High Performance Co Gas Sensor Based On Zno Nanoparticles: M. Hjiri F. Bahanan M. S. Aida L. El Mir G. Neri

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Inorganic and Organometallic Polymers and Materials (2020) 30:4063–4071

https://doi.org/10.1007/s10904-020-01553-2

High Performance CO Gas Sensor Based on ZnO Nanoparticles


M. Hjiri1,2 · F. Bahanan1 · M. S. Aida1,3 · L. El Mir2 · G. Neri4

Received: 21 March 2020 / Accepted: 17 April 2020 / Published online: 23 April 2020
© Springer Science+Business Media, LLC, part of Springer Nature 2020

Abstract
Carbon monoxide sensor was fabricated using ZnO nanoparticles, synthesized by sol–gel technique, as sensing layer. The
morphology and structure of the prepared nanopowder were analyzed using X-ray diffraction (XRD), scanning and transmis-
sion electron microscopies (SEM and TEM). Photoluminescence (PL) measurements were carried to investigate the defects in
ZnO. The sensing tests were performed by a homemade setup. XRD pattern indicate that the prepared ZnO nanopowder has
a crystallite size average around 50 nm. TEM and SEM images reveal that the ZnO nanopowder is formed of agglomeration
of spherical particles with a size of 50 nm which is in good agreement with XRD analysis. The prepared gas sensor exhibits
a response of 74% towards 80 ppm of CO gas with a response/recovery times of 21 and 70 s, respectively at 250 °C and high
stability with time. The good sensing properties of ZnO nanoparticles towards CO gas indicate their potential application
for the fabrication of low power and highly selective sensors.

Keywords Zinc oxide · Nanoparticles · Selectivity · Sensor · CO

1 Introduction acoustic wave (SAW) devices [12], transparent conducting


oxide electrodes [13], solar cells [14], blue/UV light emit-
In the last years, gas sensors based on various semiconduct- ting devices [15], gas sensors [16, 17], etc.… Its conductiv-
ing metal oxides, such as tin oxide [1], iron oxide [2], cop- ity can be tailored by controlling the deviation from stoichi-
per oxide [3], titanium oxide [4], gallium oxide [5] and zinc ometry and by doping [18].
oxide [6], have attracted the attention of several researchers. Zinc oxide is an important candidate used as sensing layer
These metal oxide sensors exhibited strong response and towards hazardous gases [19–21]. In general, ZnO gas sen-
fast response/recovery times, in addition of good selectiv- sors possess several advantages such as low cost, easy manu-
ity, stability, compatibility with microelectronic devices [7]. facturing, and small size, in comparison with the traditional
Among them, ZnO has been shown to be useful materials analytical instruments. This oxide that can be prepared with
for monitoring various pollutant gases like CO, benzene, different morphologies (nanoparticles, nanotubes, nanorods,
ammonia, ­CO2, NOx. Zinc oxide is an n-type semiconductor etc.) has been largely studied for applications in gas sensors,
with a direct wide band gap (3.3 eV) [8–10]. This material because of its important optical, morphological, microstruc-
found broad ranging applications in varistors [11], surface tural and electric properties. Nundy et al. fabricated flower
shaped ZnO and suggested that it exhibited high response
toward NOx gas at 25 °C compared to ammonia, acetone,
* M. Hjiri CO, toluene [22]. Shaikh et al. noticed that ZnO nanorods
[email protected] obtained from a simple two step chemical method showed
1
Department of Physics, Faculty of Sciences, King Abdulaziz
good selectivity toward 40 ppm of NO2 gas and excellent
University, Jeddah 21589, Saudi Arabia repeatability at lower gas concentration of 2 ppm [23].
2
Laboratory of Physics of Materials and Nanomaterials
Kanaparthi et al. demonstrated that ZnO nanoflakes were
Applied at Environment, Faculty of Sciences of Gabes, very sensitive toward 80% of NH3 at 250 °C [24]. Kim et al.
6072 Gabes, Tunisia reported that ZnO nanofibers (NFs) synthesized by the sim-
3
Center of Nanotechnology, King Abdulaziz University, ple electrospinning technique exhibited strong response and
Jeddah, Saudi Arabia high selectivity toward 10 ppm of ­H2 gas [25].
4
Department of Engineering, University of Messina,
98166 Messina, Italy

