO2 sensor were attributed to the Pt single atoms, WO3/SnO2 nanoparticles (LOD: 1 ppm) [21], NiO/SnO2 hollow spheres
which reduced the activation energy of surface reactions and enhanced (2 ppm) [22], SnO2 hollow microfibers (2 ppm) [23], TiO2/SnO2
the adsorption of TEA in the response stage and the desorption of the nanosheets (2 ppm) [24], Zn2SnO4/SnO2 microspheres (500 ppb) [25],
resultant molecules during recovery [20]. The LODs of the Pt/SnO2 and MoS2/SnO2 nanofibers (5 ppm) [26], Au/Mg-TiO2/SnO2 nanosheets
SnO2 sensors for TEA were calculated to be 7 ppb and 98 ppb, respec­ (2 ppm) [27], Pd/In2O3 microstructures (1 ppm) [28], Al2O3/α-Fe2O3
tively, which were lower than those of other sensing materials such as nanofibers (500 ppb) [29], Au/ZnO nanorods (1 ppm) [30],

4
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Co3O4/In2O3 microtubes (2 ppm) [31], α-MoO3 nanoflowers (500 ppb) whereas those of the Pt/SnO2 and SnO2 sensors were 136.2 and 20.3,
[32], Au/SnO2/α-Fe2O3 nanoneedles (2 ppm) [33], ZnCo2O4 respectively.
single-layer nanochain (5 ppm) [34], Au-TiO2/m-CN nanocomposite
(1 ppm) [35], Au/α-Fe2O3 nanorods (1 ppm) [36], In2O3 microtubes 2.1.3. Pulse-driven gas sensor containing Pd-loaded SnO2 nanoparticles
(100 ppb) [37] and Au/Co3O4/W18O49 hollow spheres (81 ppb) [38] [1-ppb-toluene, LOD: 200 ppt or 7 ppt].
(Table 1). To characterize the sensor selectivity, the thin-film sensors Suematsu et al. developed a heater-switching pulse-driven micro gas
were further exposed to a group of molecules including C6H15N, C2H6O, sensor[40] (Fig. 4a–b) composed of a microheater and a sensing elec­
CH3COCH3, CH3(CH2)3OH, CH3OH, HCHO, and C7H8. The Pt/SnO2 trode fabricated with Pd-loaded SnO2 clustered nanoparticles (CNPs).
sensor showed improved responses to all the gases. The selectivity co­ The Pd-SnO2 nanoparticles were used in the sensor because of their high
efficients calculated using the responses to the various analytes indi­ sensitivity to toluene gas[41]. A SnCl4.5 H2O solution (1 M) was added
cated that the single-atom Pt catalysts dramatically improved the to a NH4HCO3 solution (1 M) to form stannic acid gel (SnO2⋅nH2O), to
selectivity for TEA. These investigations revealed that the outstanding which Pd[(NO2)2(NH3)2] was subsequently added. NH3 solution was
sensing performance originated from the synergistic combination of the used to control the pH of the solution in the 9.3–10.6 range. The solution
optimized film thickness, which was comparable to the Debye length of was hydrothermally treated at 200 ◦ C for 3 h at 10 MPa with stirring at
SnO2, the spillover activation of oxygen by the single-atom Pt catalysts, 600 rpm. This reaction was controlled by introducing nitrogen into the
and the OVs in the SnO2 films. The Debye length, LD, is a characteristic of reactor. Micro droplets of a paste containing the Pd-SnO2 nanoparticles
a semiconductor material for a particular donor concentration and can and glycerin at a 1:1 ratio (w/w) were placed onto a MEMS-type
be calculated as follows: [39]. microsensor using a capillary tube and a micromanipulation system
( / )1/2 (PatchMan NP2, Eppendorf) (Fig. 4c–e). The sensor was sintered at
LD = ε0 εkB T q2 nc , 450 ◦ C for 12 h using a microheater under a flow of synthetic air. The
sensor was heated in pulse-heating mode, and the heater was repeatedly
where kB is the Boltzmann’s constant, ε is the dielectric constant, ε0 is switched on and off. The volatile organic compound (VOC) gas perme­
the permittivity of free space, T is the operating temperature, q is the ated through the particulate film when the sensor was not heated
electron charge, and nc is the carrier concentration, which corresponds (Fig. 4a–b). In contrast, the gas expanded when the heater was switched
to the donor concentration assuming full ionization. The carrier con­ on, and a portion of it was exhausted. The sensor response (Ra/Rg)
centration was obtained from Hall measurements. The temperature, calculated using the initial electrical resistance was 3 and 4.5 for 1 ppm
carrier concentration, mobility, and Debye length of the 4-nm-thick and 8 ppb toluene, respectively (Fig. 4f− g). Moreover, the
SnO2 film were 300 K, 2.17 × 1020 m− 3, 121.30 cm2/V⋅s, and sensor-response-related detection limit of the pulse-driven sensor con­
375.1 nm, respectively. whereas those of the 9-nm-thick SnO2 film were taining the Pd-SnO2 nanoparticles was calculated to be approximately
300 K, 9.22 × 1022 m− 3, 34.3 cm2/V⋅s, and 14.5 nm, respectively. The 200 ppt toluene.
carrier concentration of the 9-nm-thick SnO2 film was higher than that Suematsu et al. also fabricated Pd-SnO2 CNPs on a microsensor de­
of the 4-nm-thick counterpart, whereas the mobility and Debye length of vice with a double-pulse-driven mode [42]. This mode assisted in
the 9-nm-thick SnO2 film were lower than those of the 4-nm-thick switching the heater-on periods during the high-temperature preheating
equivalent. and measurement phases and the rest phase during the heater-off period
Additionally, the sensing properties of the Pt/SnO2 devices were between the preheating and measurement phases. The electrical resis­
compared with those of commercial TGS 2602 gas sensors (Figaro En­ tance in synthetic air and the sensor response to toluene increased with
gineering Inc.). The TGS 2602 gas sensor had a thick-film sensing ma­ increasing preheating temperature because of an increase in the amount
terial for VOC detection. The sensing layers of the three devices were of O2− adsorbed on the particle surfaces. The sensor response was
found to exhibit inhomogeneous surfaces, along with certain cracks in improved by extending the interval between the preheating and mea­
the sensing layers, which possibly led to unsatisfactory reproducibility surement phases. Moreover, the lower detection limit for toluene gas
of the commercial sensors. Moreover, gas sensing tests confirmed that was improved to 7 ppt using preheating and measurement temperatures
the three commercial devices delivered remarkably different sensing of 400 ◦ C and 250 ◦ C, respectively, and a rest time of 100 s
responses to TEA at an optimal operating temperature of 260 ◦ C. The
TGS 2602 gas sensors had a response (Ra/Rg) of 15–16 for 10 ppm TEA,

Table 1
Comparison of TEA detection performances of various materials [19], [21–38].
Sensing materials Operating temperature (℃) Response/ppm Response/recovery time (s) LOD (ppm) Reference

WO3/SnO2 nanoparticles 220 87/50 6/7 1 [21]


NiO/SnO2 hollow spheres 220 46.5/10 11/34 2 [22]
SnO2 hollow microfibers 270 49.5/100 14/12 2 [23]
TiO2/SnO2 nanosheets 260 52.3/100 12/22 2 [24]
Zn2SnO4/SnO2 microspheres 250 19.6/20 2/184 0.5 [25]
MoS2/SnO2 nanofibers 230 106.3/200 – 5 [26]
Au/Mg-TiO2/SnO2 nanosheets 260 30.43/50 9/95 2 [27]
Pd/In2O3 microstructures 220 47.56/50 4/17 1 [28]
Al2O3/α-Fe2O3 nanofibers 250 15.19/100 1/17 0.5 [29]
Au/ZnO nanorods 40 22/50 11/15 1 [30]
Co3O4/In2O3 microtubes 250 786.8/50 47/20 2 [31]
α-MoO3 nanoflowers 250 416/100 3/1283 0.5 [32]
Au/SnO2/α-Fe2O3 nanoneedles 300 39/100 4/203 2 [33]
ZnCo2O4 singlelayer nanochain 200 13/100 7/57 5 [34]
Au-TiO2/m-CN nanocomposite 175 78.9/50 – 1 [35]
Au/α-Fe2O3 nanorods 40 17.5/50 12/8 1 [36]
In2O3 microtubes 300 72/100 12/650 0.1 [37]
Au/Co3O4/W18O49 hollow spheres 270 283.1/50 9/14 0.08 [38]
Pt/SnO2 film (9 nm, ALD) 200 136.2/10 3/6 0.01 [19]
SnO2 film (9 nm, ALD) 260 20.3/10 6/300 0.1 [19]

5
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 4. Schematic illustrations of gas-diffusion behavior in


the (a) heating-off and (b) heating-on phases. (c) Photo­
graphic and (d) SEM images of a blank MEMS sensor. (e)
Side view of the SnO2 microsensor. (Third step) (f) Tran­
sient response curve of the Pd-SnO2 clustered nanoparticle
(CNP) microsensor to various toluene concentrations at
1.04 V (~250 ◦ C). (g) Relationship between the response
of the Pd-SnO2 CNP sensor and toluene concentration [40].
Reprinted with permission from ref. [40] Copyright 2021
American Chemical Society.

2.1.4. SnO2 1D nanostructures Pan et al. prepared SnO2 nanowires by chemical vapor deposition
Prof. Sheikh Akbar developed SnO2 nanowires [43,44], hierarchical (CVD) [65] (Fig. 5a–b). The SnO2 nanowires were modified by Ar/O2
SnO2 nanowire–TiO2 nanorod brushes on fluorine doped tin oxide (FTO) plasma treatment through preferential etching of lattice oxygen atoms
substrates [45], core–shell TiO2–Al2O3 nanowires [46,47]. The n-type (Fig. 5c). The plasma treatment gradually modified the chemical
SnO2 nanowires were coated by magnetron sputtering and chemical composition and morphology of the SnO2 nanowires. The
vapor deposition with n-type TiO2, WO3, ZnO, Nb2O5, and MoO3 as well plasma-treated SnO2 nanowires showed a higher conductance than that
as p-type Cr2O3, Co3O4 and NiO [48]. He reviewed 1D nanostructured of the as-deposited SnO2 nanowires. This was due to the effect of band
metal oxide gas sensors based on SnO2, ZnO, TiO2, In2O3, WOx, AgVO3, bending induced by surface charges associated with the additional states
CdO, MoO3, CuO, TeO2 and Fe2O3 [49]. He also reported on role of resulting from the shallow donor states of the abundant oxygen va­
oxygen vacancies [50], metal oxide heterojunctions [51–53] and gas cancies (OVs). X-ray photoelectron spectroscopy (XPS) profiles of the
selectivity mechanisms [54] using examples of SnO2 nanofibers [55], as-deposited and plasma-treated SnO2 nanowires (synthesized at 40 W)
Co3O4 nanofibers [56], Co3O4 modified SnO2 nanofibers [57], NiO/ZnO showed a distinct shift in the Sn 3d peaks toward a lower binding energy,
sheets/flowers [58], NiO/ZnO hollow microspheres [59], Co3O4 nano indicating chemical reduction of the Sn species upon exposure to radi­
rhombus [60], Co3O4 sea urchins [61], Co3O4 nanocubes [62], Co3O4 ofrequency plasma. The TEM micrograph of the nanowires showed 1D
nanosheets [63], Co3O4 nanorods [63], Co3O4 nanocubes [63], and heterostructures with a crystalline SnO2 core and an amorphous over­
Co3O4 powders [63]. He discussed gas selectivity of various p-n heter­ layer, and corroborated the anticipated structural changes by ions with a
ojunctions, n-n heterojunctions and p-p heterojunctions. It was higher kinetic energy that impinged on the nanowire surface. The rough
concluded that the process to design selective sensors involved the type amorphous overlayer in the plasma-treated samples increased the den­
of oxides, the ratio of oxides, and the morphology. Various oxide com­ sity of dangling bonds and free lattice sites, thereby favoring surface
binations may then be incorporated into a sensor array to identify gas adsorption phenomena. The plasma-treated SnO2 nanowires (40 W) had
using machine learning [64]. a gas sensing response (Ra/Rg, where Ra: resistance of the sensor in air,
and Rg: resistance of the sensor in the target gas) of 4.98 for 100 ppm
2.1.5. SnO2 nanowires with a surface amorphous layer ethanol at 250 ◦ C, which was higher than that of the as-deposited SnO2
[100-ppm-ethanol]. nanowires (1.69). The plasma treatment was effective in increasing the

6
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

initiated by the decomposition of the preformed ZnSn(OH)6 induced by


alkali etching. The diameter and density of the nanorods were controlled
by the addition of NaOH, which yielded smaller and denser nanorods.
Moreover, nucleation was facilitated by NaOH, leading to rapid growth
in all directions.

