Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Sensors & Actuators: A.

Physical 369 (2024) 115181

Contents lists available at ScienceDirect

Sensors and Actuators: A. Physical


journal homepage: www.journals.elsevier.com/sensors-and-actuators-a-physical

TiO2-ZnO Composite thin films fabricated by Sol-Gel spin coating for room
temperature impedometric acetone sensing
Muthukumar Murugesan , S.R. Meher *
Department of Physics, School of Advanced Sciences, Vellore Institute of Technology, Vellore 632014, India

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

Keywords: Metal oxide semiconductor based volatile organic compounds (VOCs) sensors lack room temperature operation
Metal oxides and have a poor limit of detection. In this work, TiO2-ZnO composite thin film based acetone sensors with room
Semiconductors temperature (29 ± 2 ◦ C) operation and good sensor response have been fabricated by sol-gel based spin coating.
Thin films
The thin films based on the single-phase TiO2 and ZnO together with their composites have been evaluated for
TiO2-ZnO
Sol-gel spin coating
their structural and surface morphological properties through X-ray diffraction and field emission scanning
Acetone sensing electron microscopy respectively. The surface chemical state of the thin films surface has been analyzed through
the X-ray photoelectron spectroscopy. The optical properties of the single-phase as well as the composite phase
thin films were studied by optical absorption spectroscopy. The acetone sensing characteristics of these thin films
based sensors were evaluated through impedance spectroscopy. The composite thin films were found to exhibit
the maximum sensor response of 93.83% with a detection threshold of 3 ppm for acetone.

1. Introduction at various concentrations which can be helpful towards the diagnosis of


diabetes.
Human exhaled breath contains many volatile organic compounds Among these MOS, titanium dioxide (TiO2) with a wide band gap of
(VOCs) owing to metabolism and the alteration of the same can be an 3.2 eV is particularly sensitive to acetone. The acetone sensing by TiO2 is
indication of various pathological disorders. Due to the rapid advance­ mostly attributed to the presence of native oxygen vacancies and anti-
ment of analytical techniques, the VOC concentration in exhaled breath site defects, which establish n-type conductivity in it [11]. The elec­
can act as a biomarker (BM) for several disease diagnoses. The acetone trical conductivity of TiO2 changes because of the change in its electron
concentration in healthy subjects ranges from 300–900 ppb whereas the concentration upon exposure to acetone. As per the literature, Navale
same rises to 1800 ppb (or 1.8 ppm) for patients suffering from diabetes et al. [12] have developed a TiO2 mesoporous sensor with an enhanced
mellitus [1]. Therefore, a compact acetone sensor with a low detection detection threshold of 500 ppb for acetone at an operating temperature
threshold that can operate under ambient conditions can be helpful for of 270 ◦ C. Meanwhile, the TiO2 nanotube-based sensor developed by
the early diagnosis and control of diabetes mellitus. Bindra et al. [13] could detect various VOCs, including acetone, at room
Chemiresistive sensors based on metal-oxide-semiconductors (MOS) temperature, although its detection threshold was greater than 100
offer trustworthy solutions to detect acetone in contrast to other sensing ppm. The major practical problems associated with the single-phase
techniques. Due to their exceptional chemical and physical features, TiO2 based acetone sensors is their high operating temperature or
easily tuneable physicochemical traits, straightforward sensing pro­ poor detection threshold at room temperature [14,15]. Numerous
cedures, ease of manufacture, and cost-effectiveness, MOSs are very studies have been conducted to modify the single-phase TiO2 in terms of
reliable for practical applications. Because of their excellent chemical doping [16], metal-insulator-semiconductor (MIS) device approach [17]
stability, quick reaction to target gas conditions, and other character­ and molecular imprinting [18] to overcome the above issues.
istics, practical obstacles during acetone sensing can be addressed [2]. Another promising approach that has received a lot of attention in
The development of MOS-based acetone sensors has been the subject of the recent past to enhance the VOC sensing characteristics is the com­
many published studies [3,4]. Materials like SnO2 [5], ZnO [6], TiO2 bination of two different MOS to form heterojunction composites. There
[7], In2O3 [8], Fe3O4 [9], and WO3 [10] are useful for detecting acetone are large number of active sites at the heterojunction interface due to the

* Corresponding author.
E-mail address: samirmeher@gmail.com (S.R. Meher).

https://doi.org/10.1016/j.sna.2024.115181
Received 10 November 2023; Received in revised form 21 January 2024; Accepted 12 February 2024
Available online 13 February 2024
0924-4247/© 2024 Elsevier B.V. All rights reserved.
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