13
Vol.:(0123456789)
4064 Journal of Inorganic and Organometallic Polymers and Materials (2020) 30:4063–4071

ZnO nanoparticles were obtained by several techniques 2 Experiments


such as hydrothermal [21], ball milling [26], co-precipitation
[27], sol–gel [28] etc… The sol–gel is the most used process 2.1 ZnO Preparation
because it is low cost, simple and easy to control.
Photoluminescence (PL) is a largely used technique for The sol–gel route was used to elaborate ZnO nanoparticles
defects detection in ZnO band gap [29–33]. It is well recog- using 16 g of zinc acetate (Zn(CH3CO2)2·2H2O) in 112 ml
nized that PL spectra of ZnO exhibits mainly two peaks, near of methyl alcohol. After magnetic stirring for 15 min, 200 ml
band emission (NBE) in the UV range, due to the excition of ethanol were added to the solution. The obtained solution
recombination [34] and a visible emission bands originating was poured in an autoclave and dried in supercritical condi-
from the extrinsic and extrinsic defects in ZnO such as Vo tions of ethyl alcohol (Tc = 243 °C; Pc = 63.6 bars) [40]. The
[35], Zni [36] and Oi [37] and Vzn [38]. obtained powders were subsequently heat treated at 400 °C
Carbon monoxide (CO), called the silent killer, is a for 2 h in air.
hazardous gas that causes death. The incomplete combus-
tion of fuels produced CO gas. This colorless and odorless 2.2 Sensing Test
toxic gas is found in the emission of automobile exhausts,
the burning of domestic fuels. Carbon monoxide affects We fabricated ZnO based sensor by printing a film of
strongly human health even at low concentration. At low thickness around 1–10 μm of the paste of ZnO powders
levels (above 70 ppm), CO can cause fatigue and nausea. on alumina substrate. The substrate had dimensions of
CO causes disorientation and death at concentrations above 6 mm × 3 mm with platinum interdigitated electrodes and
150 to 200 ppm [39]. Then it is important to fabricate and platinum heater on the backside. The sensor device and
use CO gas sensors. structure are reported in Fig. 1. The sensors were then
In present work, ZnO nanoparticles were prepared by introduced in the testing chamber. The temperature meas-
sol–gel technique for the detection of CO gas. XRD, SEM urements were in the range of 50–400 °C under a synthetic
and TEM were performed to investigate the structural and dry air. We utilized mass flow controllers to dilute the gases
morphological properties of the material. In order to have an coming from certified bottles in air at a given concentration.
idea about the defects in the structure of ZnO, PL measure- The sensor response, S, is defined as S = Ra/Rg S = R0 ∕R
ments are carried. Sensing tests were performed to show the where ­Ra is the resistance before gas injection in dry syn-
sensing properties. thetic air (20% ­O2 in nitrogen) and ­Rg is the electrical resist-
ance of the sensor after gas injection.

Fig. 1  Sensor device and


structure

13
Journal of Inorganic and Organometallic Polymers and Materials (2020) 30:4063–4071 4065

3 Results and Discussion

3.1 Microstructure and Morphology

Figure 2 presents the XRD spectrum of ZnO sample syn-


thesized by sol–gel technique and treated at 400 °C for 2 h.
The pattern shows various diffraction peaks assigned to ZnO
with no secondary phases or impurities. The obtained ZnO
powder is polycrystalline, its pattern is composed of dif-
ferent diffraction planes namely (100), (002), (101), (102),
(110), (013), (112) and (021) which are characteristic of the
Wurtzite hexagonal structure of ZnO according to JCPDS
database (Card 36-1451) [41]. The lattice parameters and
d-spacing, calculated from (002) peaks are a = 3.250 Å,
c = 5.206 Å and d = 0.26 nm which are very close to ZnO
structure [42].
The average crystallite size estimated by Scherer’s equa-
tion has been estimated to be equal to 53 nm.
0.9𝜆
D=
B cos 𝜃B (1)

where λ is the X-ray wavelength, θB is the maximum of the


Bragg diffraction peak (in radians) and B is the full width at
half maximum (FWHM) of the XRD peak.
Figure 3 reports SEM and TEM images presenting the
morphology of ZnO nanopowders after annealing at 400 °C
for 2 h. ZnO nanopowders were composed of agglomeration
of uniform spherical grains. From TEM small ZnO nanopar-
ticles with nanometric size are shown. The shape of crystal-
lites is prismatic with a narrow distribution of particle size. Fig. 3  a SEM and b TEM photographs of ZnO nanoparticles
The size of the majority of particles is around 50 nm which annealed at 400 °C
is in good agreement with the average crystallite size, D,
deduced from XRD pattern.