2.1.7. Oriented SnO2 nanorod array


[200-ppb-NO2, LOD: 132 ppt].
Song et al. fabricated SnO2 nanorods for the detection of 200 ppb
NO2 at a relative humidity (RH) of 20% and estimated a detection limit
of 132.3 ppt [67]. Ion-activated sensing, in which ionic conduction was
realized by the adsorbed hydroxide layer in the presence of humidity,
was established, which led to proton transport [68,69].
A 1-µm/550-µm-thick SiO2/Si substrate with 150-nm/30-nm-thick
Pt/Ti interdigitated electrodes was used as the sensor. The substrate
had 20 electrodes in a 1 mm × 1 mm area, which were spaced apart by
5 µm. The SnO2 nanorods were deposited on the substrate by glancing
angle deposition (GAD) using an electron beam evaporator (Fig. 7a− b).
The samples were annealed at 550 ◦ C for 2 h under a gas flow rate of
1000 sccm. The sensor resistance was measured using a DC bias voltage
of 0.5 V. The sensor showed a linear response (Rg/Ra ~ 1) in the range of
200 ppb to 10 ppm at 20% RH (Fig. 7c− d). However, the response
saturated for NO2 concentrations above 10 ppm because the active sites
were almost covered by the NO2 species. The theoretical detection limit
(signal-to-noise ratio > 3) was calculated to be 132.3 ppt NO2 under
20% RH. The gas selectivity of the SnO2 nanorods was measured under
dry and 20% RH conditions using 5 ppm NO2; 10 ppm CO, C2H5OH, and
C7H8; and 1000 ppm H2 at an applied voltage of 0.5 V. The highest
response and selectivity to NO2 were observed at 20% RH (Fig. 7e)
compared to the other tested gases because the reducing gases could not
react with the hydroxide layer. Moreover, the sensing mechanism under
high-humidity conditions was explored (Fig. 7g–j). Humidity is known
to degrade gas sensing properties, such as response, response time, and
recovery time, which is referred to as water poisoning [70]. However,
the response of the SnO2 nanorods improved with increasing RH from
0% to 20% and decreased with a further increase in RH to 80%.
Essentially, the SnO2 nanorods exhibited improved sensing properties
under humid conditions and exhibited an excellent response at 20% RH
compared to that under dry conditions only at room temperature.
Fig. 5. (a) SEM micrograph of as-deposited SnO2 nanowires. HR-TEM images of Therefore, the conventional electron exchange reaction was presumed
(b) as-deposited and (c) plasma-treated (40 W) SnO2 nanowires [65]. to occur above 100 ◦ C, and the ion-activated mechanism was operative
Reprinted with permission from ref. [65] Copyright 2021 American Chemi­ at room temperature. In addition, sensors were fabricated using NiO,
cal Society.
Cr2O3, Co3O4, and WO3 nanorods for the detection of 5 ppm NO2 at
0.5 V under 20% RH (Fig. 7f). The sensing resistances of the Cr2O3,
sensor response, and was primarily caused by the amorphous surface Co3O4, and WO3 nanorods were not significantly altered upon exposure
layers with nonstoichiometric surface compositions. Moreover, the high to NO2; however, only the SnO2 nanorods exhibited enhanced sensing
volume of dangling bonds and free lattice sites in the amorphous surface performance.
layers improved the surface adsorption phenomenon.
2.1.8. SnOx nanopillars
2.1.6. SnO2 nanorod particles with {110} facet [10-ppt-H2 in vacuum].
[44-ppm-ethanol]. D’Arsié et al. fabricated SnOx nanopillars and evaluated their sensing
Kuang et al. synthesized hierarchical SnO2 nanostructures made of performance in an ultra-high-vacuum (UHV, 10− 10 mbar) chamber for
superfine nanorods using a hydrothermal method for ethanol sensing 10 ppt H2 at 230 ◦ C [71]. This contributed to the detection of low
[66] (Fig. 6). The hierarchical SnO2 sensor (S3) was prepared by first concentrations of H2 using SnOx nanopillars, although the UHV exper­
dissolving zinc acetate (Zn(CH3COO)2; 0.05 g) in a mixture of deionized iments did not replicate the typical operational conditions of the sensor.
(DI) water (15 mL) and ethanol (15 mL) with stirring. NaOH (0.9 g) and To prepare the SnOx nanopillars, ITO/glass substrates were first
stannic chloride pentahydrate (SnCl4.5 H2O) (1 mmol) were succes­ sequentially exposed to H2 and C2H2 at ~650 ◦ C to form vertical
sively dissolved in the preceding mixture. The resulting solution was nanostructures of carbon filled with tin [72]. These specimens were
maintained at 180 ◦ C for 24 h, and the subsequently obtained powder heated at ~650 ◦ C in oxygen to incinerate the carbon shells. No further
was dissolved in ethanol to form a paste, which was coated onto an changes in the structures were observed at this temperature upon
Al2O3 tube to form an ~100–200-μm-thick film between two parallel Au incineration of the carbon. The final sensor was composed of a glass
electrodes. The SnO2 film was sintered at 400 ◦ C for 2 h in air, and the support coated with an ~120-nm-thick Sn-depleted ITO film covered
fabricated gas sensor was aged at 240 ◦ C for 72 h to improve the stability with vertical nanopillars of SnOx (Fig. 8a− b).
and repeatability. The S3 sensor showed a response (Ra/Rg) of 44 for The sensor showed variation in resistance (R − R0)/R0 [%]) for H2,
50 ppm ethanol at a heater current of 150 mA. In addition, the synthesis CH4, and O2 detection at room temperature and at 90 ◦ C, where R is the
protocol was comprehensively explained. Nucleation was found to be resistance after 100 s of gas exposure and R0 is the initial resistance prior

7
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 6. (a) Schematic for the possible growth of the hierarchical SnO2 nanostructures made of superfine nanorods. FESEM (b), TEM (c) and HRTEM (d) images of the
hierarchical SnO2 nanostructure made of superfine nanorods (S3) [66].
Reprinted with permission from ref [66]. Copyright 2021 Elsevier.

to the introduction of the gas in the experimental chamber (Fig. 8c). The (SnCl2.2 H2O) and 1.0 g of polyvinylpyrrolidone (PVP, molecular
experiment was performed under low-vacuum conditions (10− 5 mbar weight = 1,300,000) in N,N-dimethylformamide (DMF; 0.425 mL) and
base pressure) at 90 ◦ C and 10 ppm. The sensor exhibited variation in ethanol (0.425 mL), and stirring the mixture for 12 h at room temper­
resistivity (15–20%) for H2 and O2, and showed a one-order-of- ature (Fig. 9a). A solution of the Y-doped SnO2 nanofibers was synthe­
magnitude lower sensitivity for CH4. The sensor was insensitive to sized by replacing tin(II) chloride dihydrate with an equal amount of
CO2 and N2 concentrations of up to 1000 ppm. Moreover, the SnOx yttrium(III) nitrate hexahydrate (Y(NO3)3.6 H2O) in the aforementioned
nanopillar sensor responded to 10, 50, 100, and 500 ppt H2 at 230 ◦ C in protocol. The nanofibers were obtained by electrospinning from a
UHV (Fig. 8d). However, the sensor did not respond to ppt-level con­ mixture of the two aforementioned solutions at 10.0 kV between the
centrations of CH4, CO2, CO, NO2, and N2 at 230 ◦ C. This behavior can sample collector and the syringe pump. The nanofibers were collected
be explained by the effects of the gas molecules interacting with n-type on SiO2/Si substrates and subsequently calcined in air at 550 ◦ C for 1 h
SnOx nanopillars [73,74]. Negative charges trapped by oxidizing species to obtain the Y-doped SnO2 nanofibers (Fig. 9b–d).
generally cause upward band bending, whereas the reaction with The sensor with 1 at% Y-doped SnO2 nanofibers exhibited high sta­
reducing gases, such as H2, causes downward band bending. However, it bility for NO2 sensing in the RH range of 0–87% (Fig. 9e− k). The re­
is worth noting that competitive adsorption and replacement of the sponses (Ra/Rg or Rg/Ra) of the 1 at% and 5 at% Y-doped SnO2
adsorbed molecules could decrease or even reverse the band bending, nanofibers to 10 ppm NO2 at 200 ◦ C under dry conditions were 70.38
resulting in an opposite variation in the conductivity [73,74]. In addi­ and 11.06, respectively, and those under 87% RH were 97.94 and 12.17,
tion, the sensing properties of the SnOx nanopillars were compared with respectively; the changes in response (ΔS) were calculated to be 0.39
those of a commercial sensor for VOCs, NH3, and H2S (Figaro Engi­ {(97.94 − 70.38)/70.38} and 0.1 {(12.17 − 11.06)/11.06}, respec­
neering Inc., TGS 2602 [75]). Both sensors exhibited similar responses tively. The Y-doped SnO2 nanofibers exhibited superior responses and
for NH3; however, the SnOx nanopillars showed high response and re­ smaller changes in the response compared to those of other materials
covery rates compared with those of the commercial sensor for 2.2, 3.6, such as ZnO thin films [78],WS2/graphene aerogel [79], Al-doped ZnO
and 30 ppm NH3 (Fig. 8e− h). Furthermore, the authors comprehen­ [80], Cr2WO6 [81], Rb2CO3-decorated In2O3 [82], and Rb-doped In2O3
sively elucidated the energy band alignment at the ITO/SnO2 isotype [83] (Table 2). The response curves of pristine and 1 at% Y-doped SnO2
heterojunction [76] and the electron injection from the SnO2 nanopillars nanofibers as a function of NO2 concentration were used to estimate the
to the substrate upon exposure to a reducing gas (Fig. 8i − j). sensor sensitivity and LOD, which were calculated for signal-to-noise
ratios of over three. The LOD of the sensor with the 1 at% Y-doped
2.1.9. Y-doped SnO2 nanofibers SnO2 nanofibers was estimated to be 103.71 ppt NO2, which was nearly
[20-ppb-NO2, LOD: 103.71 ppt]. six times higher than that of the pristine SnO2 sensor.
Kim et al. developed Y-doped SnO2 nanofibers [77] containing OVs Furthermore, the mechanism behind the highly stable sensor oper­
and hydrophobic Y2O3 nanoparticles on their surfaces for sensing 20 ppb ation under humidity was elucidated based on experimental results and
NO2, and estimated an LOD of 103.71 ppt under humid conditions. van-der-Waals (vdW)-corrected DFT calculations (Fig. 9l− q). Surface
To synthesize the nanofibers, a viscous solution of pure SnO2 nano­ OVs were created by aliovalent doping of Y atoms into the SnO2 nano­
fibers was first prepared by dissolving 0.5 g of tin(II) chloride dihydrate fibers. These vacancies made the sensor surface energetically more

8
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 7. (a) Schematic illustration of sensor fabrication by glancing angle deposition via electron beam evaporation. (b) Planar and cross-sectional SEM images of the
SnO2 nanorods. (c) Response curves of the nanorods to 5 ppm NO2 as a function of RH at room temperature. (d) Response of the SnO2 thin film and nanorods upon
exposure to 5 ppm NO2. (e) Response of the SnO2 nanorods to different target gases at 20% RH. (f) Response to various metal-oxide-based nanorods. Schematic
illustrations of the sensor response to NO2 on (g) oxygen-adsorbed and (h) hydroxide-functionalized SnO2 surfaces. (i) Energy band structure at the surface of SnO2
featuring surface adsorbate levels of O2, OH, and NO2. The energy level ordering of the adsorbate levels follows the free-space electron affinity trends of the
respective molecules. (j) Schematics of highly ordered SnO2 nanorods and the double Schottky barrier model.
Reprinted with permission from ref. [67] Copyright 2021 John Wiley and Sons.

9
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 8. SEM images of vertical nanostructured


SnOx pillars on an ~150-nm-thick ITO/Sn-reduced
substrate acquired at tilt angles of (a) 45◦ and (b)
0 ◦ . The two images share the same scale bar;
moreover, they were obtained prior to the sensing
experiments and are representative of the post-
sensing morphology, as no changes could be
determined via SEM. (c) Response of the vertical
nanostructured SnOx pillars after exposure to
several gases at low temperatures (30 ◦ C or 90 ◦ C)
in low vacuum (10− 5 mbar) for 100 s (d) Response
of the SnOx nanopillar sensor to low-pressure H2
(corresponding to ppt amounts) at 230 ◦ C in UHV.
(Second step) (e) Variation in resistivities of the
vertical SnOx nanopillar sensor (blue profile) and a
commercial FIGARO sensor (Figaro Engineering
Inc., TGS 2602; dashed red line) after exposure to
30 ppm NH3 in air at room temperature. The NH3
exposure period is marked with a green back­
ground. The recovery times were fitted with double
exponential functions (dashed black lines). (f)
Observed fluctuations compared with (g) the signal
noise. (h) Variation in the resistivities of the verti­
cal SnOx nanopillar sensor (blue profile) and the
FIGARO sensor (red profile) after exposure to
3.6 ppm and 2.2 ppm NH3 (pulse in the millisecond
range). (i) Energy band alignment at the ITO/SnO2
isotype heterojunction. (j) Schematic of electron
injection from the SnO2 nanopillars to the substrate
upon exposure to a reducing gas. The junctions are
present only at the nanopillar/ITO interface.
Therefore, the current from source to drain does
not have to pass through any junction [71]. (For
interpretation of the references to color in this
figure legend, the reader is referred to the web
version of this article.)
(i) (adapted from a previous report [76]). (j)
Reprinted with permission from ref [71]. Copyright
2021 Springer Nature.

10
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

(caption on next page)

11
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 9. (a) Schematic illustration of the fabricated gas sensor and the concept of this work demonstrating the synergistic effect of oxygen vacancy and hydrophobic
Y2O3 in the Y-doped SnO2 nanofibers for selective NO2 adsorption in the presence of water vapor. (b) SEM images of the 1 at% Y-doped SnO2 nanofibers. (c) TEM
image of the 1 at% Y-doped SnO2 nanofiber. (d) Magnified TEM image of the 1 at% Y-doped SnO2 nanofiber. (Third step) Dynamic resistance curves of (e) pristine
SnO2 nanofibers, (f) the 0.5 at% Y-doped SnO2 nanofibers (Y0.5-SnO2), (g) the 1 at% Y-doped SnO2 nanofibers (Y1-SnO2), (h) the 5 at% Y-doped SnO2 nanofibers
(Y5-SnO2) to 10 ppm NO2 at 200 ◦ C in dry and humid conditions (RH = 0%, 17%, 52% and 82%). Resistance value of each sample in (i) air and (j) analyte gas as a
function of RH. (k) The ΔS value of the pristine and Y-doped sensors. (Fourth step) Optimized (110) slab structures of (l) pristine and (m) Y-doped SnO2. Blue-dotted
circles indicate the positions of two VO defects. Optimized (110) slab structures for NO2 adsorption in (n) pristine and (o) Y-doped SnO2. Optimized (110) slab
structures for H2O adsorption in (p) pristine and (q) Y-doped SnO2 [77].
Reprinted with permission from ref [77] Copyright 2021 Elsevier.