lattice mismatch. Meanwhile, the difference in Fermi levels of the two Table 1
MOS in the composite would result in the establishment of a potential The layer sequence of spin-coated single-phase (TiO2 and ZnO) and composite
barrier at their interfaces which is a key factor for enhancing the sensing (TiO2 – ZnO) thin films.
property of the heterojunction-based sensors [19]. The composite S. No TiO2 ZnO TiO2 - ZnO
MOS-based acetone sensors are reported to be more sensitive than 1 T Z T
single-phase materials and have improved detection threshold [20]. The 2 T Z T
composites of TiO2 - ZnO [21], TiO2 - SnO2 [22], TiO2 – reduced Gra­ 3 T Z Z
phene Oxide (rGO) [20], and Graphene Oxide (GO)- TiO2 [23] are re­ 4 T Z Z
5 T Z Z
ported to have good response with better detection threshold at lower
6 T Z T
working temperatures. Among these heterojunctions, the composites of 7 T Z T
TiO2 with ZnO have got potential features for enhancing the acetone 8 T Z Z
sensing characteristics. This is due to the wide band gap (~3.3 eV) and 9 T Z Z
n-type conductivity of ZnO which is identical to that of TiO2. The n – n 10 T Z Z
11 T Z T
type TiO2 – ZnO heterojunction facilitates easy electron transfer be­ 12 T Z T
tween the semiconductors which in turn results in fast change in the
resistance upon exposure to the target gas. Feng et al. [24] have fabri­ Note: denote the TiO2 – T; denote the ZnO – Z.
cated acetone sensors based on the TiO2 @B-TiO2 @ZnO core-shell
heterojunction. They have reported an improved acetone detection were spin-coated on cleaned glass substrates at a spinning speed of 3000
threshold of 170 ppb at 275 ◦ C which primarily has been attributed to rpm for 30 s with a ramping up for 5 s. After coating each layer, the films
the narrower band gap (~1.54 eV) of black TiO2 (B-TiO2) due to Ti+3 were pre-heated at 250 ◦ C in air ambience for 3 min. For the deposition
self-doping in TiO2. Recently, Krishna et al. [25] developed a ZnO/TiO2 of TiO2 – ZnO composite thin films, the molar concentrations for Ti and
nanocomposites-based sensor that is capable of detecting acetone at Zn precursors were kept the same whereas the layer sequence was
room temperature with a response of 179% and a low detection limit of changed. The sequence for various layers of deposition to obtain single-
100 ppb, which emphasizes the potential of the sensor for diabetes ap­ phase TiO2, composite TiO2 – ZnO and single-phase ZnO is given in
plications due to its synergy. Table 1. Finally, all the films were subjected to post-deposition
The sensors based on nanostructured materials can significantly annealing of 500 ◦ C for 1 h in air ambience. Using spectroscopic
enhance the sensor response as well as reduce the operating tempera­ ellipsometry, the thickness of single-phase TiO2, ZnO and their com­
tures. But they suffer from several issues such as repeatability and reli­ posite thin films were determined to be 297.5 nm, 272.4 nm, and 219.6
ability for practical usage. The thin film-based sensors are more robust nm, respectively.
and are commercially viable [26]. As per the current status, there are no
reports on TiO2 thin film-based acetone sensors operating at room 2.2. Thin film characterization
temperature with an appreciable low-level detection threshold. In the
present investigation, we have prepared the thin films of single-phase The phase and crystalline properties of the deposited thin films were
TiO2, ZnO and their composites by sol-gel based spin coating. The studied by the PANalytical X′Pert Pro X-ray diffractometer using a Cu Kα
impedometric room temperature (29 ± 2 ◦ C) sensing characteristics of radiation of wavelength 1.540 Å. The X-ray tube was operated at 35 kV
the fabricated sensors based on these thin films were evaluated in the voltage and 30 mA current. The optical transmittance of the thin films
range 10 – 70 ppm. was measured using a JASCO V570 double-beam spectrophotometer in
the wavelength range of 300 – 2500 nm. Spectroscopic ellipsometry (J.
2. Experimental methods A. Wollam - Alpha SE) technique was used to determine the thickness of
the films. The surface morphology, grain size distribution and elemental
2.1. Thin film deposition analysis were studied using the FEI Quanta 250 FEG field emission
scanning electron microscopy (FESEM). Further, the surface topography
Sol-gel spin coating was used to coat the single-phase as well as the and roughness were examined using the Nanosurf Easyscan 2 atomic
composite thin films of TiO2 and ZnO. The films were deposited on force microscope (AFM) (Model: 23–06-154). The chemical composition
cleaned soda-lime glass substrates of size 25 mm × 25 mm. The glass and oxidation state of the sample surface were studied using the X-ray
substrates were heated for 5 min at a temperature of 400 ◦ C and sub­ photoelectron spectrometer (Model: PHI 5000 Versa Probe III) equipped
sequently cleaned with Cedepol soap solution to remove any kind of with Al Kα X-ray source of energy 1486.6 eV.
organic contaminants. After 15 min of ultrasonic treatment with
ethanol, the substrates were treated with deionized water. The purpose 2.3. Sensor fabrication and measurements
of this process was to clean the substrates for any kind of residual con­
taminants. Finally, the substrates were flushed with dry N2 before being To fabricate the acetone sensors, Ni electrodes were coated through
spin coated. Titanium (IV) isopropoxide (C12H28O4Ti) and zinc acetate thermal evaporation in an interdigitated configuration on the deposited
dihydrate (Zn(CH3COO)2.2 H2O) were the metalorganic precursors for single-phase and composite thin films by thermal evaporation. The work
TiO2 and ZnO respectively. The 0.1 M of precursor salt for TiO2 was function of Ni (5.01 eV) is ideal for forming Ohmic contacts with n-type
dissolved in ethanol (CH2CH3OH), which served as the solvent. Under wide band gap semiconductors such as TiO2 and ZnO. The typical di­
constant stirring, a few drops of water to initiate the hydrolysis process mensions for the interdigitated configurations are depicted in Fig. 1. The
and a few drops of concentrated nitric acid (HNO3) to prevent precipi­ Ni electrode was thermally evaporated from a basket-shaped W filament
tation were added to the solution. To acquire a transparent homoge­ at a deposition rate of 5 Å/s onto the masked surface of the thin films.
neous solution, the solution was stirred for 5 h. The obtained solution The deposition rate was monitored in-situ using a quartz crystal monitor
was aged for 24 h before being spin coated. Similarly, for ZnO the Zn and the final thickness of the electrodes was ~200 nm. After the depo­
precursor with a molar concentration of 0.3 M was added to 2-methox­ sition of the electrodes, two-point electrical contacts were made using
yethanol (CH3OCH2CH2OH) solvent. An equimolar concentration of Cu wires with the help of Ag paste on top of the electrodes. Finally, the
monoethanolamine (NH2CH2CH2OH) was used as the sol-stabilizer. The sensors were dried using a hotplate for 10 min at 200 ◦ C. To study the
resultant solution was stirred at room temperature for 5 h to obtain a acetone sensing performance, we fabricated the single-phase and com­
transparent homogeneous solution. The sol was kept for 48 h before posite sensors for various acetone concentrations at room temperature
being used for spin coating. After the gelation, 12 layers of the solution (29 ± 2 ◦ C). Gas sensing measurements in this study were conducted

2
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

Fig. 1. (Colour online) Schematic of (a) Cross-sectional view of the proposed sensors (b) Top view of the electrode geometry.

impedance measurements with respect to different frequencies at


varying acetone concentrations, the sensors were initially placed in the
sensing chamber under ambient conditions. Then, the chamber was
sealed and filled with N2 gas at a flow rate of 10 sccm through a mass
flow controller. The primary purpose of N2 gas is to maintain a constant
baseline impedance for all sensors and to minimize interference from
other gases inside the sensing chamber. After 1 min, the flow of N2 gas
was stopped with the help of a gate valve. At that time, the N2 atmo­
spheric baseline sensor impedance was recorded for reference. During
acetone sensing, acetone vapor was collected from the bubbler system
and injected into the sensing chamber, where the concentration was
controlled using a needle valve. In parallel, the concentration was
monitored with the help of a VOC detector. For each concentration, once
the steady state was reached, the sensor impedance was continuously
monitored with respect to different frequencies. After the measurement
was completed, the sensing chamber was evacuated using a vacuum
pump. For the sensor response time measurement, initially, the sensor
impedance was recorded at a constant frequency of 100 Hz in the N2
atmosphere. The sensor impedance was then monitored for impedance
variations after 70 ppm of acetone was added to the chamber. For re­
covery time measurements, the sensing chamber was opened to expose it
to air, and then the sensors started to retain their original state with their
corresponding impedance values at room temperature.

3. Results and discussion

3.1. Structural, morphological and compositional studies of thin films


Fig. 2. (Colour online) XRD patterns for the single-phase (TiO2 and ZnO) and
composite (TiO2 – ZnO) thin films. The X-ray diffraction (XRD) patterns for the single-phase TiO2 and
ZnO thin films together with the TiO2 – ZnO mixed layer thin films are
using a homemade stainless steel chamber and commercially available shown in Fig. 2. All the deposited films are found to exhibit the poly­
high-purity VOCs. The sensing setup consists of a stainless steel chamber crystalline nature. The Bragg peaks for single-phase TiO2 thin films are
(350 mL) interconnected with a VOC concentration detector (Tiger XT observed at 25.28◦ , 37.8◦ , 48.04◦ , 53.8◦ and 55.06◦ corresponding to the
handheld detector), a conical flask with a bubbler system, and a PC (101), (004), (200), (105) and (211) lattice planes respectively (Fig. 2a).
interfaced with (Hioki: LCR HiTESTER 3532–50) LCR meter. By inter­ This confirms the tetragonal crystal structure of TiO2 in the anatase
facing the measuring instrument to the computer system through the phase (JCPDS: 21–1272). The growth is along the (101) lattice plane.
National Instruments software LabVIEW, the impedance data were ac­ The single-phase ZnO thin films crystallize in the hexagonal wurtzite
quired for the proposed sensors. The impedance (Z) and the phase angle structure (Fig. 2b). The corresponding lattice planes of (100), (002),
(Φ) were recorded in the frequency range of 100 Hz to 5 MHz. For (101) and (110) are confirmed from the respective diffraction peaks
centered at 31.7◦ , 34.4◦ , 36.2◦ and 56.4◦ (JCPDS: 36–1451). The pref­
Table 2 erential growth is along the c-axis [27]. On the other hand, for the mixed
Preferential orientation, FWHM and average crystallite size obtained from XRD layer films, the diffraction peaks corresponding to both anatase TiO2 and
for single-phase (TiO2 and ZnO) and composite (TiO2 – ZnO) thin films. wurtzite ZnO are observed which confirms their composite nature
Thin film Plane Peak FWHM Crystallite Size (Fig. 2c). For these composite films, the preferential growth for the ZnO
Orientation Position (nm) crystallites is along the a-axis whereas the crystallites corresponding to
(degree) (002), (101) and (110) orientations are absent. All the diffraction peaks
TiO2 (101) 25.28 0.2079 39.12 for anatase TiO2 are observed with a reduction in the intensity. The
ZnO (002) 34.45 0.1662 50.05 average crystallite size (D) was calculated by using the Debye-Scherrer
TiO2 – ZnO TiO2 – (101) 25.28 0.3116 26.09 formula given as:
ZnO – (100) 31.86 0.2818 29.24

3
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

Fig. 3. (Colour online) High resolution XPS spectra for (a) Ti-2p (b) O-1s orbitals for the single-phase TiO2 thin films.