Pure ZnO
18000
(101)

Pure ZnO
Intensity (a.u.)

16000
Intensity (a.u.)
(100)
(002)

14000
(110)

(013)
(102)

(112)

12000
(021)

10000

8000
20 30 40 50 60 70 300 400 500 600 700 800 900
2θ (degree)
Wavelenght (nm)
Fig. 2  XRD spectrum of ZnO nanoparticles calcinated at 400 °C for
2h Fig. 4  Photoluminescence spectrum of ZnO nanoparticles

13
4066 Journal of Inorganic and Organometallic Polymers and Materials (2020) 30:4063–4071

Figure 4 presents the photoluminescence spectrum 250000

5000000
recorded from pure ZnO sample in the wavelength range 200000

between 350 and 850 nm at room temperature. Two peaks

Resistance (Ω )
150000
4000000
are observed in the photoluminescence spectrum. The first
peak is located at 380 nm and it corresponds to the near
100000

Resistance (Ω )
3000000
band edge (NBE) peak, it is due to the recombination of 50000

free excitons of ZnO [43]. The second peak is centered at 2000000 0


200 250 300 350 400
around 554 nm. It is well known that the green emission Temperature (°C)

originated from the electronic defects in the ZnO forbidden 1000000


band, Therefore PL emission is a powerful method for elec-
tronic defect determination in semiconductors. 0
To determine the origin of the PL green emission cen-
50 100 150 200 250 300 350 400
tered at 544 nm, we have proceeded to a Gaussian deconvo-
Temperature (°C)
lution of this peak (Fig. 5). As shown, the green emission
spectrum is composed of four peaks. The peak located at
Fig. 6  Variation of Resistance as a function of temperature in air
466 nm (2.65 eV) originates from the transition from con-
duction band to oxygen vacancy defect V ­ o, the peak located
at 500 nm (2.47 eV) is due to the transition from conduction of electrons into the conduction band indicating the semi-
band towards the interstitial oxygen defect (Oi), the peak conductor behavior of ZnO phase. It is well established that
at 546 nm (2.26 eV) is assigned to the transition from zinc ZnO is an n-type semiconductor with a wide band gap of
interstitial defect ­(Zni) to oxygen vacancy defect level ­(Vo) 3.2 eV [44]. The broad peak at around 270 °C shown in the
and the peak at 651 nm (1.90 eV) is due to the transition insert figure is due to the oxygen chemisorption on surface
from the complex VoZni to the defect Oi. Several authors material. The chemisorbed oxygen species reduces the elec-
have observed the same defect in ZnO band gap and PL peak trons concentration, thereafter the resistance enhancement
position [35–38]. as shown in insert Fig. 6.
The measured activation energy of the electrical conduc-
3.2 Sensing Properties tivity is estimated from Arrhenius plot as shown in Fig. 7;
the obtained activation energy is about 0.48 eV. It is worth
Figure 6 presents the variation of the resistance of ZnO noting that the conductivity activation energy correspond
nanopowders versus different operating temperatures. The to the Fermi level regarding the minimum conduction posi-
higher resistance at low temperature can be explained by tion, the low value of calculated activation energy (0.48 eV)
the disorder in the lattice which enhances the efficiency of indicates clearly that the prepared ZnO material is an n-type
scattering mechanism such as phonon scattering. As seen semiconductor. On the other hand, it is well known that
in Fig. 6, increasing the measurement temperature yields to the unintentional intrinsic defects in ZnO such as oxygen
the resistance decreases because of the thermal excitation

16

18000

15
PL emission (arb.unit)

16000
Ln(R)

14000

14 Ea=0.48 eV
12000

10000 13
2,5 2,6 2,7 2,8
450 500 550 600 650 700 750
-1
Wavelength (nm) 1000/KT (K )

Fig. 5  Deconvolution of the PL green emission Fig. 7  Activation energy calculated from Arrhenius plot