Table 2
Comparison of previously reported humidity-independent NO2 detection data [77], [78–83]. Reprinted with permission from ref [77]. Copyright 2021 Elsevier.
Sensing Materials Temp. (℃) NO2 conc. (ppm) Response (Ra/Rg or Rg/Ra) Response changes (ΔS) Reference

dry wet

ZnO thin film 450 200 94 70 (RH 10%) -0.26 [78]


WS2/graphene aerogel 180 2 ~3.3 ~3.0 (RH 60%) -0.09 [79]
Al-doped ZnO 300 5 ~8 ~2.4 (RH 40%) -0.7 [80]
Cr2WO6 200 2 ~8 ~11 (RH 20%) 0.38 [81]
Rb2CO3-decorated In2O3 RT 0.25 ~65 ~140 (RH 62%) 1.16 [82]
Rb-doped In2O3 75 5 1502 660 (RH 94%) -0.56 [83]
1 at% Y-doped SnO2 nanofibers 200 10 70.38 97.94 (RH 87%) 0.39 [77]
5 at% Y-doped SnO2 nanofibers 200 10 11.06 12.17 (RH 87%) 0.1 [77]

favorable for reacting with (and thereby detecting) NO2 molecules than diameter, a few micrometers in length, and exhibited homogeneous
with water vapor. Moreover, the yttrium atoms that were not substituted morphology [85]. The HRTEM image showed the single-crystalline na­
into the SnO2 lattice formed hydrophobic Y2O3 nanoparticles over the ture of the ZnO nanowire. The interspacing between the two crystal
entire surface of the nanofibers, which acted as synergistic co-agents for planes was 0.52 nm, indicating that the preferred growth direction was
repelling water vapor. The water contact angles of the pristine SnO2 the 〈0001〉 axis [89]. The WO3 nanowires were 350 nm in diameter,
nanofibers and the 0.5, 1, 3, 5, and 10 at% Y-doped SnO2 nanofibers several micrometers in length, and exhibited a rough morphology with a
were estimated to be 80.4◦ and 94.5◦ , 107.8◦ , 121.7◦ , 144.7◦ , and nonhomogeneous diameter in which the bottom was larger than the top.
161.6◦ , respectively; essentially, the angle increased with increasing In addition, the WO3 nanowires were polygonal rather than circular, as
concentration of Y. The authors concluded that hydrophobicity could be evidenced by their vapor–solid growth model, in which the vapor of the
exploited for detecting more target gases without interference from source materials was transported to low-temperature zones and
water molecules, enabling the sensor to preserve its intrinsic sensing condensed to form small crystal clusters, which were seeded into the
ability with improved gas responses, even at high humidity levels. nanowires [90]. The HRTEM image also demonstrated that WO3 pre­
sented single crystallinity with an interspace of 0.38 nm, corresponding
2.1.10. SnO2 nanowires to the (002) planes of the monoclinic structure; thus, the growth di­
[50-ppb-Cl2, LOD: 48 ppt]. rection was determined to be the 〈001〉 axis [91]. The gas response was
Van Dang et al. fabricated SnO2 nanowires for detecting 50 ppb Cl2 at defined as S = RCl2/Rair, where RCl2 and Rair are the resistances of the
50 ◦ C with an LOD of 48 ppt. SnO2, ZnO, and WO3 nanowires were sensor in Cl2 and dry air, respectively. The SnO2 nanowires exhibited
synthesized by CVD, as described henceforth [84]. higher sensitivity to Cl2 gas than that of the ZnO and WO3 nanowires
A sensor chip with a microheater and a pair of Pt/Cr electrodes [85] (Fig. 10b). Moreover, the SnO2 nanowire sensor showed superior
(Fig. 10a− b) was placed 3 cm away from a Sn-powder-containing quartz sensitivity at 50 ◦ C compared with that in the 100–300 ◦ C range
tube. The sensor chip had arrays of catalytic Au islands and SnO2 or ZnO (Fig. 10d− e). The response (RCl2/Rair) of the SnO2 nanowire sensor to 50
nanowires to precisely grow nanowires in predefined areas; catalytic W ppb Cl2 at 50 ◦ C was approximately 57 (Fig. 10e). Additionally, the Cl2
islands were used for fabricating WO3 nanowires. The catalytic islands gas detection limit of the SnO2 nanowire sensor was calculated to be 48
had an average diameter of 5 µm and were periodically arranged be­ ppt at 50 ◦ C (Fig. 10g− h). The sensing mechanism of the SnO2 nanowire
tween the two Pt electrodes. The distance between the catalytic islands sensor was believed to involve direct adsorption of a Cl2 molecule on the
was 5 µm. The samples were heated at 750 ◦ C for 30 min to grow the metal oxide surface and its subsequent substitution with pre-adsorbed
SnO2 nanowires, with vacuum being maintained in the quartz tube oxygen, followed by lattice oxygen.
(~10− 1 Torr) by flowing a mixture of Ar (30 sccm) and O2 (0.5 sccm) The SnO2 nanowires show a superior response to Cl2 gas compared to
during the nanowire growth [86]. The nanowires grew only on the that of other materials such as Pd-doped NiFe2O4 nanocrystals [92],
predefined areas where the catalytic islands were deposited (Fig. 10c), Sb-doped SnO2 nanoparticles [93], In2O3 thin film [94],
forming nanowire–nanowire junctions between the Pt/Cr electrodes Fe2O3-modified ZnO thick film [95], MgO–In2O3 thin film [96], meso­
that acted as the sensing element. The SnO2 nanowires were 80 nm in porous SnO2 [97], Te nanotubes [98], sulfonated copper phthalocyanine
diameter and a few micrometers in length, and had homogeneous films [99], zinc phthalocyanine nanobelts [100], and nanowires of
morphology [86]. The morphologies and sizes of the ZnO and SnO2 copper phthalocyanine [101] (Table 3). Moreover, the Cl2 response of
nanowires were not significantly different. The Au-catalyst-assisted the SnO2 nanowire sensor is comparable to that of other mixed oxide
growth of the ZnO and SnO2 nanowires followed the vapor­ materials, such as CdIn2O4 powder [102], hydrothermally prepared
–liquid–solid mechanism, in which the Au islands had identical thick­ CdSnO3 [103], and CdSnO3 thick film [104]. The superior Cl2 response
nesses, which produced nanowires with similar diameters [87]. The of SnO2 than that of ZnO and WO3 was not clarified; however, this may
HRTEM image of a SnO2 nanowire (inset of Fig. 10c) showed clear lat­ be related to the influence of defect levels in the sensing materials. Ahn
tice fringes with an interspace of 0.26 nm, which corresponded to the et al. [105] suggested that the abundant OVs in the ZnO nanowires
distance between the (101) planes of tetragonal SnO2, indicating that resulted in a superior response to the oxidizing gas. The OV-related peak
the growth direction was 〈101〉 [88]. The ZnO nanowires were 80 nm in in SnO2 was the strongest and broadest in the photoluminescence

12
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 10. (a) Design of a nanowire-based sensor chip. (b) Responses of ZnO, SnO2, and WO3 nanowires to Cl2 gas. (c) SEM image of SnO2 nanowires; inset shows a
corresponding HRTEM image, with the arrow indicating the direction of growth of the nanowires. Sensing characteristics at ppb-level concentrations of Cl2: (d)
Transient resistance of the SnO2 nanowire sensor upon exposure to different concentrations of Cl2 (50–400 ppb) at various temperatures (50–300 ◦ C); (e) sensor
response as a function of Cl2 concentration measured at the investigated temperatures; and (f) selectivity of the sensor to different gases (CO, H2S, ethanol, and NH3)
measured at 400 ◦ C. (g) A fifth-order polynomial fit of ten baseline data points and the (h) linear fit of the response versus Cl2 gas concentration [84].
Reprinted with permission from ref. [84] Copyright 2021 American Chemical Society.

spectra of the SnO2, WO3, and ZnO nanowires. Van Dang et al. indicated and SnS, but at different amounts. The gas sensing response (S) was
that OVs could be responsible for the superior response of the SnO2 defined as follows: (Rgas − Rair)/Rair × 100%. The SnOx/SnS hetero­
nanowires to Cl2 compared to that of the ZnO or WO3 nanowires [84]. structures oxidized at 350 ◦ C for 1 h exhibited remarkable responses of
171% and 2735% at 1 ppb and 1 ppm NO2, respectively, which were
2.1.11. SnOx/SnS heterostructures considerably lower than the ambient air quality guidelines established
[1-ppb-NO2, LOD: 5 ppt]. by the WHO (20 ppb). The theoretical LOD was estimated to be 5 ppt,
Tang et al. demonstrated the detection of 1 ppb NO2 using SnOx/SnS according to the IUPAC definition [108,109]. This value is lower than
heterostructures and obtained an LOD of 5 ppt [106]. The hetero­ those of carbon-nanotube-based, metal-oxide-based, and
structures, which had large surface areas and abundant oxygen va­ metal-sulfide-based sensors fabricated using materials such as WO3
cancies (OVs), were synthesized by the post-oxidation of [110], HOF [111], CNT [112], SnO2 [113], SnO2–SnO [114], SnO2–ZnO
liquid-phase-exfoliated (LPE) SnS nanosheets in air at different oxida­ [115], SnO2/graphene [116], MoS2/SnO2 [117], SnS2/SnO2 [118], SnS2
tion levels, as explained henceforth. [119], SnS [120], and SnS2/SnS [121] (Table 4). The sensor responses to
SnS nanosheets were obtained by liquid-phase exfoliation in an 500 ppm NO, CO, NH3, and H2 were 377%, 181%, − 63%, and − 55%,
ethanol solution containing SnS powder [107] (Fig. 11). The SnS respectively; moreover, the sensor showed a response of 47% to 1 ppm
nanosheets were dip-coated on a sensor substrate with 10-µm-wide gold SO2. Additionally, the sensor exhibited a response of − 700% at 11% RH
interdigitated electrodes with a gap of 5 µm. The specimens were heated and a remarkable response of − 10870% at 75% RH. The SnOx/SnS gas
at 350–450 ◦ C to fabricate the SnOx/SnS-based gas sensor. XRD analysis sensor exhibited high selectivity in low-RH environments. XRD, XPS,
indicated that all heated samples contained SnO, SnO2, Sn2O3, Sn3O4, SEM, TEM, electron paramagnetic resonance (EPR), and first-principles

13
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Table 3 the crystal growth of the SnO2-nanosheet-assembled film was based on


Previously reported responses of sensing materials to chlorine gas [84], dendritic growth. The branch angles of the nanosheets were determined
[92–104]. Reprinted with permission from ref. [84] Copyright 2021 American based on the crystal structure and growth direction. The nanosheets had
Chemical Society. two types of connections with branch angles of 90◦ (type I) and 46.48◦
Sensing materials Smax (Rgas/ Cl2 T Reference (type II) (Fig. 12d− e). The nanosheets with large (101) and (− 10 − 1)
Rair) (ppm) (℃) facets were modeled using the VESTA program [125] and the crystal
Pd-doped NiFe2O4 nanocrystals 3.33 1000 250 [92] structure data of SnO2 (COD ID:1000062) [126] (Fig. 12f). These two
Sb-doped SnO2 nanoparticles ~750.00 5 RT [93] facets, which were parallel and equivalent, contained surface-bridging
In2O3 thin film 79.30 5 250 [94] oxygen ions. The image shows four facets on the sides of the nano­
CuO- modified ZnO thick film ~194.00 300 RT [95]
MgO–In2O3 thin film ~3.20 7 275 [96]
sheet, which are the equivalent, most stable facets of (110), (− 110), (−
mesoporous SnO2 26.80 5 260 [97] 1 − 10), and (1 − 10). Molecular sensors for environmental hormones in
Te nanotubes 4.00 8 RT [98] solution and hepatocellular carcinoma have been developed using the
sulfonated copper 7.25 2 RT [99] dendritic-structured SnO2 nanosheet film [127].
phthalocyanine films
The coating of a low-heat-resistant polymer film with a metal oxide is
zinc phthalocyanine nanobelts 7.55 1.5 RT [100]
nanowires of copper 7.70 1.5 RT [101] challenging because a metal oxide, such as SnO2, requires a high tem­
phthalocyanine perature of several hundred degrees Celsius for crystallization. Howev­
CdIn2O4 powder 53,000.00 10 255 [102] er, this can be realized by crystallization in an aqueous solution. For
CdSnO3 (hydrothermal 1339.00 5 25 [103] instance, high-temperature annealing is considered indispensable in the
method)
field of ceramics. Cold crystallization, which has been a long-standing
CdSnO3 thick film 8.00 0.1 250 [104]
ZnO nanowires 3.13 5 400 [84] issue, has been realized using the aqueous solution process.
WO3 nanowires 2.90 5 250 [84]
SnO2 nanowires 1785.00 5 50 [84] 3.2. Theoretical calculations of the {110} or {101} facet of SnO2 for gas
sensing
analyses (DFT simulations) indicated that the remarkable response of
the OV-containing heterostructures to the gas molecules was possibly Theoretical calculations of the (110) or (101) facet revealed the ef­
due to the Schottky nature of the metal–SnOx/SnS system, abundant fects of different surfaces on the sensor properties. Zakaryan et al.
adsorption sites, bandgap modulation, and active electron transfer in the studied the adsorption of CO on the (110), (100), (101), and (001)
sensing interface layer. surfaces of SnO2 using DFT calculations [128] (Fig. 13). The electronic
The EPR results revealed that the intensities of the signals in the density of states combined with Bader charge analysis showed that the
samples at 325 ◦ C and 350 ◦ C were considerably higher than those (101) and (001) surface orientations gathered more electrons than the
observed at 400 ◦ C and 450 ◦ C, indicating higher OV contents at 325 ◦ C other orientations, such as (110). Moreover, the (101) or (001) surface
and 350 ◦ C. Therefore, the samples at 325 and 350 ◦ C exhibited stronger of SnO2 was determined to be a superior platform for CO gas sensing.
gas sensing responses. However, NO2 was chemisorbed on the surface of Jiang et al. conducted theoretical calculations of the (110), (101), and
the OV-containing SnOx/SnS heterostructures; the gas molecule (221) facets of SnO2 for NO and NO2 gas sensing [129]. Optimized
desorption increased in difficulty with an increase in the amount of OVs structures of the NO molecules adsorbed on the (110), (101), and (211)
on the sample surfaces. Thus, the gas response of the 325 ◦ C samples facets were obtained. The total energy was calculated using DFT to
could not be easily recovered to their initial state. Moreover, UV illu­ investigate the adsorption of NO molecules. The exchange and correla­
mination was employed to accelerate recovery. tion energies were calculated using the Perdew–Wang (PW) functional
with the generalized gradient approximation (GGA) and a double nu­
merical plus polarization (DNP) basis set. The configuration of the
3. Development of SnO2 nanosheets with cold crystallization
adsorbed NO molecule was more stable in the nitrogen orientation than
in the oxygen orientation on the (101) crystal surface. The adsorption
Morphological control of metal oxides has been actively studied
energy (binding energy) of NO on the (110) facet was the lowest,
[122]. Recently, the cold crystallization of metal oxides has been suc­
whereas that on the (211) facet was the highest. Moreover, the (101)
cessfully realized in aqueous solutions. Achieving control of aspects such
facet was found to be the most beneficial among the three crystal sur­
as crystal structure, morphology, crystal facets, crystallinity, chemical
faces for both NO and NO2 gas detection based on the determined
composition, and defects can offer scientific and industrial benefits.
adsorption energies and electron transfer. Abokifa et al. investigated
SnO2 nanosheets and dendritic-structured SnO2 nanosheet films are
the sensing mechanism of SnO2 for detecting ethanol and acetone at
discussed in this section.
room temperature using first-principles DFT calculations and ab initio
molecular dynamics (AIMD) simulations [130]. Both the ethanol and
3.1. Dendritic-structured SnO2 nanosheet film acetone molecules had stronger interactions with the pre-adsorbed O−2
species on the reduced (101) surface than on the reduced (110) surface
Cold crystallization of SnO2 was first realized in 2009 [123,124]. at room temperature. The most cited sensing mechanism involves in­
SnO2 is typically synthesized by the reaction of SnF2 and H2O in an teractions with the ionosorbed O− species that exhibit high activity for
aqueous solution. In addition, crystallization of SnO2, which typically oxidizing the target gas molecules. However, the less active superoxide
requires a high temperature of several hundred degrees Celsius, has been molecule (O−2 ) constituted the majority of the pre-adsorbed oxygen
realized at lower temperatures. Because the most stable crystal facet of species on the SnO2 surface at room temperature. Neither the ethanol
SnO2 is {110}, particles surrounded by {110} crystal facets are typically nor acetone molecules could exothermically bind to the pre-occupied
produced. However, SnO2 nanosheets with {101} crystal facets, which vacancies and showed no interactions with the pre-adsorbed O−2 spe­
are metastable crystal facets, can be synthesized by the facet-controlled cies on the reduced (110) surface. In contrast, the OV sites simulta­
growth method. Furthermore, a dendritic-structured SnO2 nanosheet neously accommodated the O−2 species and the ethanol or acetone
film has been prepared by controlling the nucleation and crystal growth molecules on the reduced (101) surface, forcing the pre-adsorbed O−2
in the solution; details of the dendritic crystal growth were clarified by species to release the minor charge back to the surface. AIMD simula­
FE-SEM (Fig. 12a− b), high-resolution TEM (Fig. 12c), and crystal tions demonstrated that both molecules preferentially bonded to the
structure models [17]. The models that were analyzed using fast Fourier unoccupied OV sites (that is, at the undercoordinated Sn4c). Once all the
transform (FFT) images and the calculated SAED pattern indicated that OV sites were filled with the pre-adsorbed oxygen from the ambient