Fig. 4. (Colour online) High resolution XPS spectra for (a) Ti-2p (b) Zn-2p (c) O-1s orbitals for the composite thin films (d) Single and composite phase comparison of
Ti-2p orbitals.

4
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

O 1 s peaks through Voigt profile with Shirley background correction.


The peak-to-peak distance and peak area ratio constraints were applied
during the deconvolution of peaks [28]. The high-resolution XPS spectra
for Ti 2p and O 1 s orbitals for the single-phase TiO2 thin films are shown
in Fig. 3. The signature peaks corresponding to Ti 2p3/2 and 2p1/2 are
observed at 458.51 eV and 464.25 eV respectively [29]. The peak po­
sitions and the spin-orbit energy difference of 5.74 eV confirms the + 4
oxidation state of Ti [30]. The high-resolution O 1 s spectra shows the
high-intensity peak centered at 529.80 eV corresponding to the lattice
oxygen (OL). The low-intensity peak at 531.34 eV is associated with the
O- adsorbed species at the oxygen vacancy (OV) sites [31,32]. The peaks
corresponding to Ti 2p3/2 and Ti 2p1/2 for the composite thin film are
observed at 458.45 eV and 464.22 eV, respectively (Fig. 4a). For the
composite thin films, the peaks corresponding to Ti 2p3/2 and 2p1/2
states shift slightly towards lower binding energy as compared to the
single-phase TiO2 (Fig. 4a). The Zn2+ chemical state is confirmed from
the XPS peaks located at 1021.71 eV and 1044.83 eV corresponding to
2p3/2 and 2p1/2 states [33] (Fig. 4b). The peak area ratio is subjected to
constraints specified by the (2J +1) condition. This condition is based on
the total angular momentum quantum number (J), which is equal to the
Fig. 5. (Colour online) Valence band (VB) spectra for Single-phase TiO2 sum of orbital angular momentum (l), and the spin quantum number
thin film. (s = ±12). Accordingly, the peak area ratio of (Ti or Zn 2p1/2) and (Ti or
Zn 2p3/2) is 12 [34,35]. The increase in the relative intensity of the peak
0.9λ
D= (1) corresponding to OV signifies the formation of a greater number of ox­
β cos θ ygen vacancy sites in the composite thin films (Fig. 4c). For the com­
Here, λ is the wavelength of the X-ray, θ is the peak center and β is the posite thin films, the peaks corresponding to Ti 2p3/2 and 2p1/2 states
full width at half maximum (FWHM) corrected by the instrumental shift slightly towards lower binding energy as compared to the
broadening parameter determined from the monocrystalline Si diffrac­ single-phase TiO2 shown in inset of (Fig. 4d). The partial electron
tion line. The spherical crystallites are assumed for all the films. The transfer from ZnO to TiO2 causes the shift of the Ti 2p states towards
FWHM and the corresponding average crystallite size for all the films are lower binding energy due to increase in the electron concentration [36].
given in Table 2. A decrease in the crystallite size is observed for the The absence of any other chemical oxidation states of Zn or Ti suggests
composite films compared to the single-phase films. that no other ternary phases were formed during the deposition.
The chemical state and elemental composition of Ti, Zn and O atoms The valence band maximum (VBM) relative to Fermi energy level
at the film surface were characterized by X-ray photoelectron spec­ was determined from the valence band spectra through a linear
troscopy (XPS). Being a surface-sensitive technique with a penetration extrapolation of its lower energy region. The same has been depicted for
depth limit of ~10 nm, all the films were sputter etched for 2 min using single phase TiO2 thin films in Fig. 5. The valence band spectra between
Ar ions to remove the chemically adsorbed impurity atoms from the 1.0 to 9.0 eV contains two peaks centered at 3.02 eV and 5.0 eV for the π
surface. The XPS spectra were calibrated with respect to the adventitious (non-bonding) and σ (bonding) O 2p orbitals, respectively. The corre­
C 1 s peak corresponding to 284.6 eV. XPS PEAKFIT software was used sponding VBM is found to be 1.28 eV from the linear extrapolation. The
to deconvolute the high-resolution core-level spectra of Ti 2p, Zn 2p, and intensity of the σ orbital is higher than that of the π orbital which is

Fig. 6. (Colour online) Planar view FESEM micrograph of (a) TiO2 (b) ZnO, and (c) TiO2 - ZnO thin films.

5
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

Fig. 7. (Colour online) AFM images (2 µm × 2 µm) of (a) TiO2 (b) ZnO (c) TiO2 - ZnO thin films with roughness.

composite thin films. The AFM micrographs with a scan area of 2 µm


× 2 µm are shown in Fig. 7. The root mean square (RMS) roughness for
single-phase TiO2 films is found to be 12.16 nm with a random distri­
bution of the grains. This is in accordance with the FESEM results. The
ZnO thin films are found to be much smoother with an RMS roughness of
5.99 nm. The composite films exhibit a slightly higher RMS roughness of
23.16 nm. Both ZnO and the composite films are found to have hillock
shape domains. According to these results, the composite thin films have
distinct features with rough surface which can be attributed to the lattice
mismatch and the internal strain developed in them. Moreover, higher
surface roughness results in an increase in the number of surface reactive
sites which plays an essential role in the sensing of analytes [37].

3.2. Optical properties

The transmittance spectra for the deposited thin films in the wave­
length region 300 – 2500 nm are shown in Fig. 8. The TiO2 thin films are
found to have higher transmittance (> 80%) than the ZnO thin films in
the visible region. In the near-infrared (NIR) region, both the films
exhibit ~80 – 85% transparency. The oscillations corresponding to the
Fig. 8. (Colour online) UV–Vis-NIR transmittance spectra for single-phase and characteristic thin film interference indicates good optical quality of the
composite films. The Inset image shows the absorption edge. deposited thin films with a thickness of few 100 s of nm. The ultraviolet
(UV) absorption corresponding to the band edge is well-defined for the
consistent with the literature [28]. For the composite thin films in the single-phase films. In contrast, a broad absorption edge is observed for
absence of any well-defined VBM, it was not possible to accurately the composite thin films signifying the presence of band tail states. The
determine the same from the valence band spectroscopy. band gap corresponding to the single-phase TiO2 and ZnO were deter­
The FESEM micrographs for the single-phase TiO2 and ZnO thin films mined through Tauc’s plot method Fig. 9. The band gap type (direct/
and the TiO2 – ZnO composite thin films are shown in Fig. 6. The TiO2 indirect) of TiO2 is debatable in the literature. But majority of the
surface is free of any cracks with randomly distributed spherical grains experimental reports suggest an indirect band gap for anatase. In the
(Fig. 6a). The typical grain size ranges from 25 – 35 nm. Fig. 6b depicts present study also, we have assumed the indirect band transition in TiO2
the morphology of single-phase ZnO thin films, which reveals that the and have calculated the band gap value of 3.10 eV. On the other hand,
irregularly connected grains are distributed across the surface with ZnO is a well-known direct band gap semiconductor and the calculated
random agglomerated particles on top of the grains. In the case of band gap value is 3.22 eV. Because of the presence of broad absorption
composite TiO2 – ZnO thin films, we observe an uneven distribution of edge, the parabolic band approximation is invalid for the composite thin
TiO2 and ZnO grains like a plum-pudding type morphology (Fig. 6c). The films. Therefore, the corresponding band gap cannot be determined
ZnO grains appear to be diffused into the TiO2 layer because of the post- from the optical absorption.
deposition annealing.
Atomic force microscopy (AFM) in contact mode was used to 3.3. Acetone sensing performance
examine the morphology and surface roughness of the single-phase and
The room temperature (29 ± 2 ◦ C) acetone sensing performance was

6
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

Fig. 9. (Colour online) The plot of (a) (αhν)1/2 versus hν for indirect bandgap for TiO2 (b) (αhν)2 versus hν for direct bandgap for ZnO thin films.