13
Journal of Inorganic and Organometallic Polymers and Materials (2020) 30:4063–4071 4067

vacancy (Vo), interstitial oxygen (­ Oi) and interstitial Zn remarkable drift when we have tested the sensor with differ-
­(Zni) act as donor defects, this explain the origin of n-type ent CO concentrations. The response was enhanced when the
conduction in undoped ZnO. Thereafter, the low activation operating temperature increased to 250 °C and reached 3.8
energy is consistent with the presence of donor defects (Oi, and the drift disappeared. When the temperature increases to
­Zni and ­Vo) in the forbidden band as deduced from PL meas- 300–350 °C, the sensor response gets decreases. The sens-
urement (Fig. 5). ing mechanism can be explained as follow: Stable oxygen
ions species were O ­ 2− below 200 °C, O
­ − between 200 and
2−
3.2.1 Response Towards CO Gas 300 °C, and O ­ above 300 °C [45]. The reactions of the
oxygen species with CO molecules at different operating
The variation of ZnO sensor response versus temperature temperatures can be described using the following equations:
towards 80 ppm of carbon monoxide is shown in Fig. 8.
2CO + O−2 → 2CO2 + e− (2)
The response exhibits a Gaussian shape. Starting from
150 °C, the response increases and goes through a maxi-
mum value located at 250 °C and then decreases. The CO + O− → CO2 + e− (3)
CO sensor response depends on an accurate equilibrium
between adsorption and desorption rate of carbon monox- CO + O2− → CO2 + 2e− (4)
ide. It depends also on the reactivity of the CO surface with
adsorbed oxygen. Enhancing temperature to some value At temperature below 200 °C, Eq. 2 plays major role. In
leads to CO chemisorption and reaction rate occurring on the this case CO gas reacts slowly with O ­ 2− species leading to
ZnO surface which favors increase of the gas response. Rais- low response of the gas sensor. Going from 250 to 300 °C,
ing temperature, we showed a reduction on the gas response the chemisorbed oxygen species present at the surface of
because of CO desorption; and this produces a decrease of ZnO are ­O− and ­O2−. This means that both Eqs. (3) and (4)
the amount of carbon monoxide adsorbed on the surface of participated in CO adsorption operation causing an increase
ZnO. of gas response. The previous mechanism can also explained
The dynamic responses of ZnO based sensor tested to in Fig. 10. Above 300 °C, only Eq. (3) appears in CO adsorp-
different carbon monoxide concentrations from 5 to 80 ppm tion; and this explains the decrease of the response.
in air at different operating temperatures (200–350 °C) are The response/recovery times is defined as the time which
reported in Fig. 9. The injection of CO gas decreases the occurs to reach 90% of the final resistance after the exposure
electrical resistance of the layer. The signal could be back to target and reference air, respectively [46, 47].
to its primary value after many cycles indicating that CO The response and recovery time observed for ZnO sam-
adsorption was reversible. The CO gas was then desorbed ple with different operating temperature were 17–63 s and
when the gas was turn off. 30–102 s respectively as seen in Fig. 11. Faster response
The operating temperature is an important parameter time is observed for the operating temperature of 300 °C
that influences the sensor response; for 200 °C there was a and faster recovery time is observed at 350 °C. Lower
response and recovery time, respectively 63 s and 102 s,
4
were observed for the operating temperature of 200 °C. This
is due to CO adsorption and desorption in the surface of the
material.
Sensor response (R0/R)

3 ZnO
80 ppm CO 3.2.2 Selectivity of ZnO Sensor

The fabricated device was tested to different gases such as


2 CO, ­NO2 and ­CO2 at 250 °C in order to show the selectivity
of the sensor. As seen in Fig. 12, the sensor device is very
sensitive to CO gas. The response to oxidizing gases (such
as ­NO2) appears instead to be more limited. The selectivity
1
of our sensor towards CO gas id attributed to many factors
150 200 250 300 350 such as metal oxide shape and size. The operating tempera-
Temperature (°C) ture [48] is a key parameter for the selectivity of the gas
sensor. CO molecules are very reactive from 250 to 350 °C.
Fig. 8  Response to 80 ppm CO of the ZnO nanoparticles as a func- As result the sensor is selective to CO gas compared to C
­ O2
tion of the temperature and ­NO2.