14
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 11. (a) Illustration of the fabrication of the SnOx/SnS-hetero­


structure-based gas sensor. (b) XRD patterns of the as-synthesized sam­
ples. (c) SEM, (d) TEM, and (e) high-resolution TEM (HRTEM) images of
the SnOx/SnS heterostructures; the inset image in (e) shows selected area
electron diffraction (SAED) rings. Band structures of (f) SnS and (g) SnOx
determined using first-principles calculations. (h) Calculated band
alignment between SnS and SnOx, and the position of the work function
of Au. The black dashed lines represent the Fermi level, which is a
Schottky contact. Energy band structures of the SnOx/SnS heterojunction
contact with Au in (i) air and (j) a NO2 atmosphere. (k) Formation of OV-
rich SnOx/SnS heterostructure. (l) Charge density difference (CDD) plot
showing NO2 on the SnOx/SnS heterostructure. The blue and yellow re­
gions indicate charge accumulation and depletion, respectively. (m)
Electron localization function (ELF) plot of NO2 on the SnOx/SnS heter­
ostructure; an isosurface value of 0.015 e Å− 3 was used [106].
Reprinted with permission from ref [106] Copyright 2021 IOP
Publishing.

15
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Table 4
Comparison of various gas sensors with respect to the range of detected NO2, operating temperature, response and recovery times, and response [106], [110–121].
Reprinted with permission from ref [106]. Copyright 2021 IOP Publishing.
Sensing NO2 conc. LOD Temperature Response Recovery Response Reference
materials (ppb) (ppb) (℃) time (s) time (s) (%)

WO3 500 10 200 ~3600 ~1800 150 [110]


HOF 100 40 25 17.6 19.1 ~6.9 [111]
CNT 1 0.1 25 120 — ~25 [112]
SnO2 1 0.2 25 ~1800 ~900 ~90 [113]
SnO2–SnO 100 100 50 ~150 ~400 26 ppm− 1 [114]
SnO2–ZnO 400 — 200 ~300 ~300 600 [115]
SnO2/graphene 10 0.024 150 43 408 200–300 [116]
MoS2/SnO2 500 — 25 408 162 0.6 [117]
SnS2/SnO2 125 — 100 299 143 ~90 [118]
SnS2 10,000 — 120 ~170 ~140 ~3600 [119]
SnS 1000 — 25 ~25 ~25 60 [120]
SnS2/SnS 400 75 25 365 1216 660 [121]
SnOx/SnS 1 0.005 25 1800 36 (UV) 171 [106]
SnOx/SnS 1000 0.005 25 1800 36 (UV) 2735 [106]

CNT: carbon nanotube, HOF: hydrogen-bonded organic framework material.

atmosphere, the target molecule tended to momentarily share the pre­ 3.4. SnO2-nanosheet-based MEMS
occupied vacancies with the pre-adsorbed oxygen molecules prior to its
eventual binding to the surface Sn5c sites. Therefore, the adsorption [lung cancer detection].
configuration resembled that of stoichiometric surfaces. Additionally, An MEMS-type gas sensor was subsequently developed using SnO2
the authors reported that adsorption was generally stronger for ethanol nanosheets [133], which were crystallized on a Si3N4 (SIN) substrate on
than for acetone because of the bipolar nature of the hydroxyl group (in which a Pt electrode and a Pt heater were formed. The MEMS sensor was
ethanol) that interacted with the surface via two distinct charge-transfer evaluated at 250 ◦ C (Fig. 15). The SnO2-nanosheet-type sensors
modes. These results suggest that the sensing mechanism of SnO2 for exhibited superior response (Ra/Rg) to nonanal (C9H18O), a lung cancer
polar VOCs at room temperature can be explained by their direct marker, compared to that of CO, nitrogen dioxide (NO2), acetone
adsorption on the surface rather than through oxidation by the ion­ (CH3COCH3), hydrogen (H2), ethanol (C2H6O), ammonia (NH3),
osorbed oxygen species. Feng et al. investigated the relationship be­ hydrogen sulfide (H2S), formaldehyde (HCHO), acetaldehyde
tween the morphology and humidity sensing properties of SnO2 using (CH3CHO), and butyraldehyde (C4H8O). In addition, the responses to
DFT calculations [131]. The calculations indicated that the {101} facet various aldehydes were evaluated. The results showed that the response
adsorbed more water molecules than the {110} facet. This result was increased with increasing molecular weight in the order of formalde­
consistent with the experimental data obtained using a nanosensor hyde (HCHO), acetaldehyde (CH3CHO), butyraldehyde (C4H8O,
based on three-dimensional (3D) hierarchical SnO2 dodecahedral C3H7CHO), and nonanal (C9H18O, C8H17CHO). The response to nonanal
nanocrystals (DNCs), which exhibited superior humidity-sensing per­ increased from 1.383 for 100 ppb to 2.00 for 300 ppb, showing a strong
formance, including fast response and recovery times, narrow hysteresis concentration dependence. The high sensitivity and molecular recog­
loop, high sensitivity, excellent linearity response, and decent stability, nition characteristics of the SnO2 nanosheets to nonanal molecules
compared with those of 3D hierarchical SnO2 nanorods (NRs) and SnO2 enable the early detection of lung cancer.
nanoparticles (NPs). The authors attributed the enhanced sensing
properties of the DNCs to the peculiar 3D open nanostructures and high
chemical activity of the exposed {101} facet. In summary, 3D open 3.5. H2/CH4 sensing properties of the dendritic-structured SnO2
nanostructures can promote the penetration and diffusion of water nanosheet film
molecules, and the exposed {101} facet can facilitate their adsorption.
The most thermodynamically stable crystal facet of tin oxide is
3.3. SnO2 nanosheet/SnO2 nanoparticle film with Pt, Au, and Pd for 1- {110}. However, the large surface of the SnO2 nanosheets has a meta­
nonanal/lung cancer detection stable {101} facet, which is a characteristic feature of these nanosheets.
Choi et al. studied the anisotropic growth of SnO2 nanosheets and the
Several types of gas sensors and odor sensors have been developed effects of crystal facets on gas sensing properties [134] (Fig. 16).
using dendritic-structured SnO2 nanosheet films, as discussed below. Time-varying XRD analysis of the crystallite size revealed that the ver­
These dendritic-structured films were was shown to exhibit excellent tical growth of the {101} facet of the SnO2 nanosheets was suppressed
performance as a gas sensor and as a molecular sensor in a solution. and that the growth proceeded in the in-plane direction. In addition, XPS
The exhaled breath of lung cancer patients is known to contain a analysis revealed that the SnO2 nanosheets contained Sn2+ in addition to
higher concentration of nonanal molecules than that of healthy in­ Sn4+. Moreover, the presence of several OVs in the SnO2 nanosheets was
dividuals. Therefore, the use of gas sensors to screen for lung cancer can confirmed. These features assisted in enabling the excellent sensor
offer tremendous benefits. A representative nonanal-detecting gas response. Furthermore, the H2 and CH4 gas sensing characteristics were
sensor was fabricated by first preparing a SnO2-nanoparticle film con­ evaluated using rectangular SnO2 particles with the {110} facet and
taining nanoparticles of Pt, Au, and Pd as a catalyst on the sensor sub­ SnO2 nanosheets. The response to CH4 was 1.5 for the SnO2 particles
strate as the first layer [132] (Fig. 14a). The substrate was immersed in a with the {110} facet, whereas that of the SnO2 nanosheets was slightly
SnF2 aqueous solution to grow SnO2 nanosheets on the surface of the lower at 1.3. However, the response of the SnO2 particles to H2 was 3.0,
SnO2-nanoparticle film. The gas sensor showed an increased response whereas that of the SnO2 nanosheets was 7.5, which was more than
with increasing concentration of nonanal. In particular, the SnO2 double the former. Moreover, the 90% response time (T90 response) and
nanosheet showed an improved rate of increase in the response to an 90% recovery time (T90 recovery) to reach 90% of the saturation
increase in the nonanal concentration from 1 to 10 ppm (Fig. 14b). resistance in the gas and air were evaluated. The SnO2 particles showed
T90 responses of 24 s and 18 s for H2 or CH4, respectively, whereas the
SnO2 nanosheets showed fast response times of 18 s and 6 s,

16
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 12. Surface (a) and cross-sectional (b) SEM images of the SnO2
(101)-nanosheet-assembled film. (c) Cross-sectional TEM image of the
SnO2 (101)-nanosheet-assembled film. (d) Cross-sectional TEM image of
the gradient structure of the SnO2 (101)-nanosheet-assembled film on an
FTO substrate with branch angles of 90◦ (type I; red circle) and 46.48◦
(type II; blue circle) between the SnO2 nanosheets. (e) Crystal growth
models of the SnO2 nanosheets that were connected at branch angles of
90◦ (right, red circle) and 46.48◦ (left, blue circle), respectively. The
models were observed from the direction of B = <1 − 1 − 1 > of SnO2.
(f) Model of the SnO2 nanosheet with large (101) and (− 10 − 1) facets
[17]. (For interpretation of the references to color in this figure legend,
the reader is referred to the web version of this article.)
Reprinted with permission from ref [17] Copyright 2021 Springer
Nature.

17
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

respectively. The SnO2 particles showed T90 recoveries of 36 s and 44 s


for H2 and CH4, respectively, whereas the SnO2 nanosheets showed fast
recovery times of 28 s and 12 s, respectively. Further clarification of the
characteristics of the SnO2 nanosheets can facilitate the development of
highly sensitive gas sensors and odor sensors.

3.6. The dendritic-structured SnO2 nanosheet film

[200-ppb-acetone sensing].
Detection of low-concentration acetone in gases that are exhaled and
released via skin can enable early detection of diabetes. In addition,
acetone detection is crucial for monitoring odors in rooms. Kim et al.
developed a low-concentration acetone gas sensor using a dendritic-
structured SnO2 nanosheet film [135] (Fig. 17a). The size and aspect
ratio of the SnO2 nanosheets were controlled by varying the duration of
synthesis. The effects of the synthesis period on sensor characteristics
were evaluated. The dendritic-structured SnO2 nanosheet film synthe­
sized by crystal growth over 6 h exhibited a superior response than those
of films synthesized for 2 h or 24 h. The operating temperature of the
sensor was investigated in the range of 200–300 ◦ C, with an excellent
response being obtained at 280 ◦ C. Detection of low-concentration
acetone gas was realized using the dendritic-structured SnO2 nano­
sheet film synthesized for 6 h (Fig. 17b− d). A response value of
approximately 3 was achieved for 200 ppb acetone (Fig. 17e− f). Addi­
tionally, the response to 1 ppm acetone reached 10.2, which is higher
than the acetone response values of the other SnO2 sensors, such as
porous flower-like SnO2 [136], hierarchical SnO2 hollow microspheres
[137], SnO2 microspheres with Au and NiO [138], co-catalyzed SnO2
nanospheres [139], nanotubular SnO2 [140], and mesoporous SnO2
[141] (Table 5). In addition, the gas selectivity was analyzed. The
response to 1 ppm acetone (10.2) was higher than those for 1 ppm
acetaldehyde, isoprene, xylene, cyclohexane, toluene, or ammonia (1–3;
Fig. 17g− h). As mentioned earlier, the dendritic-structured SnO2
nanosheet film was confirmed to exhibit a high sensitivity for detecting
low-concentration acetone (200 ppb) and a high gas selectivity in terms
of selectively detecting acetone from a plurality of gases. The
dendritic-structured SnO2 nanosheet film can be effectively employed to
detect diabetes and for indoor air management.