Fig. 10. (Colour online) Impedance vs acetone concentration of proposed Fig. 11. (Colour online) Sensor response of single-phase (TiO2 or ZnO) and
sensors at 100 Hz. The inset image shows the single-phase sensors impedance composite (TiO2-ZnO) sensors at 100 Hz.
changes in the kΩ range.
157.7 kΩ is observed for the composite thin films when exposed to
examined for the single-phase TiO2 and ZnO thin film sensors and TiO2 – 70 ppm acetone. Together with this an improvement in the detection
ZnO composite thin film sensors by observing its change in the imped­ threshold from 10 ppm to 3 ppm is observed for the composite sensors
ance with respect to the N2 baseline for various acetone concentrations. compared to the single-phase ones. Acetone sensors with extreme
Fig. 10 shows the variation in the impedance at 100 Hz for different detection threshold of 20 ppb and 170 ppb have been reported in the
sensors exposed to acetone concentrations ranging from 10 – 70 ppm. literature [22,24]. But these sensors require higher temperatures to
The lower frequency of 100 Hz has been chosen to ensure space charge achieve this detection threshold. A potential acetone sensor for the early
redistribution [38]. Upon exposure to acetone, we observe a significant diagnosis of diabetes requires lower detection threshold together with
change in the impedance for the TiO2 – ZnO composite sensors room temperature operation [25]. In this context, the detection
compared to the single-phase TiO2 and ZnO based sensors. The com­ threshold of 3 ppm is a significant improvement towards acetone
posite sensors have a higher baseline impedance (2.55 MΩ @100 Hz) sensing.
due to the electron transfer from ZnO to TiO2 establishing a high space The relative sensor response (%) is defined as the ratio between the
charge region. In addition, the higher concentration of oxygen vacancies change in impedance upon exposure to acetone and the N2 baseline
in the composite thin films is favorable for chemisorbed O- ions on the impedance. The following expression was used to calculate the same.
sensor surface. These observations are in good agreement with the XPS
ZN2 − Zg
spectra which suggest relatively large number of OV related active sites SensorResponse(%) = × 100 (2)
ZN2
at the surface. This leads to an improvement in the acetone sensing
performance for the composite sensors than the single-phase TiO2 and Here, ZN2 and Zg denotes the impedance of the sensor at N2 baseline
ZnO based sensors. A change in the impedance from 2.55 MΩ to and various acetone concentrations. The sensor response (%) for all the

7
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

Fig. 12. (Colour online) Response and recovery plot (a) TiO2 (b) ZnO (c) TiO2 - ZnO sensors at 100 Hz.

acetone sensors under investigation are depicted in Fig. 11. The com­ phase TiO2 and ZnO based sensors are 53 s and 165 s respectively.
posite thin film based sensors exhibit the highest sensor response of The room temperature acetone sensors based on single-phase TiO2 is
93.83% compared to single-phase TiO2 (40.3%) and ZnO (52.41%) for found to have faster response and recovery times but poor sensor
70 ppm acetone. The response for 3 ppm acetone concentration is found response.
to be 3.95% for the composite sensors. The higher surface roughness and The bulk and surface contribution of the sensing material upon
more OV sites are the significant reasons for the increased sensor exposure to acetone was evaluated through electrochemical impedance
response exhibited by the TiO2 – ZnO composite thin film based sensors spectroscopy (EIS). A sinusoidal peak-to-peak voltage of 1 V with fre­
[39,40]. Together with this, the contact potentials associated with quency varying from 100 Hz to 5 MHz was applied to all the sensors
localized states at the heterojunction interface can easily trap and pro­ resulting in a sinusoidal current. The obtained electrical impedance
mote the reaction with acetone molecules which in turn affects the Z(ω) = Z′(ω) +iZ′ (ω) was analysed for different sensors through the
impedance of the composite sensors [41]. Cole-Cole plots. Fig. 13 shows the Cole-Cole plots for the proposed
In addition to the sensor response, the response and recovery times sensors for acetone concentrations ranging from 10 to 70 ppm. The
are crucial parameters for evaluating the merit of a sensor. In the present single-phase TiO2 and ZnO-based sensors exhibit an elongated semi­
case, the response time is defined as the time required to lose 90% of its circle with a small tail towards the lower frequency. For the TiO2 – ZnO
impedance measured at 100 Hz when the acetone concentration is based composite sensors, a perfect semicircle is observed without any
increased from 0 to 70 ppm. The same is called as the recovery time tail-like features at lower frequency. For all the sensors, the higher
when the concentration is decreased. The dynamic response and re­ intercept on the X-axis is found to be shifting towards lower values with
covery times of all the proposed sensors are illustrated in Fig. 12. We an increase of acetone concentration. This is due to the release of the
have recorded three complete cycles of the response and recovery with trapped electrons because of the interaction between acetone molecules
good repeatability. The estimated response time for the composite sen­ and chemisorbed oxygen species on the active sites of the sensor surface
sors is 42 s whereas it is found to be 51 s and 130 s for the single-phase [42]. The shift of this intercept towards lower values is the highest for
TiO2 and ZnO based sensors respectively. With further increase of the composite thin film based sensors. This is attributed to the higher
acetone concentration beyond 70 ppm, the sensing behaviour starts to surface roughness and a greater number of oxygen vacancy sites as
saturate for the composite sensors resulting in a recovery time of 149 s validated by AFM and XPS investigations. The Col-Cole plots were fitted
for the composite sensors. The recovery time observed for the single- with the relevant equivalent circuit using the EISSA software [26]. In the

8
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

Fig. 13. (Colour online) Cole-Cole plot of (a) TiO2 (b) ZnO (c) TiO2 - ZnO sensors and the inset referred to fitted equivalent circuit.