13
4068 Journal of Inorganic and Organometallic Polymers and Materials (2020) 30:4063–4071

60000

5000

40000 4000
Resistance (KΩ )

Resistance (KΩ )
5 ppm

3000 10 ppm

20000
Pure ZnO 20 ppm
Pure ZnO
2000 CO
CO
250°C
200°C
50 ppm
0 1000
80 ppm
28000 29000 30000 26000 27000 28000
Time (s) Time (s)

500

70
Resistance (KΩ )

5 ppm 5 ppm
10 ppm
400 20 ppm 10 ppm
Resistance (KΩ )

Pure ZnO Pure ZnO (400°C/2h) 20 ppm


50 ppm CO (5-80 ppm)
CO (5-80 ppm) 50 ppm
300°C 65 350°C
50 ppm 50 ppm
80 ppm
80 ppm
300
10000 11000 12000 13000
7000 8000 9000 10000
Time (s)
Time (s)

Fig. 9  Dynamic response of pure ZnO at different operating temperature to various concentrations of CO in air

110

100
60
90

80
Response time (s)

70 Recovery time (s)


40
60

50

40
20
30
Fig. 10  CO sensing mechanism of ZnO gas sensor
20
180 200 220 240 260 280 300 320 340 360

Temperature (°C)
In order to show the stability of the sample, ZnO was
tested towards 80 ppm of CO at 250 °C several times along
Fig. 11  Variation of response/recovery times versus operating tem-
1 months as seen in Fig. 13. We noticed that the sensor perature
response decreases from 74 to 71%. This is probably due to
the surface reaction with its surroundings along the meas-
urements. The sample exhibited a high durability of around
95%.

13
Journal of Inorganic and Organometallic Polymers and Materials (2020) 30:4063–4071 4069

5000
4000

Pure ZnO 4000


Resistance (KΩ )

Resistance (KΩ )
NO2 5 ppm
3500
250°C
1 ppm
3000 10 ppm

3000 0.25 ppm


Pure ZnO 20 ppm
5 ppm 2000 CO
250°C
0.5 ppm
50 ppm
2500 1000
80 ppm
3000 4000 5000 6000 26000 27000 28000
Time (s) Time (s)
0,66

ZnO c) 70
5%
60
1%
Resistance (MΩ)

Response (%)
50
0.5%
0,64 0.25% 40

30

20

10
CO2 T=250 °C
0,62 0
9200 9400 9600 9800
CO (80ppm) CO2 (500 ppm) NO2 (5ppm)

Time (s)

Fig. 12  Response to different gases

6000000
80 ppm CO
72 ZnO
5000000
Sensor response (%)

69
Resistance (ΚΩ )

4000000

66 80 ppm CO
3000000

63
2000000

60
0 5 10 15 20 25 30 1000000
27400 27600 27800 28000 28200 28400
Days
Time (s)

Fig. 13  Stability of the sensor with days towards 80 ppm of CO gas


Fig. 14  Reproducibility tests to 80 ppm of CO at 250 °C

Figure 14 presents the reproducibility of the sensor when


exposed to three consecutive pulses of 80 ppm of CO gas at As shown in Table 1, the developed sensor based on
250 °C. It is clearly observed that the response and recovery ZnO nanopowders shows a very high response and quick
characteristics are almost reproducible. response/recovery times compared to a previous works

13
4070 Journal of Inorganic and Organometallic Polymers and Materials (2020) 30:4063–4071

Table 1  Comparison of Sample Response (%) CO (ppm) Tempera- Response Recovery Refs.
different CO sensing properties ture (°C) time (s) time (s)
with other works
ZnO 74 80 250 21 70 This work
Sm2O3 34 50 250 35 110 [49]
TiO2/LSCN 38.4 400 200 – – [50]
LaCo1−xFexO3 40.9 100 250 187 100 [51]
CuO 47.2 800 150 40 90 [52]