3.7. Bridge-type dendritic-structured SnO2 nanosheet film

3.7.1. [lung cancer detection]


Choi et al. developed a bridge-type dendritic-structured SnO2
nanosheet film [142], which was separated from the substrate by
approximately 100 nm (Fig. 18a). A SiO2 substrate with comb-shaped Pt
electrodes was used for sensing. The space was formed between the
bridge-type dendritic-structured SnO2 nanosheet film and the SiO2
substrate owing to the etching effect of SiO2 in the aqueous SnF2 solu­
tion. For comparison, a SnO2 dendritic-structured nanosheet film was
also formed on an alumina substrate with a comb-shaped Pt electrode
(Fig. 18b). Approximately 150-nm-thick dendritic-structured SnO2
nanosheet films were synthesized in both scenarios by immersion in an
aqueous SnF2 solution (28 mM, 90 ◦ C) for 6 h. The effects of the bridge
structure were examined by comparing the two SnO2 nanosheet films
(Fig. 18c). The bridged version exhibited superior responses to hydrogen
(500 ppm), methane (500 ppm), and nonanal (0.85 ppm) than those of
the SnO2-based film on the alumina substrate. Both sides of the bridge
structure membrane were presumed to participate in the reaction with
the gas. In addition, the space under the bridge structure was considered
to function as a reaction field for additional oxidation of partially
oxidized molecules. The bridge-type dendritic-structured SnO2 nano­
sheet film showed a stronger response to hydrogen than to methane. In
Fig. 13. Different surfaces of SnO2 featuring possible adsorption sites: (a) - addition, considering the employed gas concentration, the response to
(110), (b) - (100), (c) - (101), and (d) - (001) [128].
nonanal was considerably superior to those of hydrogen and methane.
Reprinted with permission from ref [128] Copyright 2021 International Fre­
The bridge-type dendritic-structured SnO2 nanosheet film had a unique
quency Sensor Association (IFSA) Publishing.
structure, and both sides of the film were in contact with the gas, which

18
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 14. (a) Schematic view of the reaction of a 1-nonanal molecule on a sensor consisting of SnO2 nanosheets, SnO2 nanoparticles, and noble metal catalysts. (b)
Changes in sensitivity with 1-nonanal gas concentration for NS(6 h)/PF (solid circles and lines) and PF (open circles and dotted lines) at 250 ◦ C and 300 ◦ C (blue and
red profiles, respectively; NS: SnO2 nanosheet, PF: SnO2 particulate film). (For interpretation of the references to color in this figure legend, the reader is referred to
the web version of this article.)
[132] Reprinted with permission from ref [132] Copyright 2021 Springer Nature.

Fig. 15. (a) Photograph of a sensor containing


three micro sensor tips. (b) Photograph of
sensor tip array, with the inset showing a
magnified photograph of the sensor tip. (c) SEM
micrograph of the sensor tip. Magnified SEM
micrograph of the sensor tip showing the SiN
(d) and Pt areas (e). B, Magnified SEM micro­
graph of SnO2 nanosheets in the SiN area. (f)
Sensor sensitivities to carbon monoxide, nitro­
gen dioxide, acetone, hydrogen, ethanol,
ammonia, hydrogen sulfide, formaldehyde,
acetaldehyde, and nonanal; a nonanal concen­
tration of 0.1 ppm at 250 ◦ C was used. (g)
Change in sensitivity with aldehyde carbon
number in a series featuring formaldehyde,
acetaldehyde, butanal, and nonanal; a nonanal
concentration of 0.3 ppm at 250 ◦ C was used
[133].
Reprinted with permission from ref. [133]
Copyright 2021 John Wiley and Sons.

facilitated the reaction with the target molecules. In addition, the film 3.8. Surface molecular separator for selective gas sensing
had a space between the substrate for gas trapping. These characteristics
are effective for sensing molecules that have a large molecular weight Choi et al. introduced a metal–organic framework (MOF) layer of
and cannot readily undergo complete oxidation, such as nonanal. zeolitic imidazolate framework-8 (ZIF-8) on the surface of a SnO2
nanosheet sensor [143]. The MOF layer functioned as a surface

19
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 16. (a) Variation in crystallite size perpendicular to the (110), (101), and (211) planes of a SnO2 nanosheet as a function of growth period. (b) Primary growth
direction of a SnO2 nanosheet. XPS profiles of the SnO2 nanosheet and the SnO2 nanoparticle gas sensors: O-1 s data of the (c) nanosheets and (d) nanoparticles.
Cross-sections of crystal structure models of the (e) (110) surface of the SnO2 nanoparticles and the (f) (101) surface of the SnO2 nanosheets [134].
Reprinted with permission from ref [134] Copyright 2021 Elsevier.

molecular separator for selective gas sensing, and the gas selectivity of (In2O3:ZIF-8 = 4:1) for 1 ppm NO2 was 0.3 [146]. The
the sensor was analyzed, as described below. operating-temperature-dependent results for 1 ppm NO2 showed that
The bridge-type SnO2 nanosheet sensors were dipped into a 0.1 M the optimal testing temperature for In2O3/ZIF-8 was 140 ◦ C, which was
aqueous zinc nitrate solution and heated at 200 ◦ C for 6 h to form a ZnO slightly lower than that for pristine In2O3 nanofibers (160 ◦ C). The
seed layer. The resulting product was immersed in an aqueous solution In2O3/ZIF-8 (4:1)-based sensor exhibited the highest NO2 response,
containing zinc nitrate hexahydrate and 2-methylimidazole at 50 ◦ C for followed by In2O3/ZIF-8 (2:1), In2O3/ZIF-8 (8:1), and pristine In2O3 NFs
4 h to grow a ZIF-8 layer (Fig. 19). ZIF-8 operates as a molecular sieve at 140 ◦ C (Rg/Ra = 16.4, 10.7, 10.4, and 4.9, respectively). Moreover,
for approximately 4.0–4.2-Å-sized particles. Compared to the SnO2- the response ratio of the ZIF-8/SnO2 nanosheets to 50 ppm C2H4
nanosheet-based gas sensor, the ZIF-8/SnO2-nanosheet equivalent reached 0.85. The sensor signal response to C2H4, a small gas molecule,
exhibited a decreased signal response for large molecules, for example, increased after the introduction of ZIF-8, similar to trends described in
C3H6 (4.120 Å), 1-C4H8 (5.430 Å), and C9H18O (12.011 Å) (Fig. 19g− i). previous reports [145–147]. This enhancement was presumably caused
However, the response of the ZIF-8/SnO2-nanosheet-based sensor to by the improvement of the activity, which was attributed to the changes
smaller gas molecules, such as C2H4 (3.058 Å), slightly increased owing in electrons at the interface between the sensor material and ZIF-8.
to electron transfer at the ZIF-8–SnO2 interface. The control of gas In addition, the separation effect has been reported in previous
selectivity based on molecular size was realized by introducing ZIF-8 as studies. The responses of pristine ZnO nanowires to 10, 30, and
a surface molecular separator. 50 ppm H2were 2.34, 2.47, and 2.62, respectively, [148] which
Gas selectivity achieved via introduction of ZIF-8 can be classified decreased to 1.17, 1.34, and 1.44, respectively, for the ZnO@ZIF-8
into two categories (Table 6); one involves an increase in the sensor nanowires. The response ratios (SM/SM+ZIF-8) for 10, 30, and
signal response to target gases, and the other involves a decrease in the 50 ppm H2 were 2, 1.84, and 1.82, respectively. The pristine ZnO
signal response to nontarget gases. In the former case, ZIF-8 contributes nanowires responded to C6H6 and C7H8; however, the responses of the
to an enhancement of the gas selectivity to relatively small gas mole­ ZnO@ZIF-8 nanowires to C6H6 and C7H8 almost vanished and
cules, such as H2 and NO2. This enhancement effect can be monitored approached 1.0. The response ratios (SM/SM+ZIF-8) for 10, 30, and
using the response ratio (SM/SM+ZIF-8), which is calculated using the 50 ppm C6H6 and 10, 30, and 50 ppm C7H8 were calculated to be 1.33,
sensor signal responses before (SM) and after (SM/SM+ZIF-8) the 1.43, 1.58, and 1.24, 1.29, and 1.40, respectively. The response ratios of
introduction of ZIF-8. The response ratios of ZnO@ZIF-8 for 5, 10, 50, the ZnO@ZIF-8 nanorod arrays for 50 ppm H2, NH3, CH3CH2OH,
100, 200, and 500 ppm CH3CH2OH were reported to be 0.92, 0.62, 0.40, CH3COCH3, and C6H6 were 1.12, 1.25, 5.5, 7, and 7, respectively, [149].
0.41, 0.43, and 0.41, respectively [144], whereas those of pristine ZnO ZIF-8-coated Pd/ZnO nanowires were reported to have response ratios
were approximately 3.39, 4.80, 11.27, 14.54, 19.98, and 32.0, respec­ of 1.28, 2.75, 4.70, 3.50, and 3.30 for 50 ppm H2, CH3CH2OH,
tively. In contrast, responses of approximately 3.68, 7.70, 28.13, 35.90, CH3COCH3, C6H6, and C7H8, respectively; [150] those of the ZIF-8/SnO2
46.65, and 78.0 were observed at the aforementioned ethanol concen­ nanosheets (SM/SM+ZIF-8) for 50 ppm C3H6, C4H8, and C9H18O were
trations for the ZnO@ZIF-8 core–shell material. Moreover, the response 1.18, 1.67, and 2.15, respectively. In the case of a decreasing signal
ratio of ZnO@ZIF-8 nanorods to 50 ppm H2 was 0.67 [145]. The response, ZIF-8 enabled the separation by operating as a molecular
ZnO@ZIF-8 nanorods exhibited a higher H2 sensitivity than that of the sieve. Although decreases were observed in the sensor signal responses
pristine ZnO nanorods at 250 ◦ C (Ra/Rg = 2.15 and 3.28, respectively, at to both large and small gas molecules, the sensor signal response to large
50 ppm). The response ratio of ZIF-8-coated In2O3 composite fibers gas molecules decreased considerably. In addition, no sensor signal

20
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 17. (a) Schematic of SnO2 nano­


sheet. (b) Cross-sectional TEM image of
the dendritic-structured SnO2 nanosheet
film, (c) HR-TEM image of a single
nanosheet, and (d) atomic arrangements
in a single nanosheet. (e) Dynamic
response curve of NS-6 from 200 to 1000
ppb, (f) responses of NS-6 from 200 to
1000 ppb, (g) responses of NS-
6–1000 ppb acetone, acetaldehyde,
isoprene, xylene, toluene, and ammonia,
and (h) appearance energies of acetone,
acetaldehyde, isoprene, xylene, and
toluene (excluding cyclohexane and
ammonia) for the dissociation of a
methyl cation [135].
Reprinted with permission from ref.
[135] Copyright 2021 American Chemi­
cal Society.

21
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Table 5
Various types of SnO2-based acetone gas sensors and their responses to specific
concentrations [135], [136–141]. Reprinted with permission from ref [135].
Copyright 2021 American Chemical Society.
Sensing material Acetone conc. Response Refecence
(ppm)

porous flower-like SnO2 20 30 [136]


hierarchical SnO2 hollow microspheres 20 2.4 [137]
SnO2 microsphere with Au and NiO 1 5 [138]
Co-catalyzed SnO2 nanosphere 1 1.8 [139]
nanotubular SnO2 1 1.6 [140]
mesoporous SnO2 1 1.5 [141]
SnO2 nanosheet 1 10.2 [135]

responses were recorded for certain large gas molecules after ZIF-8
introduction under several conditions [148,150]. In these studies, the
sensor signal response to large gas molecules, such as C3H6, C4H8, and
C9H18O, decreased after ZIF-8 introduction, which was mainly attrib­
uted to the molecular sieving effect of ZIF-8. The dominant effect of the
enhancement or separation of ZIF-8 on gas sensing is possibly influenced
by various factors such as thickness and operating temperature. By
controlling these factors, an effective improvement in gas selectivity can
be realized via the introduction of a molecular separation layer, such as
ZIF-8. This variation in selectivity with the molecular size of the gas can
facilitate the identification of gas species by data processing. In partic­
ular, the ZIF-8/SnO2 nanosheet thin film can be applied in the medical
field to improve the reliability of selective detection of biomarkers, such
as the C9H18O lung cancer biomarker.