frequency dependent system, a typical equivalent circuit for EIS consists significant increase in capacitance (C) is also observed for the TiO2 –
of constant phase element (CPE), contact resistance (RC), grain boundary ZnO composite thin film-based sensors with acetone concentration. This
resistance (RGB) and Warburg element (W). The impedance corre­ is due to the change in the dielectric properties of the material because
sponding to the constant phase element is given by ZCPE = Q(i21πf)n . Here A of the increase in the surface electron concentration caused by acetone
is a real parameter, f is the frequency and the index n varies from 0 to 1. molecules. The Bode plot analysis shows the maximum impedance
For n = 0, the CPE behaves as a pure resistor and for n = 1, it behaves as changes between 100 Hz and 50 kHz for the single phase TiO2 and ZnO
a pure capacitor. In the present study, the CPE describes the capacitance thin films (Fig. 15). A relatively larger change in the impedance is
between the sensor surface and the electrode. RC represents the resis­ observed between 100 Hz and 5 kHz for the composite thin film based
tance at the interface between the electrode and the sensing layer. The sensors. At lower frequencies, the polarization of the chemisorbed
grain boundary resistance RGB varies due to the conduction variation acetone molecules in composite sensors increases which leads to a
between the grains. The Warburg impedance (W) appears as a straight decrease in the impedance.
line with a semicircle in the lower frequency region, which is due to the In order to check the selectivity of the proposed sensors, several
diffusion of ions or electron transfer at the sensor/electrode interface other VOCs such as ethanol, isopropyl alcohol, acetophenone, methanol
[43]. The corresponding equivalent circuit is provided in the inset of the and 2-methoxy ethanol at 70 ppm concentrations were studied. The
Cole-Cole plots. The CPE is connected in parallel to the RGB and W. The corresponding sensors response to these VOCs are shown in Fig. 16. All
variation in the fitted values of RGB and C with respect to the acetone the proposed sensors were found to have higher response towards
concentration for all the proposed sensors is shown in Fig. 14. A acetone than other VOCs studied. The composite sensors have a response
remarkable change in the RGB is observed for the composite sensors which is 2.32 and 1.79 times higher than that of single-phase TiO2 and
when compared with the single-phase ones. This is primarily attributed ZnO respectively. In particular, the response of the composite sensors
to the increased electron densities at the grain boundaries because of the towards acetone (93.83%) is significantly higher than the other VOCs.
trapped chemisorbed electrons. This enhances the grain boundary con­ The composite sensors exhibit the sensor response of 39.92%, 13.4%,
ductivity with increase in the acetone concentration [42]. Similarly, a 25%, 21.5%, and 43.81% for acetophenone, 2-methoxy ethanol,
ethanol, methanol, and isopropyl alcohol, respectively. Among these,

9
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

Fig. 14. (Colour online) Calculated (a) Grain boundary resistance (RGB) and (b) C – capacitance of the sensors at different acetone concentrations.

isopropyl alcohol has moderate response, which is 0.46 times lower than oxygen molecules capture an electron from the conduction band of the
the sensor response to acetone. Moreover, 2-methoxy ethanol exhibits a n-type metal oxide and form a chemisorbed anion through a dangling
very poor response to the composite sensor. bond. Different oxygen species (O-2, O- and O2-) are involved in the
When the analyte VOC (acetone) interacts with the surface of the chemisorption depending on the temperature of the environment [50].
sensor, which operates at room temperature (29 ± 2 ◦ C), humidity is The depletion region formed between the grains causes an increase in
one of the main elements impacting the performance of the sensor [44]. the impedance and the concentration of free electrons at the surface is
Therefore, the baseline impedance of the TiO2-ZnO composite sensor decreased by the adsorbed oxygen molecules. When the sensor surface is
with respect to various relative humidity (RH) percentages and its exposed to acetone, it interacts with the chemisorbed anions according
associated sensor response for 70 ppm of acetone were examined. As per to the following equations [51]:
the report [26,45], we created and maintained constant relative hu­
CH3 COCH 3(gas) +4O−2(ads) →3H2 O(vap) +3CO2(gas) +4e− (return to sensor)
midity levels of 12%, 32%, 54%, 75%, and 90% using a saturated so­
lution of potassium hydroxide (KOH), magnesium chloride (MgCl2), (3)
magnesium nitrate (MgNO3), sodium chloride (NaCl), and sodium car­
CH3 COCH3(gas) + 8O−(ads) →3H2 O(vap) −
+ 3CO2(gas) + 8e (return to sensor)
bonate (Na2CO3), respectively. The percentage of RH was then contin­
uously monitored using a DHT11 temperature and humidity sensor (4)
interfaced with the Arduino-UNO IDE. Fig. 17a depicts the change in
sensor impedance as a function of RH (%). The baseline impedance of CH3 COCH 3(gas) +8O2−(ads) →3H2 O(vap) +3CO2(ads) +16e− (return to sensor)
the composite sensor decreased at a higher RH%, which is due to the (5)
Grotthuss chain reaction, as a result of the adsorption of water molecules
Acetone adsorption causes an increase in electrical conductivity
on the sensor surface [26]. The sensor response of the composite thin
because of the return of the trapped electrons to the sensor surface. This
films towards 70 ppm of acetone at different RH% has been shown in
leads to a drop in the overall impedance. For room temperature sensing,
Fig. 17b. At 90% RH, the sensor response dropped from 93.83% to
O−2 is the most dominant species involved in the chemisorption and
76.61%, due to the physisorbed and chemisorbed water molecules on
hence the sensing in the present study takes place through Eq. 3.
the surface of the composite sensor, which limit the interaction of the
In composite sensors, an additional band offset is created between
acetone molecule with the active sites of the sensor [46]. That is the
the grains of TiO2 and ZnO. The band offset type depends on the posi­
main cause of the sensor response being reduced at 90% relative
tions of the conduction band minimum (CBM) and valence band
humidity.
maximum (VBM) of TiO2 and ZnO. The positions of CBM and VBM of
TiO2 and ZnO were calculated by the Mulliken electronegativity formula
[52]. However, this method ignores the lattice mismatch and the
3.4. Acetone sensing mechanism
structural factors of the semiconductors. The approximate positions of
the VBM (EVB) and CBM (ECB) can be expressed as,
The grain boundary model [47,48] is widely accepted for the sensing
of VOCs by metal oxides. It involves the adsorption and desorption of gas EVB = X − Ee + 0.5Eg (6)
molecules on the sensor surface which in turn alters its conductivity. The
acetone sensing mechanism for single-phase TiO2 and ZnO thin films is ECB = EVB − Eg (7)
explained in Fig. 18 which is adopted from Xu et al. [49]. Since both
Here, X is the electronegativity of TiO2 or ZnO (5.82 eV and 5.78 eV
these materials are n-type semiconductors, we have outlined the basic
[53] respectively), Ee is the energy of free electrons at the hydrogen scale
mechanism of acetone sensing by n-type single-phase metal oxide. When
(= 4.5 eV vs. SHE) and Eg is the band gap of the semiconductor. The
an n-type metal oxide (TiO2 or ZnO) is exposed to air, atmospheric ox­
positions of the CBM for TiO2 and ZnO were calculated to be − 0.23 eV
ygen molecules get adsorbed on the sensors surface. These chemisorbed

10
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

Fig. 15. (Colour online) Bode plots of (a) TiO2 (b) ZnO and (c) TiO2 - ZnO sensors as a function of frequency.

and − 0.33 eV respectively. The corresponding positions for the VBM


were found to be 2.87 eV and 2.89 eV respectively. Based on these
values, the TiO2 – ZnO heterojunction exhibits type – I band offset as
shown in Fig. 19. This is in accordance with our XPS results for Ti-2p
associated peaks in composite thin films which showed a slight shift
towards lower binding energy values. In order to equilibrate the Fermi
energy, the electron transfer takes place from the CB of ZnO to the CB of
TiO2 because the position of the former is more negative than the latter.
As a result, electrons accumulate on the TiO2 side facilitating a larger
concentration of active sites at the interface. The charge redistribution
close to the heterojunction interface caused by the type-I band offset
significantly widens the depletion region. In addition to this, more
electrons are trapped at the heterojunction interface due to oxygen
adsorption. The increased thickness of the depletion region leads to an
increase in the impedance of the sensor. Upon exposure to acetone, the
chemisorbed oxygen readily interacts with it and the excess electrons
return to the sensor surface. This results in the narrowing of the deple­
tion region and hence a decrease in the impedance.
Table 3 summarizes the performance of the proposed sensors in
comparison with the TiO2 based sensors reported in the literature.
Fig. 16. (Colour online) Selectivity plot for the proposed sensors with
Compared to the other reports, the TiO2 – ZnO composite thin film based
various VOCs. sensors exhibit a better detection threshold of 3 ppm under room tem­
perature operations. The acetone sensors reported by Bhowmik et al.
[54] have a higher relative response of 178.6% at 150 ◦ C compared to

11
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

Fig. 17. (Colour online) (a) Different percentage of RH vs composite sensor impedance (b) Sensor response to 70 ppm acetone at different RH% of the compos­
ite sensor.