[49–52]. A response, (ΔR/R0) % of 74 at 80 ppm, is 8. K. Omri, A. Bettaibi, K. Khirouni, L. El Mir, Phys. B 537,
obtained which indicates the high sensor response of our 167–175 (2018)
9. K. Omri, A. Alyamani, L. El Mir, J. Mater. Sci. 30, 16606–
device as CO gas sensor. 16612 (2019)
10. K. Omri, I. Najeh, L. El Mir, Ceram. Int. 42, 8940–8948 (2016)
11. R.N. Viswanath, S. Ramasamy, R. Ramamoorthy, P. Jayavel, T.
Nagarajan, Nanostruct. Mater. 6, 993–996 (1995)
4 Conclusion 12. M.S. Wu, A. Azuma, T. Shiosaki, A. Kawabata, I.E.E.E. Trans,
Ultrason. Ferroelec. Freq. Contr. 136, 442–445 (1989)
Pure ZnO nanoparticles have been prepared by sol–gel 13. M.K. Jayaraj, A. Antony, M. Ramachandran, Bull. Mater. Sci.
25, 227–230 (2002)
technique and heat treated at 400 °C for 2 h in air. ZnO 14. K. Keis, C. Bauer, G. Boschloo, A. Hagfeldt, K. Westermark, H.
nanoparticles have a hexagonal wurtzite structure with an Rensmo, H. Siegbahn, J. Photochem. Photobiol. A 148, 57–64
average crystallite size of 50 nm. TEM showed spherical (2002)
and agglomerated particles with a size of about 50 nm. PL 15. P. Yang, H. Yan, S. Mao, R. Russo, J. Johnson, R. Saykally,
N. Morris, J. Pham, R. He, H. Choi, Adv. Funct. Mater. 12,
measurements showed two peaks are observed in the photo- 323–331 (2002)
luminescence spectrum. The first peak is located at 380 nm 16. S. Roy, S. Basu, Bull. Mater. Sci. 25, 513–515 (2002)
and it corresponds to the near band edge (NBE) peak and 17. A.R. Raju, C.N.R. Rao, Sens. Actuators B 3, 305–310 (1991)
the second peak is centered at around 554 nm. It is well 18. L. El Mir, Z. Ben Ayadi, M. Saadoun, H.J. Von Bardeleben,
K. Djessas, A. Zeinert, Phys. Status Solidi (a) 204, 3266–3277
known that the green emission originated from the electronic (2007)
defects in the ZnO forbidden band. The realized gas sensor 19. T. Yamazaki, S. Wada, T. Noma, T. Suzuki, Sens. Actuators B
exhibited a high response of 74% towards 80 ppm of CO 13–14, 594–595 (1993)
gas at 250 °C. The response/recovery times are 21 and 70 s, 20. M. Hjiri, L. El Mir, S.G. Leonardi, A. Pistone, L. Mavilia, G.
Neri, Sens. Actuators B 196, 413–420 (2014)
respectively. The fabricated sensor has been tested to C ­ O2 21. E. Pál, V. Hornok, R. Kun, A. Oszkó, T. Seemann, I. Dékány,
and ­NO2 gases. The results reveal its high response towards M. Busse, J. Colloid Interface Sci. 378, 100–109 (2012)
CO in comparison to the other studied gases. In addition our 22. S. Nundy, T. Eom, J. Kang, J. Suh, M. Cho, J.S. Park, H.J. Lee,
sensor exhibited good reproducibility and long stability. The Ceram. Int. 46, 5706–5714 (2020)
23. S.K. Shaikh, V.V. Ganbavale, S.V. Mohite, U.M. Patil, K.Y.
good sensing properties of ZnO nanoparticles towards CO Rajpure, Superlattices Microstruct. 120, 170–186 (2018)
gas suggests its potential application for the detection of this 24. S. Kanaparthi, S.G. Singh, Mater. Sci. Energy Technol. 3, 91–96
hazardous gas. (2020)
25. J.H. Kim, A. Mirzaei, H.W. Kim, P. Wu, S.S. Kim, Sens. Actua-
tors B 293, 210–223 (2019)
26. J. Singh, S. Sharma, S. Soni, S. Sharma, R. Chand Singh, Mater.
References Sci. Semicond. Process. 98, 29–38 (2019)
27. P. Shunmuga Sundaram, S. Stephen Rajkumar Inbanathan, G.
Arivazhagan, Phys. B 574, 411668 (2019)
1. Z.D. Lin, W.L. Song, H.M. Yang, Sens. Actuators B 173, 22–27 28. M. Hjiri, R. Dhahri, L. El Mir, A. Bonavita, N. Donato, S.G.
(2012) Leonardi, G. Neri, J. Alloys, Compounds 634, 187–192 (2015)
2. P. Sun, W.N. Wang, Y.P. Liu, Y.F. Sun, J. Ma, G.Y. Lu, Sens. 29. J.L. Yu, Y.F. Lai, S.Y. Cheng, Q. Zheng, Y.H. Chen, J. Lumin.
Actuators B 173, 52–57 (2012) 161, 330–334 (2015)
3. H. Kim, C. Jin, S. Park, S. Kim, C. Lee, Sens. Actuators B 161, 30. J. Lv, M. Fang, Mater. Lett. 218, 18–21 (2018)
594–599 (2012) 31. S. Pal, A. Sarkar, P. Kumar, D. Kanjilal, T. Rakshit, S.K. Ray,
4. P.M. Perillo, D.F. Rodríguez, Sens. Actuators B 171–172, 639– D. Jana, J. Lumin. 169, 326–333 (2016)
643 (2012) 32. J.D. McNamara, N.M. Albarakati, M.A. Reshchikov, J. Lumin.
5. C. Baban, Y. Toyoda, M. Ogita, J. Optoelectron. Adv. Mater. 7, 178, 301–306 (2016)
891–896 (2005) 33. J. Wang, S. Yu, H. Zhang, Optik 180, 20–26 (2019)
6. J.J. Hassana, M.A. Mahdia, C.W. China, H. Abu-Hassan, Z. Has- 34. C. Wende, W. Ping, Z. Xingquan, X. Tan, J. Appl. Phys. 100, 5
san, Sens. Actuators B 176, 360–367 (2013) (2006)
7. P. Singh, V.N. Singh, K. Jain, T.D. Senguttuvan, Sens. Actuators
B 166–167, 678–684 (2012)