4. Conclusion

Various SnO2 nanomaterials have been developed to improve the


sensitivity, response speed, selectivity, and durability of sensors. Low-
concentration gas sensing is crucial for the detection of biological
gases and spatial odor, which require further improvements in sensi­
tivity. Control of nanostructures, crystal planes, and defects is pro­
gressing, and the acquired information can be leveraged for the
development of materials other than SnO2. This review is anticipated to
facilitate the development of advanced high-sensitivity gas sensors and
odor sensors for solving problems related to healthcare, indoor air
control, food transportation and storage, and plant safety and
construction.
Recent advances in SnO2 nanostructure based gas sensors suggest
Fig. 18. Schematic illustrations of sensing reaction for a (a) SnO2 nanosheet
valuable guidelines for development of SnO2 nanostrures. It is revealed thin film with a bridge-type structure (BS) and a (b) typical SnO2 nanosheet thin
that dimension and morphology affect the sensing performance with film on a Pt interdigitated-electrodes-printed Al2O3 substrate (NS). (c) Sensor
reported studies. The following are important guidelines. Crystallite response ratios of BS to NS for 500 ppm H2, 500 ppm CH4, and
sizes of oxide semiconductors should be as small as possible. Sensitizers 0.85 ppm C9H18O gases [142].
should be dispersed as finely as possible. Sensing layer thickness and Reprinted with permission from ref. [142] Copyright 2021 Elsevier.
porosity should also be optimized to improve selectivity and durability.
Furthermore, state-of-the-art liquid phase crystal growth control nanostructures, etc. can be applied to improve the selectivity. Surface
technology has made it possible to control the SnO2 nanostructure. This modification with nanomaterials including metal oxides, polymers, self-
has made it possible to realize the following design guidelines for the assembled monolayers, carbon materials, etc. can be also utilized. MOF
development of SnO2 nanomaterials. SnO2 nanomaterials having nano- surface molecular separator, p-n hybrid structures, temperature control
meter size which is similar to size of a depletion layer, and controlled of the sensor, excitation light will be helpful. Additionally, one of the
microstructure such as a nanosheet structure shows high sensing per­ resolutions for the issue is sensor array which includes various type of
formance. The dendritic structure in which 2D nanosheets are connected metal oxide semiconductor-type gas sensors with/without different type
by crystal growth points the direction for future sensor development. sensors. Sensing signals, which are obtained from the sensors, can be
Furthermore, the high gas adsorption performance and reactivity of the analyzed to recognize gas or odor automatically with AI or machine
metastable crystal plane will be a guide for future sensor development. learning techniques. Furthermore, MEMS type sensors have advantages
Metal oxide semiconductor-type gas sensors have advantages for in low power consumption, miniaturization, and weight reduction. The
detecting low concentration gases. However, gas selectivity of them is heating elements are connected by narrow bridges to float in some
not enough high. Control of surface crystal facet, crystal defects, MEMS devices. Therefore, the MEMS devices have the advantage that

22
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Fig. 19. (a) Schematic diagram of the interactions between ZIF-8 and SnO2 nanosheets. (b) Surface FE-SEM image and (c) a cross-section TEM image of the ZIF-8/
SnO2 nanosheet thin film. (d) HAADF-STEM image and STEM-EDS mappings of the ZIF-8/SnO2 nanosheet thin film: Sn (e) and Zn (f). Variations in the (g) electrical
resistance, (h) sensor signal response, and (i) response ratio of the SnO2 nanosheet thin-film and ZIF-8/SnO2 nanosheet thin-film gas sensors for 50 ppm C2H4, C3H6,
and C4H8, and 0.85 ppm C9H18O. (j) Crystal structure of ZIF-8 and molecular structures of (k) C2H4, (l) C3H6, (m) C4H8, and (n) C9H18O [143].
Reprinted with permission from ref [143] Copyright 2021 American Chemical Society.

23
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

Table 6
Comparison of previously reported ZIF-8-incorporated metal oxide gas sensors [143], [144–146,148–150]. Reprinted with permission from ref [143]. Copyright 2021
American Chemical Society.
Sensing materials Temp. Target gas Conc. Response ratioa Ref.
(℃) (ppm) (SM/SM+ZIF-8),
Response of SM,
Response of SM+ZIF-8

enhancement ZnO@ZIF-8 160 CH3CH2OH 5 0.92 (3.39/3.68) [144]


10 0.62 (4.80/7.70)
50 0.40 (11.27/28.13)
100 0.41 (14.54/35.90)
200 0.43 (19.98/46.65)
500 0.41 (32.0/78.0)
ZnO@ZIF-8 nanorods 250 H2 50 0.66 (2.15/3.28) [145]
ZIF-8-coated In2O3 composite fibers 140 NO2 [146]
(In2O3:ZIF-8 =4:1) 1 0.30 (4.9/16.4)
(In2O3:ZIF-8 =2:1) 1 0.46 (4.9/10.7)
(In2O3:ZIF-8 =8:1) 1 0.47 (4.9/10.4)
ZIF-8/SnO2 nanosheets 100 C2H4 50 0.85 [143]
separation ZnO@ZIF-8 nanowires 300 H2 10 2 (2.34/1.17) [148]
30 1.84 (2.47/1.34)
50 1.82 (2.62/1.44)
C6H6 10 1.33b (1.33/1)
30 1.43b (1.43/1)
50 1.58b (1.58/1)
C7H8 10 1.24b (1.24/1)
30 1.29b (1.29/1)
50 1.40b (1.40/1)
ZnO@ZIF-8 nanorod arrays 250 H2 50 1.12 [149]
NH3 50 1.25
CH3CH2OH 50 5.5
CH3COCH3 50 7
C6H6 50 7
ZIF-8-coated Pd/ZnO nanowires 200 H2 50 1.28 [150]
CH3CH2OH 50 2.75b
CH3COCH3 50 4.70b
C6H6 50 3.50b
C7H8 50 3.30b
ZIF-8/SnO2 nanosheets 100 C3H6 50 1.18 [143]
C4H8 50 1.67
C9H18O 50 2.15
a
SM and SM+ZIF are the sensor signal responses before and after introduction of ZIF-8, respectively.
b
No sensor signal response after incorporation of ZIF-8.

they do not get hot and do not affect the surrounding devices. Some of [9] M. Madou, S.R. Morrison, Chemical Sensing with Solid State Devices, Academic
Press, Boston, 1989.
these techniques have been introduced in this review. I hope that these
[10] G. Korotcenkov, Chemical Sensors: Fundamentals of Sensing Materials,
technologies will help to detect low-concentration gases and odors in the Momentum Press, New Jersey, 2011.
healthcare and environmental fields. [11] H.-J. Kim, J.-H. Lee, Highly Sensitive and Selective Gas Sensors Using p-type
Oxide Semiconductors: Overview, Sens. Actuators B: Chem. 192 (2014) 607–627.
[12] M. Osada, T. Sasaki, Two-dimensional dielectric nanosheets: novel
nanoelectronics from nanocrystal building blocks, Adv. Mater. 24 (2012)
Declaration of Competing Interest 210–228.
[13] Z. Liu, R. Ma, M. Osada, N. Iyi, Y. Ebina, K. Takada, et al., Synthesis, anion
exchange, and delamination of co− al layered double hydroxide: assembly of the
The authors declare that they have no known competing financial
exfoliated nanosheet/polyanion composite films and magneto-optical studies,
interests or personal relationships that could have appeared to influence J. Am. Chem. Soc. 128 (2006) 4872–4880.
the work reported in this paper. [14] M. Osada, T. Sasaki, Exfoliated oxide nanosheets: new solution to
nanoelectronics, J. Mater. Chem. 19 (2009) 2503–2511.
[15] M. Osada, Y. Ebina, H. Funakubo, S. Yokoyama, T. Kiguchi, K. Takada, et al.,
References High-κ dielectric nanofilms fabricated from titania nanosheets, Adv. Mater. 18
(2006) 1023–1027.
[1] K. Toko, Biomimetic Sensor Technology, Cambridge University Press, UK, 2000. [16] M. Osada, K. Akatsuka, Y. Ebina, H. Funakubo, K. Ono, K. Takada, et al., Robust
[2] K. Toko, Biochemical Sensor-mimicking Gustatory and Olfactory Senses, Pan high-κ response in molecularly thin perovskite nanosheets, ACS Nano 4 (2010)
Stanford Publishing, Singapore, 2013. 5225–5232.
[3] Y. Masuda, T. Ito, Development of Odor Monitoring Technology for Living Space, [17] Y. Masuda, Facet controlled growth mechanism of SnO2 (101) nanosheet
Environ. Meas. Technol. 46 (2019) 3–7. assembled film via cold crystallization, Sci. Rep. 11 (2021) 11304.
[4] I. Nikkei, Business Publications, Nikkei Business Publications, Inc, Japan, 2017. [18] Y. Li, N. Chen, D. Deng, X. Xing, X. Xiao, Y. Wang, Formaldehyde detection: SnO2
〈https://project.nikkeibp.co.jp/mirakoto/atcl/wellness/h_vol19/〉. microspheres for formaldehyde gas sensor with high sensitivity, fast response/
[5] K. Toma, M. Tsujii, T. Arakawa, Y. Iwasaki, K. Mitsubayashi, Dual-target gas- recovery and good selectivity, Sens. Actuators B: Chem. 238 (2017) 264–273.
phase biosensor (bio-sniffer) for assessment of lipid metabolism from breath [19] Y. Xu, W. Zheng, X. Liu, L. Zhang, L. Zheng, C. Yang, et al., Platinum single atoms
acetone and isopropanol, Sens. Actuators B: Chem. 329 (2021), 129260. on tin oxide ultrathin films for extremely sensitive gas detection, Mater. Horiz. 7
[6] K. Minami, G. Yoshikawa, Effects of partial attachment at the interface between (2020) 1519–1527.
receptor and substrate on nanomechanical cantilever sensing, Sens. Actuators A: [20] S.-J. Kim, S.-J. Choi, J.-S. Jang, H.-J. Cho, W.-T. Koo, H.L. Tuller, et al.,
Phys. 319 (2021), 112533. Exceptional high-performance of Pt-based bimetallic catalysts for exclusive
[7] N. Yamazoe, K. Shimanoe, 1 - Fundamentals of semiconductor gas sensors, in: detection of exhaled biomarkers, Adv. Mater. 29 (2017), 1700737.
R. Jaaniso, O.K. Tan (Eds.), Semiconductor Gas Sensors, Woodhead Publishing, [21] V.K. Tomer, S. Devi, R. Malik, S.P. Nehra, S. Duhan, Highly sensitive and selective
2013, pp. 3–34. volatile organic amine (VOA) sensors using mesoporous WO3–SnO2 nanohybrids,
[8] J. Watson, The Tin Oxide Gas Sensor and Its Applications, in: Sensors and Sens. Actuators B: Chem. 229 (2016) 321–330.
Actuators, 5, 1984, pp. 29–42.