Fig. 18. (Colour online) Schematic diagram of the single-phase (TiO2 or ZnO) acetone sensing mechanism.

Fig. 19. (Colour online) Schematic diagram of the TiO2-ZnO heterojunction acetone sensing mechanism.

12
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

Table 3
Comparison of acetone sensing performance of existing TiO2 based sensors with the proposed TiO2 – ZnO composite thin film based sensors.
Material Fabrication technique Acetone Sensor Response/ Operating Limit of Ref.
concentration response Recovery time Temperature detection
(ppm)a; (ppb)b (s) ( ◦ C) (ppm/ppb)

TiO2 – ZnO Sol-gel spin coating technique 70a 93.83a 42/149 Room 3 ppm This
Thin films temperature work
(29 ◦ C ± 2)
TiO2/ SiO2 Dip coating technique 500a 178.6a 42.11/44.05 150 500 ppm [54]
Thin films
Pd/TiO2/Si Dip coating 50a 33a 7/20 200 0.5 ppm [17]
Thin films technique
a b
TiO2 Hydrothermal method 250 33.72 46/24 330 10 ppm [55]
flower-like
Nanomaterials
W – doped Hydrothermal method 500a 173.6b 14/9 240 5 ppm [56]
Nanoporous TiO2
TiO2/SnO2 Pulsed Laser deposition method 100a 301.5b 15/57 300 0.02 ppm [22]
Thin films
a a
GO/TiO2 Sol-gel technique 200 18.6 80/- Room 100 ppm [23]
Nanocomposites temperature
TiO2 @B-TiO2 @ZnO CVD method followed by 100a 80.45b 5/26 275 170 ppb [24]
NPs coprecipitation technique
TiO2/ZnO Co-precipitation method 100b 179a 13/11 Room 100 ppb [25]
Nanocomposite temperature
a a
TiO2/RGO Hummers method followed by 100 81.24 10/23 Room 20 ppm [20]
Nanocomposites Ultrasonication temperature

a Ra − Rg Ra
Note: denote the sensor response = × 100(%); b denote the sensor response =
Ra Rg

the TiO2 – ZnO composite sensors proposed in the present study. How­ editing. Murugesan Muthukumar: Data curation, Investigation,
ever, the major advantage of the proposed sensors is their room tem­ Methodology, Writing – original draft.
perature operation which is ideal for the easy diagnosis of diabetes
mellitus. However, the proposed sensors have moderate response time Declaration of Competing Interest
and slow recovery time compared to the other acetone sensors which
need to be improved. The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
4. Conclusions the work reported in this paper.

The thin films of single-phase TiO2 and ZnO and their composites Data availability
were coated on glass substrates by the commercially viable sol-gel spin
coating. XRD reveals the anatase and wurtzite phase for TiO2 and ZnO Data will be made available on request.
respectively. The occurrence of diffraction peaks corresponding to both
anatase and wurtzite confirmed the composite nature of TiO2 – ZnO Acknowledgment
mixed layer films. The formation of heterojunction between TiO2 and
ZnO grains is confirmed from the shift of the Ti-2p XPS peak towards The authors would like to acknowledge Department of Science and
lower binding energies compared to the single-phase ones. XPS also Technology (DST), New Delhi, India for providing the financial support
reveals the presence of more oxygen vacancy sites in composite thin through FIST (Fund for Improvement of S&T Infrastructure in Higher
films than the single phase films. The fabricated sensors based on these Education Institution) project [no. SR/FST/ETI-015/2011]. The authors
thin films were studied for their impedometric sensing performance also would like to acknowledge Vellore Institute of Technology, Vellore,
towards various VOCs. The composite sensors exhibit superior selec­ India for providing the financial support through VIT SEED GRANT
tivity towards acetone than other VOCs studied. They exhibit an (RGEMS) [SG20230043]. The authors would also like to acknowledge
improved LOD of 3 ppm for acetone and a maximum sensor response of Prof. Zachariah C. Alex for useful discussions and suggestions.
93.83% for 70 ppm acetone when compared with single-phase of TiO2
and ZnO thin film based sensors. The higher acetone sensing perfor­ References
mance of TiO2 – ZnO composite thin film based sensors are attributed to
the higher surface roughness and the presence of greater number of [1] V. Saasa, T. Malwela, M. Beukes, M. Mokgotho, C.P. Liu, B. Mwakikunga, Sensing
technologies for detection of acetone in human breath for diabetes diagnosis and
oxygen vacancies. In addition, the composite sensors have better monitoring, Diagnostics 8 (2018) 1–17, https://doi.org/10.3390/
response time of 42 s compared to the single-phase TiO2 or ZnO based diagnostics8020012.
sensors. The room temperature operability of the proposed acetone [2] N. Goel, K. Kunal, A. Kushwaha, M. Kumar, Metal oxide semiconductors for gas
sensing, Eng. Rep. (2022) 1–22, https://doi.org/10.1002/eng2.12604.
sensors makes them ideal for the easy diagnosis of diabetes mellitus. [3] M. Ahmadipour, A.L. Pang, M.R. Ardani, S.Y. Pung, P.C. Ooi, A.A. Hamzah, M.F.M.
However, the slow recovery time of 149 s for the composite thin film R. Wee, M.A.S.M. Haniff, C.F. Dee, E. Mahmoudi, A. Arsad, M.Z. Ahmad, U. Pal, K.
based acetone sensors need improvement through the aid of possible M. Chahrour, S.A. Haddadi, Detection of breath acetone by semiconductor metal
oxide nanostructures-based gas sensors: a review, Mater. Sci. Semicond. Process.
functional materials. 149 (2022) 106897, https://doi.org/10.1016/j.mssp.2022.106897.
[4] V. Amiri, H. Roshan, A. Mirzaei, G. Neri, A.I. Ayesh, Nanostructured metal oxide-
CRediT authorship contribution statement based acetone gas sensors: a Review, Sensors 20 (2020) 1–25, https://doi.org/
10.3390/s20113096.