13
Journal of Inorganic and Organometallic Polymers and Materials (2020) 30:4063–4071 4071

35. T.M. Børsetha, B.G. Svensson, A.Y. Kuznetsov, Appl. Phys. Lett. 46. B.T. Raut, P.R. Godse, S.G. Pawar, M.A. Chougule, B. Patil, J.
89, 26 (2006) Mater. Sci. 23, 956–963 (2012)
36. K. Ravichandran, A. Anbazhagan, M. Baneto, N. Dineshbabu, C. 47. L.E. Mathevulaa, L.L. Notoa, B.M. Mothudia, M. Chithambo,
Ravidhas, G. Muruganandam, Mater. Sci. Semicond. Process. 41, M.S. Dhlamini, J. Lumines. 192, 879–887 (2017)
150–154 (2016) 48. M. Hjiri, L. El Mir, S.G. Leonardi, N. Donato, G. Neri, Nanoma-
37. P. Murkute, H. Ghadi, S. Saha, S.K. Pandey, S. Chakrabarti, terials 3, 357–369 (2013)
Mater. Sci. Semicond. Process. 66, 1–8 (2017) 49. S. Rasouli Jamnani, H. Milani Moghaddam, S.G. Leonardi, N.
38. P.U. Londhe, N.B. Chaure, Mater. Sci. Semicond. Process. 60, Donato, G. Neri, Appl. Surf. Sci. 487, 793–900 (2019)
5–15 (2017) 50. K.-C. Hsu, T.-H. Fang, Y.-J. Hsiao, P.-C. Wu, J. Alloy. Compd.
39. G. Neri, A. Bonavita, G. Micali, G. Rizzo, E. Callone, G. Car- 794, 576–584 (2019)
turan, Sens. Actuators B 132, 224–233 (2008) 51. J.-C. Ding, H.-Y. Li, T.-C. Cao, Z.-X. Cai, X.-X. Wang, X. Guo,
40. L. El Mir, J. El Ghoul, S. Alaya, M. Ben Salem, C. Barthou, H.J. Solid State Ionics 303, 97–102 (2017)
von Bardeleben, Phys. B 403, 1770–1774 (2008) 52. R. Molavi, M.H. Sheikhi, Mater. Sci. Semicond. Proc. 106,
41. Y. Chen, D.M. Bagnall, H.K. Koh, K.T. Park, K. Hiraga, Z.Q. 104767 (2020)
Zhu, T. Yao, J. Appl. Phys. 84, 3912 (1998)
42. Y. Chem, D.M. Bagnall, H.K. Koh, K.T. Park, K. Hiraga, Z.Q. Publisher’s Note Springer Nature remains neutral with regard to
Zhu, T. Yao, J. Appl. Phys. 84, 3912 (1998) jurisdictional claims in published maps and institutional affiliations.
43. V.A. Fonoberov, K.A. Alim, A.A. Balandin, F. Xiu, J. Liu, Phys.
Rev. B 73, 165317–165326 (2006)
44. D.R. Lide, Handbook of Chemistry and Physics (CRC Press,
Florida, 1992), pp. 4–163
45. H. Gong, J.Q. Hu, J.H. Wang, C.H. Ong, F.R. Zhu, Sens. Actua-
tors B 115, 247–251 (2006)

13

You might also like