24
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

[22] D. Ju, H. Xu, Q. Xu, H. Gong, Z. Qiu, J. Guo, et al., High triethylamine-sensing [49] M.M. Arafat, B. Dinan, S.A. Akbar, A.S.M.A. Haseeb, Gas sensors based on one
properties of NiO/SnO2 hollow sphere P–N heterojunction sensors, Sens. dimensional nanostructured metal-oxides: a review, Sensors 12 (2012)
Actuators B: Chem. 215 (2015) 39–44. 7207–7258.
[23] Y. Zou, S. Chen, J. Sun, J. Liu, Y. Che, X. Liu, et al., Highly efficient gas sensor [50] M. Al-Hashem, S. Akbar, P. Morris, Role of oxygen vacancies in nanostructured
using a hollow SnO2 microfiber for triethylamine detection, ACS Sens. 2 (2017) metal-oxide gas sensors: a review, Sens. Actuators B: Chem. 301 (2019), 126845.
897–902. [51] D.R. Miller, S.A. Akbar, P.A. Morris, Nanoscale metal oxide-based heterojunctions
[24] H. Xu, J. Ju, W. Li, J. Zhang, J. Wang, B. Cao, Superior triethylamine-sensing for gas sensing: a review, Sens. Actuators B: Chem. 204 (2014) 250–272.
properties based on TiO2/SnO2 n–n heterojunction nanosheets directly grown on [52] J.M. Walker, S.A. Akbar, P.A. Morris, Synergistic effects in gas sensing
ceramic tubes, Sens. Actuators B: Chem. 228 (2016) 634–642. semiconducting oxide nano-heterostructures: a review, Sens. Actuators B: Chem.
[25] X. Yang, Q. Yu, S. Zhang, P. Sun, H. Lu, X. Yan, et al., Highly sensitive and 286 (2019) 624–640.
selective triethylamine gas sensor based on porous SnO2/Zn2SnO4 composites, [53] P. Karnati, S. Akbar, P.A. Morris, Conduction mechanisms in one dimensional
Sens. Actuators B: Chem. 266 (2018) 213–220. core-shell nanostructures for gas sensing: a review, Sens. Actuators B: Chem. 295
[26] X.-Q. Qiao, Z.-W. Zhang, D.-F. Hou, D.-S. Li, Y. Liu, Y.-Q. Lan, et al., Tunable (2019) 127–143.
MoS2/SnO2 P–N heterojunctions for an efficient trimethylamine gas sensor and [54] J. Walker, P. Karnati, S.A. Akbar, P.A. Morris, Selectivity mechanisms in resistive-
4-nitrophenol reduction catalyst, ACS Sustain Chem. Eng. 6 (2018) 12375–12384. type metal oxide heterostructural gas sensors, Sens. Actuators B: Chem. 355
[27] H. Xu, D. Ju, Z. Chen, R. Han, T. Zhai, H. Yu, et al., A novel hetero-structure (2022), 131242.
sensor based on Au/Mg-doped TiO2/SnO2 nanosheets directly grown on Al2O3 [55] L. Cheng, S.Y. Ma, T.T. Wang, X.B. Li, J. Luo, W.Q. Li, et al., Synthesis and
ceramic tubes, Sens. Actuators B: Chem. 273 (2018) 328–335. characterization of SnO2 hollow nanofibers by electrospinning for ethanol
[28] X. Liu, K. Zhao, X. Sun, C. Zhang, X. Duan, P. Hou, et al., Rational design of sensing properties, Mater. Lett. 131 (2014) 23–26.
sensitivity enhanced and stability improved TEA gas sensor assembled with Pd [56] J.-W. Yoon, J.-K. Choi, J.-H. Lee, Design of a highly sensitive and selective
nanoparticles-functionalized In2O3 composites, Sens. Actuators B: Chem. 285 C2H5OH sensor using p-type Co3O4 nanofibers, Sens. Actuators B: Chem. 161
(2019) 1–10. (2012) 570–577.
[29] L. Guo, C. Wang, X. Kou, N. Xie, F. Liu, H. Zhang, et al., Detection of triethylamine [57] D.-d Chen, Z. Li, X. Jin, J.-x Yi, Highly responsive and selective ethanol gas sensor
with fast response by Al2O3/α-Fe2O3 composite nanofibers, Sens. Actuators B: based on Co3O4-Modified SnO2 Nanofibers, Chin. J. Chem. Phys. 30 (2017)
Chem. 266 (2018) 139–148. 474–478.
[30] X. Song, Q. Xu, H. Xu, B. Cao, Highly sensitive gold-decorated zinc oxide [58] Y. Lu, Y. Ma, S. Ma, S. Yan, Hierarchical heterostructure of porous NiO
nanorods sensor for triethylamine working at near room temperature, J. Colloid nanosheets on flower-like ZnO assembled by hexagonal nanorods for high-
Interface Sci. 499 (2017) 67–75. performance gas sensor, Ceram. Int. 43 (2017) 7508–7515.
[31] S. Shi, F. Zhang, H. Lin, Q. Wang, E. Shi, F. Qu, Enhanced triethylamine-sensing [59] C. Liu, L. Zhao, B. Wang, P. Sun, Q. Wang, Y. Gao, et al., Acetone gas sensor based
properties of P-N heterojunction Co3O4/In2O3 hollow microtubes derived from on NiO/ZnO hollow spheres: fast response and recovery, and low (ppb) detection
metal–organic frameworks, Sens. Actuators B: Chem. 262 (2018) 739–749. limit, J. Colloid Interface Sci. 495 (2017) 207–215.
[32] L.-l Sui, Y.-M. Xu, X.-F. Zhang, X.-L. Cheng, S. Gao, H. Zhao, et al., Construction of [60] Z. Wen, L. Zhu, W. Mei, L. Hu, Y. Li, L. Sun, et al., Rhombus-shaped Co3O4
three-dimensional flower-like α-MoO3 with hierarchical structure for highly nanorod arrays for high-performance gas sensor, Sens. Actuators B: Chem. 186
selective triethylamine sensor, Sens. Actuators B: Chem. 208 (2015) 406–414. (2013) 172–179.
[33] H. Xu, W. Li, R. Han, T. Zhai, H. Yu, Z. Chen, et al., Enhanced triethylamine [61] S.F. Shen, M.L. Xu, D.B. Lin, H.B. Pan, The growth of urchin-like Co3O4 directly
sensing properties by fabricating Au@SnO2/α-Fe2O3 core-shell nanoneedles on sensor substrate and its gas sensing properties, Appl. Surf. Sci. 396 (2017)
directly on alumina tubes, Sens. Actuators B: Chem. 262 (2018) 70–78. 327–332.
[34] N. Luo, G. Sun, B. Zhang, Y. Li, H. Jin, L. Lin, et al., Improved TEA sensing [62] C. Sun, X. Su, F. Xiao, C. Niu, J. Wang, Synthesis of nearly monodisperse Co3O4
performance of ZnCo2O4 by structure evolution from porous nanorod to single- nanocubes via a microwave-assisted solvothermal process and their gas sensing
layer nanochain, Sens. Actuators B: Chem. 277 (2018) 544–554. properties, Sens. Actuators B: Chem. 157 (2011) 681–685.
[35] R. Malik, V.K. Tomer, N. Joshi, T. Dankwort, L. Lin, L. Kienle, Au–TiO2-loaded [63] K.-I. Choi, H.-R. Kim, K.-M. Kim, D. Liu, G. Cao, J.-H. Lee, C2H5OH sensing
cubic g-C3N4 nanohybrids for photocatalytic and volatile organic amine sensing characteristics of various Co3O4 nanostructures prepared by solvothermal
applications, ACS Appl. Mater. Interfaces 10 (2018) 34087–34097. reaction, Sens. Actuators B: Chem. 146 (2010) 183–189.
[36] X. Song, Q. Xu, T. Zhang, B. Song, C. Li, B. Cao, Room-temperature, high [64] U. Yaqoob, M.I. Younis, Chemical gas sensors: recent developments, challenges,
selectivity and low-ppm-level triethylamine sensor assembled with Au and the potential of machine learning—a review, Sensors 21 (2021) 2877.
decahedrons-decorated porous α-Fe2O3 nanorods directly grown on flat [65] J. Pan, R. Ganesan, H. Shen, S. Mathur, Plasma-modified SnO2 nanowires for
substrate, Sens. Actuators B: Chem. 268 (2018) 170–181. enhanced gas sensing, J. Phys. Chem. C. 114 (2010) 8245–8250.
[37] W. Yang, L. Feng, S. He, L. Liu, S. Liu, Density gradient strategy for preparation of [66] X. Kuang, T. Liu, D. Shi, W. Wang, M. Yang, S. Hussain, et al., Hydrothermal
broken In2O3 microtubes with remarkably selective detection of triethylamine synthesis of hierarchical SnO2 nanostructures made of superfine nanorods for
vapor, ACS Appl. Mater. Interfaces 10 (2018) 27131–27140. smart gas sensor, Appl. Surf. Sci. 364 (2016) 371–377.
[38] Y. Xu, T. Ma, L. Zheng, L. Sun, X. Liu, Y. Zhao, et al., Rational design of Au/ [67] Y.G. Song, Y.-S. Shim, J.M. Suh, M.-S. Noh, G.S. Kim, K.S. Choi, et al., Ionic-
Co3O4-functionalized W18O49 hollow heterostructures with high sensitivity and activated chemiresistive gas sensors for room-temperature operation, Small 15
ultralow limit for triethylamine detection, Sens. Actuators B: Chem. 284 (2019) (2019), 1902065.
202–212. [68] Z. Chen, C. Lu, Humidity sensors: a review of materials and mechanisms, Sens.
[39] V.V. Sysoev, B.K. Button, K. Wepsiec, S. Dmitriev, A. Kolmakov, Toward the Lett. 3 (2005) 274–295.
nanoscopic “electronic nose”: hydrogen vs carbon monoxide discrimination with [69] H. Farahani, R. Wagiran, M.N. Hamidon, Humidity sensors principle, mechanism,
an array of individual metal oxide nano- and mesowire sensors, Nano Lett. 6 and fabrication technologies: a comprehensive review, Sensors 14 (2014)
(2006) 1584–1588. 7881–7939.
[40] K. Suematsu, W. Harano, T. Oyama, Y. Shin, K. Watanabe, K. Shimanoe, Pulse- [70] C.-H. Kwak, T.-H. Kim, S.-Y. Jeong, J.-W. Yoon, J.-S. Kim, J.-H. Lee, Humidity-
driven semiconductor gas sensors toward ppt level toluene detection, Anal. Chem. independent oxide semiconductor chemiresistors using terbium-doped SnO2
90 (2018) 11219–11223. yolk–shell spheres for real-time breath analysis, ACS Appl. Mater. Interfaces 10
[41] K. Suematsu, Y. Shin, Z. Hua, K. Yoshida, M. Yuasa, T. Kida, et al., Nanoparticle (2018) 18886–18894.
cluster gas sensor: controlled clustering of SnO2 nanoparticles for highly sensitive [71] L. D’Arsié, V. Alijani, S.T.S. Brunelli, F. Rigoni, G. Di Santo, M. Caputo, et al.,
toluene detection, ACS Appl. Mater. Interfaces 6 (2014) 5319–5326. Improved recovery time and sensitivity to H2 and NH3 at room temperature with
[42] K. Suematsu, W. Harano, S. Yamasaki, K. Watanabe, K. Shimanoe, One-trillionth SnOx vertical nanopillars on ITO, Sci. Rep. 8 (2018) 10028.
level toluene detection using a dual-designed semiconductor gas sensor: material [72] L. D‘Arsié, M. Fanetti, C. Cepek, L. Casalis, P. Parisse, L. Gregoratti, et al., Tubular
and sensor-driven designs, ACS Appl. Electron. Mater. 2 (2020) 4122–4126. Sn-filled carbon nanostructures on ITO: nanocomposite material for multiple
[43] D.R. Miller, R.E. Williams, S.A. Akbar, P.A. Morris, D.W. McComb, STEM- applications, Carbon 65 (2013) 13–19.
cathodoluminescence of SnO2 nanowires and powders, Sens. Actuators B: Chem. [73] E. Leblanc, L. Perier-Camby, G. Thomas, R. Gibert, M. Primet, P. Gelin, NOx
240 (2017) 193–203. adsorption onto dehydroxylated or hydroxylated tin dioxide surface. Application
[44] D.R. Miller, R.E. Williams, S.A. Akbar, P.A. Morris, D.W. McComb, Measuring to SnO2-based sensors, Sens. Actuators B: Chem. 62 (2000) 67–72.
optical properties of individual SnO2 nanowires via valence electron energy-loss [74] G. Sberveglieri, Gas Sensors, Kluwer Academic Publishing, Berlin, Germany,
spectroscopy, J. Mater. Res. 32 (2017) 2479–2486. 1992.
[45] D.R. Miller, S.A. Akbar, P.A. Morris, Synthesis of hierarchical SnO2 [75] Figaro Engineering Inc, TGS 2602, 〈https://www.figaro.co.jp/en/product/entry/
nanowire–TiO2 nanorod brushes anchored to commercially available FTO-coated tgs2602.html〉.
glass substrates, Nano-Micro Lett. 9 (2017) 33. [76] T. Liu, X. Zhang, J. Zhang, W. Wang, L. Feng, L. Wu, et al., Interface study of ITO/
[46] M.M. Arafat, A.S.M.A. Haseeb, S.A. Akbar, Growth and characterization of the ZnO and ITO/SnO2 complex transparent conductive layers and their effect on
oxide scales and core/shell nanowires on Ti-6Al-4V particles during thermal CdTe solar cells, Int. J. Photo 2013 (2013), 765938.
oxidation, Ceram. Int. 41 (2015) 4401–4409. [77] K. Kim, J.K. Park, J. Lee, Y.J. Kwon, H. Choi, S.-M. Yang, et al., Synergistic
[47] M.M. Arafat, A.S.M.A. Haseeb, S.A. Akbar, M.Z. Quadir, In-situ fabricated gas approach to simultaneously improve response and humidity-independence of
sensors based on one dimensional core-shell TiO2-Al2O3 nanostructures, Sens. metal-oxide gas sensors, J. Hazard. Mater. 424 (2022), 127524.
Actuators B: Chem. 238 (2017) 972–984. [78] R. Lontio Fomekong, B. Saruhan, Influence of humidity on NO2-Sensing and
[48] D. Miller, S. Akbar, P. Morris, R.E.A. Williams, D.W. McComb, Correlative STEM- Selectivity of Spray-CVD Grown ZnO thin film above 400 ◦ C, Chemosensors 7
Cathodoluminescence and Low-Loss EELS of Semiconducting Oxide Nano- (2019) 42.
Heterostructures for Resistive Gas-Sensing Applications, Microsc. Microanal. 21
(2015) 1255–1256.