Meher Samir Ranjan: Supervision, Validation, Writing – review &

13
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

[5] H.J. Cho, S.J. Choi, N.H. Kim, I.D. Kim, Porosity controlled 3D SnO2 spheres via [31] A. Lakshmanan, Z.C. Alex, S.R. Meher, Cu2O thin films grown by magnetron
electrostatic spray: selective acetone sensors, Sens. Actuators B: Chem. 304 (2020) sputtering as solar cell absorber layers, Mater. Sci. Semicond. Process. 148 (2022)
127350, https://doi.org/10.1016/j.snb.2019.127350. 106818, https://doi.org/10.1016/j.mssp.2022.106818.
[6] Q. Jia, H. Ji, Y. Zhang, Y. Chen, X. Sun, Z. Jin, Rapid and selective detection of [32] A. Sahai, N. Goswami, Probing the dominance of interstitial oxygen defects in ZnO
acetone using hierarchical ZnO gas sensor for hazardous odor markers application, nanoparticles through structural and optical characterizations, Ceram. Int. 40
J. Hazard. Mater. 276 (2014) 262–270, https://doi.org/10.1016/j. (2014) 14569–14578, https://doi.org/10.1016/j.ceramint.2014.06.041.
jhazmat.2014.05.044. [33] M.M. Alam, A.M. Asiri, M.T. Uddin, M.A. Islam, M.M. Rahman, Wet-chemically
[7] N. Chen, Y. Li, D. Deng, X. Liu, X. Xing, X. Xiao, Y. Wang, Acetone sensing prepared low-dimensional ZnO/Al2O3/Cr2O3 nanoparticles for xanthine sensor
performances based on nanoporous TiO2 synthesized by a facile hydrothermal development using an electrochemical method, RSC Adv. 8 (2018) 12562–12572,
method, Sens. Actuators B: Chem. 238 (2017) 491–500, https://doi.org/10.1016/j. https://doi.org/10.1039/c8ra01734d.
snb.2016.07.094. [34] P. Van der Heide. X-Ray Photoelectron Spectroscopy: An Introduction to Principles
[8] F. Chen, M. Yang, X. Wang, Y. Song, L. Guo, N. Xie, X. Kou, X. Xu, Y. Sun, G. Lu, and Practices, 1st ed..,, John Wiley & Sons, Inc, Hoboken, New Jersey, 2011.
Template-free synthesis of cubic-rhombohedral-In2O3 flower for ppb level acetone [35] M. Perez-Gonzalez, S.A. Tomas, Surface chemistry of TiO2-ZnO thin films doped
detection, Sens. Actuators B: Chem. 290 (2019) 459–466, https://doi.org/ with Ag. Its role on the photocatalytic degradation of methylene blue, Catal. Today
10.1016/j.snb.2019.04.013. 360 (2021) 129–137, https://doi.org/10.1016/j.cattod.2019.08.009.
[9] Q. Zhai, B. Du, R. Feng, W. Xu, Q. Wei, A highly sensitive gas sensor based on Pd- [36] Y. Li, H. Yang, J. Tian, X. Hu, H. Cui, Synthesis of In2O3 nanoparticle/TiO2
doped Fe3O4 nanoparticles for volatile organic compounds detection, Anal. nanobelt heterostructures for near room temperature ethanol sensing, RSC Adv. 7
Methods 6 (2014) 886–892, https://doi.org/10.1039/c3ay41867g. (2017) 11503–11509, https://doi.org/10.1039/c7ra00011a.
[10] S. Wei, S. Li, R. Wei, S. Liu, W. Du, Different morphologies of WO3 and their [37] Y. Dessie, S. Tadesse, R. Eswaramoorthy, Surface roughness and electrochemical
exposed facets-dependent acetone sensing properties, Sens. Actuators B: Chem. 329 performance properties of biosynthesized α-MnO2/NiO-based polyaniline ternary
(2021) 129188, https://doi.org/10.1016/j.snb.2020.129188. composites as efficient catalysts in microbial fuel cells, J. Nanomater 3 (2021)
[11] X. Tian, X. Cui, T. Lai, J. Ren, Z. Yang, M. Xiao, B. Wang, X. Xiao, Y. Wang, Gas 1–21, https://doi.org/10.1155/2021/7475902.
sensors based on TiO2 nanostructured materials for the detection of hazardous [38] A.A. Ward, State Art. Dielectr. Mater. Adv. Appl. (2015) 1–70.
gases: a review, Nano Mater. Sci. 3 (2021) 390–403, https://doi.org/10.1016/j. [39] M. Kumar, B. Singh, P. Yadav, V. Bhatt, M. Kumar, K. Singh, A.C. Abhyankar,
nanoms.2021.05.011. A. Kumar, J.H. Yun, Effect of structural defects, surface roughness on sensing
[12] W. Ge, S. Jiao, Z. Chang, X. He, Y. Li, Ultrafast response and high selectivity toward properties of Al doped ZnO thin films deposited by chemical spray pyrolysis
acetone vapor using hierarchical structured TiO2 nanosheets, ACS Appl. Mater. technique, Ceram. Int. 43 (2017) 3562–3568, https://doi.org/10.1016/j.
Interfaces 12 (2020) 13200–13207, https://doi.org/10.1021/acsami.9b23181. ceramint.2016.11.191.
[13] S.T. Navale, Z.B. Yang, C. Liu, P.J. Cao, V.B. Patil, N.S. Ramgir, R.S. Mane, F. [40] E. Ciftyurek, Z. Li, K. Schierbaum, Adsorbed oxygen ions and oxygen vacancies:
J. Stadler, Enhanced acetone sensing properties of titanium dioxide nanoparticles their concentration and distribution in metal oxide chemical sensors and
with a sub-ppm detection limit, Sens. Actuators B: Chem. 255 (2018) 1701–1710, influencing role in sensitivity and sensing mechanisms, Sensors 23 (2023) 1–30,
https://doi.org/10.1016/j.snb.2017.08.186. https://doi.org/10.3390/s23010029.
[14] D. Zhao, X. Zhang, W. Wang, L. Sui, X. Bai, H. Song, C. Guo, Y. Xu, X. Cheng, [41] A.J.T. Naik, I.P. Parkin, R. Binions, Gas sensing studies of an n-n heterojunction
S. Gao, L. Huo, Ultra-small TiO2 nanocubes with highly active (0 0 1) facet for metal oxide semiconductor sensor array based on WO3 and ZnO composites, Proc.
acetone fast detection and diagnosis of diabetes, Microchem. J. 184 (2023) 108122 IEEE Sens. (2013), https://doi.org/10.1109/ICSENS.2013.6688509.
https://doi.org/10.1016/j.microc.2022.108122. [42] M.I. Ikim, G.N. Gerasimov, V.F. Gromov, O.J. Ilegbusi, L.I. Trakhtenberg,
[15] P. Bindra, A. Hazra, Selective detection of organic vapors using TiO2 nanotubes Synthesis, structural and sensor properties of nanosized mixed oxides based on
based single sensor at room temperature, Sens. Actuators B: Chem. 290 (2019) In2O3 particles, Int. J. Mol. Sci. 24 (2023) 1–23, https://doi.org/10.3390/
684–690, https://doi.org/10.1016/j.snb.2019.03.115. ijms24021570.
[16] Z. Feng, L. Zhang, W. Chen, Z. Peng, Y. Li, A strategy for supportless sensors: [43] M. Velumani, S.R. Meher, Z.C. Alex, Impedometric humidity sensing characteristics
fluorine doped TiO2 nanosheets directly grown onto Ti foam enabling highly of SnO2 thin films and SnO2–ZnO composite thin films grown by magnetron
sensitive detection toward acetone, Sens. Actuators B: Chem. 322 (2020) 128633 sputtering, J. Mater. Sci. Mater. Electron. 29 (2018) 3999–4010, https://doi.org/
https://doi.org/10.1016/j.snb.2020.128633. 10.1007/s10854-017-8342-z.
[17] B. Bhowmik, A. Hazra, K. Dutta, P. Bhattacharyya, Nanocrystalline p-TiO2 based [44] G.K. Mani, J.B.B. Rayappan, A highly selective room temperature ammonia sensor
MIS device for efficient acetone detection, 2014-Decem, Proc. IEEE Sens. (2014) using spray deposited zinc oxide thin film, Sens. Actuators B: Chem. 183 (2013)
293–296, https://doi.org/10.1109/ICSENS.2014.6984991. 459–466, https://doi.org/10.1016/j.snb.2013.03.132.
[18] W. Yang, H. Shen, J. Ge, B. Xu, Improving TiO2 gas sensing selectivity to acetone [45] L. Greenspan, Humidity fixed points of binary saturated aqueous solutions, J. Res.
and other gases via a molecular imprinting method, Nanotechnology 32 (2021) Natl. Bur. Stand. A. Phys. Chem. 81A (1977) 89–96, https://doi.org/10.6028/
155503, https://doi.org/10.1088/1361-6528/abd818. jres.081A.011.
[19] S. Yang, G. Lei, H. Xu, Z. Lan, Z. Wang, H. Gu, Metal oxide based heterojunctions [46] X. Chen, J. Hu, P. Chen, M. Yin, F. Meng, Y. Zhang, UV-light assisted NO2 gas
for gas sensors: a review, Nanomaterials 11 (2021) 1–26, https://doi.org/10.3390/ sensor based on WS2/PbS heterostructures with full recoverability and reliable
nano11041026. anti-humidity ability, Sens. Actuators B: Chem. 339 (2021) 129902 https://doi.
[20] M. Daneshnazar, B. Jaleh, M. Eslamipanah, R.S. Varma, Optical and gas sensing org/10.1016/j.snb.2021.129902.
properties of TiO2/RGO for methanol, ethanol, and acetone vapors, Inorg. Chem. [47] N.A. Isaac, I. Pikaar, G. Biskos, Metal oxide semiconducting nanomaterials for air
Commun. 145 (2022) 110014, https://doi.org/10.1016/j.inoche.2022.110014. quality gas sensors: operating principles, performance, and synthesis techniques,
[21] M.M. Hashemi, A. Nikfarjam, H. Hajghassem, N. Salehifar, Hierarchical dense Microchim. Acta 189 (2022) 1–22, https://doi.org/10.1007/s00604-022-05254-0.
array of ZnO nanowires spatially grown on ZnO/TiO2 nanofibers and their [48] A. Dey, Semiconductor metal oxide gas sensors: a review, Mater. Sci. Eng. B. 229
Ultraviolet activated gas sensing properties, J. Phys. Chem. C. 124 (2020) (2018) 206–217, https://doi.org/10.1016/j.mseb.2017.12.036.
322–335, https://doi.org/10.1021/acs.jpcc.9b07207. [49] H. Xu, J. Xu, H. Li, W. Zhang, Y. Zhang, Z. Zhai, Highly sensitive ethanol and
[22] B. Sharma, A. Sharma, J.H. Myung, Highly selective detection of acetone by TiO2- acetone gas sensors with reduced working temperature based on Sr-doped BiFeO3
SnO2 heterostructures for environmental biomarkers of diabetes, Sens. Actuators B: nanomaterial, J. Mater. Res. Technol. 17 (2022) 1955–1963, https://doi.org/
Chem. 349 (2021) 130733, https://doi.org/10.1016/j.snb.2021.130733. 10.1016/j.jmrt.2022.01.137.
[23] V. Kumar, R. Madan, D. Mohan, Sol–gel derived GO/TiO2 nanocomposites for room [50] H.J. Kim, J.H. Lee, Highly sensitive and selective gas sensors using p-type oxide
temperature acetone gas sensing application, J. Mater. Sci. Mater. Electron. 33 semiconductors: overview, Sens. Actuators B: Chem. 192 (2014) 607–627, https://
(2022) 7655–7667, https://doi.org/10.1007/s10854-022-07914-6. doi.org/10.1016/j.snb.2013.11.005.
[24] Y. Feng, W. Yang, Y. Li, H. Shen, The effect of near-surface electron trapping layer [51] C. Li, P.G. Choi, K. Kim, Y. Masuda, High performance acetone gas sensor based on
on the acetone sensing performance of black TiO2 capped with ZnO, ultrathin porous NiO nanosheet, Sens. Actuators B: Chem. 367 (2022) 132143,
Nanotechnology 33 (2022) 275712, https://doi.org/10.1088/1361-6528/ac5aea. https://doi.org/10.1016/j.snb.2022.132143.
[25] K.G. Krishna, S.R. Parne, P. Nagaraju, Facile synthesis and enhanced sensing [52] C. Chen, W. Bi, W. Yuan, L. Li, Hydrothermal synthesis of the CuWO4/ZnO
properties of ZnO/TiO2 nanocomposite at room temperature, Mater. Sci. Eng. B. composites with enhanced photocatalytic performance, ACS Ome 5 (2020)
297 (2023) 116734, https://doi.org/10.1016/j.mseb.2023.11673. 13185–13195, https://doi.org/10.1021/acsomega.0c01220.
[26] M. Velumani, S.R. Meher, Z.C. Alex, Composite metal oxide thin film based [53] N. Bai, X. Liu, Z. Li, X. Ke, K. Zhang, Q. Wu, High-efficiency TiO2/ZnO
impedometric humidity sensors, Sens. Actuators B: Chem. 301 (2019) 127084, nanocomposite photocatalysts by sol-gel and hydrothermal methods, J. Solgel Sci.
https://doi.org/10.1016/j.snb.2019.127084. Technol. 99 (2021) 92–100, https://doi.org/10.1007/s10971-021-05552-8.
[27] R.S. Gaikwad, G.R. Patil, B.N. Pawar, R.S. Mane, S.H. Han, Liquefied petroleum gas [54] B. Bhowmik, K. Dutta, N. Banerjee, A. Hazra, P. Bhattacharyya, Low temperature
sensing properties of sprayed nanocrystalline zinc oxide thin films, Sens. Actuators acetone sensor based on Sol-gel grown nano TiO2 thin film, IEEE Int. Conf. Emerg.
A: Phys. 189 (2013) 339–343, https://doi.org/10.1016/j.sna.2012.10.005. Trends Comput. Commun. Nanotechnology, ICE-CCN 2013, 2013, pp. 553–557,
[28] M. Perez-Gonzalez, S.A. Tomas, J. Santoyo-Salazar, M. Morales-Luna, Enhanced https://doi.org/10.1109/ICE-CCN.2013.6528561.
photocatalytic activity of TiO2-ZnO thin films deposited by dc reactive magnetron [55] W. Yang, Q. Ou, X. Yan, L. Liu, S. Liu, H. Chen, Y. Liu, High sensing performance
sputtering, Ceram. Int. 43 (2017) 8831–8838, https://doi.org/10.1016/j. toward acetone vapor using TiO2 Flower-Like nanomaterials, Nanoscale Res. Lett.
ceramint.2017.04.016. 17 (1) (2022) 9, https://doi.org/10.1186/s11671-022-03721-4.
[29] B. Bharti, S. Kumar, H.N. Lee, R. Kumar, Formation of oxygen vacancies and Ti3+ [56] L. Wang, X. Xing, N. Chen, R. Zhao, Z. Wang, T. Zou, Z. Wang, Y. Wang, W-doped
state in TiO2 thin film and enhanced optical properties by air plasma treatment, nanoporous TiO2 for high performances sensing material toward acetone gas,
Sci. Rep. 6 (1) (2016) 12, https://doi.org/10.1038/srep32355. J. Nanostruct. 10 (2020) 148–156, https://doi.org/10.22052/JNS.2020.01.016.
[30] U. Diebold, T.E. Madey, TiO2 by XPS, Surf. Sci. Spectra 4 (1996) 227–231, https://
doi.org/10.1116/1.1247794.

14
M. Murugesan and S.R. Meher Sensors and Actuators: A. Physical 369 (2024) 115181

Muthukumar M received his M.Phil. Degree from Alagappa University, Karaikudi, India, S. R. Meher currently working as an Associate Professor in the Department of Physics,
in 2020. He is pursuing his Ph.D. in the Department of Physics, School of Advanced Sci­ School of Advanced Sciences, VIT – Vellore, India. His current research interests include
ences VIT – Vellore, India. His research interests include the development of thin film and thin film and nanostructured Semiconductors for gas sensing, optoelectronic and self-
nanostructure based sensors for Volatile Organic Compound (VOC) sensing applications. cleaning smart window applications.

15

You might also like