25
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

[79] W. Yan, M.A. Worsley, T. Pham, A. Zettl, C. Carraro, R. Maboudian, Effects of [106] H. Tang, C. Gao, H. Yang, L.N. Sacco, R. Sokolovskij, H. Zheng, et al., Room
ambient humidity and temperature on the NO2 sensing characteristics of WS2/ temperature ppt-level NO2 gas sensor based on SnOx/SnS nanostructures with
graphene aerogel, Appl. Surf. Sci. 450 (2018) 372–379. rich oxygen vacancies, 2D Mater. 8 (2021), 045006.
[80] Z. Ling, C. Leach, The effect of relative humidity on the NO2 sensitivity of a [107] J.R. Brent, D.J. Lewis, T. Lorenz, E.A. Lewis, N. Savjani, S.J. Haigh, et al., Tin(II)
SnO2/WO3 heterojunction gas sensor, Sens. Actuators B: Chem. 102 (2004) sulfide (SnS) nanosheets by liquid-phase exfoliation of herzenbergite: IV–VI main
102–106. group two-dimensional atomic crystals, J. Am. Chem. Soc. 137 (2015)
[81] Y. Wu, M. Yan, C. Tian, Y. Liu, Z. Hua, NO2 Sensing Properties of Cr2WO6 Gas 12689–12696.
Sensor in Air and N2 Atmospheres, Front. Chem. 7 (2020). [108] E. Bernal, Limit of detection and limit of quantification determination in gas
[82] N.M. Hung, N.M. Hieu, N.D. Chinh, T.T. Hien, N.D. Quang, S. Majumder, et al., chromatography, Adv. Gas. Chromatogr. 3 (2014) 57–63.
Rb2CO3-decorated In2O3 nanoparticles for the room-temperature detection of [109] M. Li, D. Zhou, J. Zhao, Z. Zheng, J. He, L. Hu, et al., Resistive gas sensors based
sub-ppm level NO2, Sens. Actuators B: Chem. 313 (2020), 128001. on colloidal quantum dot (CQD) solids for hydrogen sulfide detection, Sens.
[83] Y. Wang, L. Yao, L. Xu, W. Wu, W. Lin, C. Zheng, et al., Enhanced NO2 gas sensing Actuators B: Chem. 217 (2015) 198–201.
properties based on Rb-doped hierarchical flower-like In2O3 microspheres at low [110] L. Parellada-Monreal, et al., WO3 processed by direct laser interference
temperature, Sens. Actuators B: Chem. 332 (2021), 129497. patterning for NO2 detection, Sens Actuators B 305 (2020).
[84] T. Van Dang, N. Duc Hoa, N. Van Duy, N. Van Hieu, Chlorine gas sensing [111] Y. Wang, D. Liu, J. Yin, Y. Shang, J. Du, Z. Kang, et al., An ultrafast responsive
performance of on-chip grown zno, wo3, and sno2 nanowire sensors, ACS Appl. NO2 gas sensor based on a hydrogen-bonded organic framework material, Chem.
Mater. Interfaces 8 (2016) 4828–4837. Commun. 56 (2020) 703–706.
[85] J. Zhang, X. Jia, D. Lian, J. Yang, S. Wang, Y. Li, et al., Enhanced selective acetone [112] P. Qi, O. Vermesh, M. Grecu, A. Javey, Q. Wang, H. Dai, et al., Toward large
gas sensing performance by fabricating ZnSnO3/SnO2 concave microcube, Appl. arrays of multiplex functionalized carbon nanotube sensors for highly sensitive
Surf. Sci. 542 (2021), 148555. and selective molecular detection, Nano Lett. 3 (2003) 347–351.
[86] X. Zheng, H. Fan, H. Wang, B. Yan, J. Ma, W. Wang, et al., ZnO–SnO2 nano- [113] X. Wang, N. Aroonyadet, Y. Zhang, M. Mecklenburg, X. Fang, H. Chen, et al.,
heterostructures with high-energy facets for high selective and sensitive chlorine Aligned epitaxial SnO2 nanowires on sapphire: growth and device applications,
gas sensor, Ceram. Int. 46 (2020) 27499–27507. Nano Lett. 14 (2014) 3014–3022.
[87] L. Liu, Y. Wang, Y. Dai, G. Li, S. Wang, T. Li, et al., In situ growth of NiO@SnO2 [114] L. Li, C. Zhang, W. Chen, Fabrication of SnO2–SnO nanocomposites with p–n
hierarchical nanostructures for high performance H2S sensing, ACS Appl. Mater. heterojunctions for the low-temperature sensing of NO2 gas, Nanoscale 7 (2015)
Interfaces 11 (2019) 44829–44836. 12133–12142.
[88] T. Nguyen, N. Adjeroud, M. Guennou, J. Guillot, Y. Fleming, A.-M. Papon, et al., [115] J.A. Park, J. Moon, S.J. Lee, S.H. Kim, H.Y. Chu, T. Zyung, SnO2–ZnO hybrid
Controlling electrical and optical properties of zinc oxide thin films grown by nanofibers-based highly sensitive nitrogen dioxides sensor, Sens Actuators B 145
thermal atomic layer deposition with oxygen gas, Results Mater. 6 (2020), (2010) 592–595.
100088. [116] V. van Quang, N. van Dung, N. Sy Trong, N. Duc Hoa, N. van Duy, N. van Hieu,
[89] J.-H. Kim, J.-Y. Kim, J.-H. Lee, A. Mirzaei, H.W. Kim, S. Hishita, et al., Indium- Outstanding gas-sensing performance of graphene/SnO2 nanowire Schottky
implantation-induced enhancement of gas sensing behaviors of SnO2 nanowires junctions, Appl. Phys. Lett. 105 (2014).
by the formation of homo-core–shell structure, Sens. Actuators B: Chem. 321 [117] S. Cui, Z. Wen, X. Huang, J. Chang, J. Chen, Stabilizing MoS2 nanosheets through
(2020), 128475. SnO2 nanocrystal decoration for high-performance gas sensing in air, Small 11
[90] N. Van Hieu, H. Van Vuong, N. Van Duy, N.D. Hoa, A morphological control of (2015) 2305–2313.
tungsten oxide nanowires by thermal evaporation method for sub-ppm NO2 gas [118] J. Hao, D. Zhang, Q. Sun, S. Zheng, J. Sun, Y. Wang, Hierarchical SnS2/SnO2
sensor application, Sens. Actuators B: Chem. 171–172 (2012) 760–768. nanoheterojunctions with increased active-sites and charge transfer for
[91] Z. Khatoon, H. Fouad, H.K. Seo, M. Hashem, Z.A. Ansari, S.G. Ansari, Feasibility ultrasensitive NO2 detection, Nanoscale 10 (2018) 7210–7217.
study of doped SnO2 nanomaterial for electronic nose towards sensing [119] J.Z. Ou, W. Ge, B. Carey, T. Daeneke, A. Rotbart, W. Shan, et al., Physisorption-
biomarkers of lung cancer, J. Mater. Sci.: Mater. Electron. 31 (2020) based charge transfer in two-dimensional SnS2 for selective and reversible NO2
15751–15763. gas sensing, ACS Nano 9 (2015) 10313–10323.
[92] G. Zhang, H. Zeng, J. Liu, K. Nagashima, T. Takahashi, T. Hosomi, et al., [120] J. Wang, G. Lian, Z. Xu, C. Fu, Z. Lin, L. Li, et al., Growth of large-size SnS thin
Nanowire-based sensor electronics for chemical and biological applications, crystals driven by oriented attachment and applications to gas sensors and
Analyst 146 (2021) 6684–6725. photodetectors, ACS Appl. Mater. Interfaces 8 (2016) 9545–9551.
[93] M. Sun, N. Liu, H. Yu, X.-T. Dong, Y. Yang, L. Xia, Fabrication of porous WO3/ [121] Q. Sun, J. Wang, J. Hao, S. Zheng, P. Wan, T. Wang, et al., SnS2/SnS p–n
SBA-15 composite NOx gas sensor with a high sensitivity, J. Porous Mater. 28 heterojunctions with an accumulation layer for ultrasensitive room-temperature
(2021) 1031–1039. NO2 detection, Nanoscale 11 (2019) 13741–13749.
[94] K. Yuan, C.-Y. Wang, L.-Y. Zhu, Q. Cao, J.-H. Yang, X.-X. Li, et al., Fabrication of a [122] Y. Masuda, Ceramic nanostructures of SnO2, TiO2, and ZnO via aqueous crystal
micro-electromechanical system-based acetone gas sensor using CeO2 nanodot- growth: cold crystallization and morphology control, J. Ceram. Soc. Jpn. 128
decorated WO3 nanowires, ACS Appl. Mater. Interfaces 12 (2020) 14095–14104. (2020) 718–737.
[95] H. Zhang, Z. Jin, M.D. Xu, Y. Zhang, J. Huang, H. Cheng, et al., Enhanced [123] Y. Masuda, K. Kato, Aqueous synthesis of nano-sheet assembled tin oxide particles
isopropanol sensing performance of the CdS nanoparticle decorated ZnO Porous and Their N2 adsorption characteristics, J. Cryst. Growth 311 (2009) 593–596.
nanosheets-based gas sensors, IEEE Sens. J. 21 (2021) 13041–13047. [124] Y. Masuda, T. Ohji, K. Kato, Room-temperature synthesis of tin oxide nano-
[96] J. Chu, X. Yang, A. Yang, D. Wang, H. Yuan, X. Wang, et al., Multivariate electrodes in aqueous solutions, Thin Solid Films 518 (2009) 850–852.
evaluation method for screening optimum gas-sensitive materials for detecting [125] K. Momma, F. Izumi, VESTA 3 for three-dimensional visualization of crystal,
SF6 decomposition products, ACS Sens. 5 (2020) 2025–2035. volumetric and morphology data, J. Appl. Crystallogr. 44 (2011) 1272–1276.
[97] Y.-X. Li, Z. Guo, Y. Su, X.-B. Jin, X.-H. Tang, J.-R. Huang, et al., Hierarchical [126] W.H. Baur, A.A. Khan, Rutile-type compounds. IV. SiO2, GeO2 and a comparison
morphology-dependent gas-sensing performances of three-dimensional SnO2 with other rutile-type structures, Acta Crystallogr. Sect. B 27 (1971) 2133–2139.
nanostructures, ACS Sens. 2 (2017) 102–110. [127] Y. Masuda, T. Ohji, K. Kato, Tin oxide nanosheet assembly for hydrophobic/
[98] N.H. Ly, H.H. Kim, S.-W. Joo, On-site detection for hazardous materials in hydrophilic coating and cancer sensing, ACS Appl. Mater. Interfaces 4 (2012)
chemical accidents, Bull. Korean Chem. Soc. 42 (2021) 4–16. 1666–1674.
[99] W.S. Chi, C.S. Lee, H. Long, M.H. Oh, A. Zettl, C. Carraro, et al., Direct [128] H. Zakaryan, V. Aroutiounian, C.O. Gas, Adsorption on SnO 2 surfaces: density
organization of morphology-controllable mesoporous SnO2 using amphiphilic functional theory study, Sens. Transducers 212 (2017) 50–56.
graft copolymer for gas-sensing applications, ACS Appl. Mater. Interfaces 9 [129] C. Jiang, G. Zhang, Y. Wu, L. Li, K. Shi, Facile synthesis of SnO2 nanocrystalline
(2017) 37246–37253. tubes by electrospinning and their fast response and high sensitivity to NOx at
[100] T. Zhou, T. Zhang, R. Zhang, Z. Lou, J. Deng, L. Wang, Hollow ZnSnO3 cubes with room temperature, CrystEngComm 14 (2012) 2739–2747.
controllable shells enabling highly efficient chemical sensing detection of [130] A.A. Abokifa, K. Haddad, J. Fortner, C.S. Lo, P. Biswas, Sensing mechanism of
formaldehyde vapors, ACS Appl. Mater. Interfaces 9 (2017) 14525–14533. ethanol and acetone at room temperature by SnO2 nano-columns synthesized by
[101] R.B. Shinde, A.S. Patil, S.V. Sadavar, Y.M. Chitare, V.V. Magdum, N.S. Padalkar, aerosol routes: theoretical calculations compared to experimental results,
et al., Polyoxotungstate intercalated self-assembled nanohybrids of Zn-Cr-LDH for J. Mater. Chem. A 6 (2018) 2053–2066.
room temperature Cl2 sensing, Sens. Actuators B: Chem. 352 (2022), 131046. [131] H. Feng, C. Li, T. Li, F. Diao, T. Xin, B. Liu, et al., Three-dimensional hierarchical
[102] T.K. Dang, N. Van Toan, C.M. Hung, N. Van Duy, N.N. Viet, L.V. Thong, et al., SnO2 dodecahedral nanocrystals with enhanced humidity sensing properties,
Investigation of zinc electronucleation and growth mechanisms onto platinum Sens. Actuators B: Chem. 243 (2017) 704–714.
electrode from a deep eutectic solvent for gas sensing applications, J. Appl. [132] Y. Masuda, T. Itoh, W. Shin, K. Kato, SnO2 nanosheet/nanoparticle detector for
Electrochem. (2021). the sensing of 1-nonanal gas produced by lung cancer, Sci. Rep. 5 (2015) 10122.
[103] S. Yan, J. Zhang, Y. Huang, Y. Lu, Preparation of coral-like SnO2 hierarchical [133] Y. Masuda, K. Kato, M. Kida, J. Otsuka, Selective nonanal molecular recognition
nanostructures and its application in ethanol gas-sensing performance, IOP Conf. with SnO2 nanosheets for lung cancer sensor, Int J. Appl. Ceram. Technol. 16
Ser.: Mater. Sci. Eng. 611 (2019), 012039. (2019) 1807–1811.
[104] E. Boateng, S.S. Thind, S. Chen, A. Chen, Synthesis and electrochemical studies of [134] P.G. Choi, N. Izu, N. Shirahata, Y. Masuda, Improvement of sensing properties for
WO3-based nanomaterials for environmental, energy and gas sensing SnO2 gas sensor by tuning of exposed crystal face, Sens. Actuators B: Chem. 296
applications, Electrochemical Science Advances, n/a e2100146. (2019), 126655.
[105] M.-W. Ahn, K.-S. Park, J.-H. Heo, J.-G. Park, D.-W. Kim, K.J. Choi, et al., Gas [135] K. Kim, Pg Choi, T. Itoh, Y. Masuda, Catalyst-free highly sensitive SnO2 nanosheet
sensing properties of defect-controlled ZnO-nanowire gas sensor, Appl. Phys. Lett. gas sensors for parts per billion-level detection of acetone, ACS Appl. Mater.
93 (2008), 263103. Interfaces 12 (2020) 51637–51644.

26
Y. Masuda Sensors and Actuators: B. Chemical 364 (2022) 131876

[136] L. Cheng, S.Y. Ma, T.T. Wang, J. Luo, Synthesis and enhanced acetone sensing [144] G. Ren, Z. Li, W. Yang, M. Faheem, J. Xing, X. Zou, et al., ZnO@ZIF-8 core-shell
properties of 3D porous flower-like SnO2 nanostructures, Mater. Lett. 143 (2015) microspheres for improved ethanol gas sensing, Sens. Actuators B: Chem. 284
84. (2019) 421–427.
[137] J. Li, P. Tang, J. Zhang, Y. Feng, R. Luo, A. Chen, et al., Facile synthesis and [145] X. Wu, S. Xiong, Z. Mao, S. Hu, X. Long, A Designed ZnO@ZIF-8 core–shell
acetone sensing performance of hierarchical SnO2 hollow microspheres with nanorod film as a gas sensor with excellent selectivity for H2 over CO, Chem. – A
controllable size and shell thickness, Ind. Eng. Chem. Res 55 (2016) 3588–3595. Eur. J. 23 (2017) 7969–7975.
[138] X. Wang, F. Liu, X. Chen, G. Lu, X. Song, J. Tian, et al., SnO2 core-shell hollow [146] Y. Liu, R. Wang, T. Zhang, S. Liu, T. Fei, Zeolitic imidazolate framework-8 (ZIF-8)-
microspheres Co-modification with Au and NiO nanoparticles for acetone gas coated In2O3 nanofibers as an efficient sensing material for ppb-level NO2
sensing, Powder Technol. 364 (2020) 159–166. detection, J. Colloid Interface Sci. 541 (2019) 249–257.
[139] Y. Xu, L. Zheng, C. Yang, X. Liu, J. Zhang, Highly sensitive and selective electronic [147] K.S. Park, Z. Ni, A.P. Côté, J.Y. Choi, R. Huang, F.J. Uribe-Romo, et al.,
sensor based on Co Catalyzed SnO2 nanospheres for acetone detection, Sens Exceptional chemical and thermal stability of zeolitic imidazolate frameworks,
Actuator B-Chem. 304 (2020), 127237. Proc. Natl. Acad. Sci. 103 (2006) 10186–10191.
[140] S. Zhu, D. Zhang, J. Gu, J. Xu, J. Dong, J. Li, Biotemplate fabrication of SnO2 [148] M. Drobek, J.-H. Kim, M. Bechelany, C. Vallicari, A. Julbe, S.S. Kim, MOF-based
nanotubular materials by a sonochemical method for gas sensors, J. Nanopart. membrane encapsulated ZnO nanowires for enhanced gas sensor Selectivity, ACS
Res 12 (2010) 1389. Appl. Mater. Interfaces 8 (2016) 8323–8328.
[141] L. Li, H. Lin, F. Qu, Synthesis of mesoporous SnO2 nanomaterials with selective [149] T. Zhou, Y. Sang, X. Wang, C. Wu, D. Zeng, C. Xie, Pore size dependent gas-
gas-sensing properties, J. Sol. -Gel Sci. Technol. 67 (2013) 545. sensing selectivity based on ZnO@ZIF nanorod arrays, Sens. Actuators B: Chem.
[142] P.G. Choi, N. Shirahata, Y. Masuda, Tin oxide nanosheet thin film with bridge 258 (2018) 1099–1106.
type structure for gas sensing, Thin Solid Films 698 (2020), 137845. [150] M. Weber, J.-H. Kim, J.-H. Lee, J.-Y. Kim, I. Iatsunskyi, E. Coy, et al., High-
[143] P.G. Choi, Z. Liu, N. Hara, Y. Masuda, Surface molecular separator for selective performance nanowire hydrogen sensors by exploiting the synergistic effect of Pd
gas sensing, Ind. Eng. Chem. Res. 59 (2020) 17894–17900. nanoparticles and metal–organic framework membranes, ACS Appl. Mater.
Interfaces 10 (2018) 34765–34773.

27

You might also like