Sie sind auf Seite 1von 22

Accepted Manuscript

Title: Highly sensitive acetone sensors based on


flame-spray-made La2 O3 -doped SnO2 nanoparticulate thick
films

Authors: N. Tammanoon, A. Wisitsoraat, D. Phokharatkul, A.


Tuantranont, S. Phanichphant, V. Yordsri, C. Liewhiran

PII: S0925-4005(18)30267-3
DOI: https://doi.org/10.1016/j.snb.2018.01.238
Reference: SNB 24087

To appear in: Sensors and Actuators B

Received date: 13-10-2017


Revised date: 17-1-2018
Accepted date: 31-1-2018

Please cite this article as: { https://doi.org/

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
1
 

<AT>Highly sensitive acetone sensors based on flame-spray-made La2O3-doped SnO2


nanoparticulate thick films
<AU>N. Tammanoona,b, A. Wisitsoraatc,d,e, D. Phokharatkulc,d, A. Tuantranontc,f, S.
Phanichphantc, V. Yordsrig, C. Liewhirana,c,h,* ##Email##cliewhiran@gmail.com##/Email##
<AFF>aDepartment of Physics and Materials Science, Faculty of Science, Chiang Mai
University, Chiang Mai 50200, Thailand
<AFF>bGraduate School, Chiang Mai University, Chiang Mai 50200, Thailand
<AFF>cCenter of Advanced Materials for Printed Electronics and Sensors, Materials Science

PT
Research Center, Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand
<AFF>dCarbon-based Devices and Nanoelectronics Laboratory, National Electronics and
Computer Technology Center, National Science and Technology Development Agency,
Pathumthani 12120, Thailand

RI
<AFF>eDepartment of Common and Graduate Studies, Sirindhorn International Institute of
Technology, Thammasat University, Pathumthani 12120, Thailand

SC
<AFF>fThailand Organic and Printed Electronics Innovation Center, National Electronics and
Computer Technology Center, National Science and Technology Development Agency,
Pathumthani 12120, Thailand
<AFF>gNational Metal and Materials Technology Center, National Science and Technology

U
Development Agency, Klong Luang, Pathumthani 12120, Thailand
<AFF>hCenter of Excellence in Materials Science and Technology, Chiang Mai University,
Chiang Mai 50200, Thailand N
<AFF>Tel.: +66-81-408-2324; Fax: +66-53-943-445
A
<PA>*,a
Corresponding author.
M

<ABS-Head><ABS-HEAD>Graphical abstract
<ABS-P>
D

<ABS-P><xps:span class="xps_Image">fx1</xps:span>
TE

<ABS-HEAD>Highlights► Flame-made SnO2 nanoparticles were incorporated with 0–2 wt%


La. ► Structural analyses suggested that La2O3 crystallites were soluted in SnO2 matrix. ►
Response to 400 ppm acetone at 350°C was greatly enhanced from 8.2 to 3,626 with 0.5 wt% La.
EP

► The optimal sensor exhibited high acetone selectivity against SO2, H2S, NO2, C6H6, C7H8,
C8H10 and CH2O. ► The results were attributed to distributed p-n heterjunctions, strong La2O3
basicity and reduced structural sizes.
CC

<ABS-HEAD>ABSTRACT
<ABS-P>In the present work, flame-spray-made La2O3-doped SnO2 nanoparticles with 0.1−2
wt% La contents were systematically studied for acetone detection. The particle and sensing film
properties were characterized by X-ray diffraction, nitrogen adsorption, electron microscopy,
A

energy dispersive X-ray spectroscopy and X-ray photoelectron spectroscopy. The sensing films
were tested towards 0.1−400 ppm acetone at operating temperatures ranging from 150 to 400°C
in dry air. Gas-sensing results demonstrated that the SnO2 sensing film with the optimal La
content of 0.5 wt% exhibited a very high response of 3,626 toward 400 ppm acetone with a short
response time of 2.8 s at the optimal operating temperature of 350°C. Moreover, the sensors
displayed high acetone selectivity against SO2, H2S, NO2, C6H6, C7H8, C8H10 and CH2O.
Therefore, the La2O3-doped SnO2 sensors are promising for sensitive and selective detections of
acetone at low concentrations.
2
 

<KWD>Keywords: Flame spray pyrolysis; Acetone; SnO2; La2O3 doping; Gas sensor.

Introduction
Acetone (C3H6O) is a highly volatile organic compound (VOC) commonly used as a solvent and
a chemical intermediate in many chemical laboratories and industries. Its ingestion and
inhalation at significant levels can bring about low acute and chronic poisoning [1]. It is thus one
of the most important air pollutants that should be monitored and controlled. The acetone
concentration in polluted air is typically in the range of 1–200 ppm [2]. In addition, it is naturally

PT
produced from endogenous lipolysis in human and is present in blood, urine as well as breathe.
The concentration of acetone in breathe can be used to indicate the ketotic state of diabetes
relating to the levels of glucose and fat loss. Insulin-dependent diabetic patients having ketosis
(high ketone level in blood) will release high acetone concentrations (>1.8 ppm) in breathe,

RI
which are more than twice of that of normal people (<0.9 ppm) [3]. Therefore, techniques for
acetone detection in breathe must be highly sensitive and selective. Presently, breath analysis

SC
requires analytical instruments such as gas chromatography and spectroscopic techniques, which
are expensive, cumbersome and only available in laboratory. It is thus compelling to develop
highly sensitive and selective acetone sensors for the rapid determination of air-quality as well as
diabetes and related illnesses. Acetone gas sensors based on nanostructures of semiconducting

U
metal oxides particularly tin oxide (SnO2) are among promising candidates for this purpose due
to its low cost, small size, high sensitivity, fast response, good stability and simple electronic
N
interface. However, they still suffer from low sensitivity and selectivity towards acetone [3, 4].
Two practical approaches to minimize these short comings include the reduction of structural
A
sizes and doping or loading with effective additives. The structural sizes can be reduced by the
selection of an appropriate synthesis process to obtain higher specific surface area, larger number
M

of gas adsorptive sites and consequently higher sensitivity while the additive may be chosen
from a wide variety of elements including noble metals, transition metals and post-transition
metals as well as rare-earth elements based on its electronic and catalytic properties towards a
D

particular gas. Some dopants including La, Ni, Y and Eu have recently been reported to
substantially enhance the gas-sensing performances of SnO2 towards VOCs especially acetone
TE

[5–9]. Among these, La (the first element of lanthanide transition metal group) is a particularly
attractive choice due to its high chemical reactivity, unique catalytic activity, strong surface
basicity in oxide form (La2O3) and moderate cost [5–7]. La has only one stable oxidation state of
EP

La3+ and normally presents in oxide or other compound forms due to its high reactivity. In
addition, La2O3 is a p-type semiconducting oxide with interesting gas-sensing properties [7].
Gas-sensing performances of SnO2 sensors with various additives and other advanced materials
CC

prepared by different production methods towards VOCs particularly acetone are summarized in
Table 1. Firstly, La-SnO2 nanobelts prepared by thermal evaporation exhibited optimal sensor
responses at 230°C to 100 ppm acetone, 100 ppm ethanol and 100 ppm ethanediol of 2.28, 3.75
and 8.76, respectively [6]. In addition, the 5 at% La-SnO2 pellet sensor synthesized by a sol–gel
A

technique displayed good responses of 59, 57 and 55 to 500 ppm acetone, 500 ppm ethanol and
500 ppm methanol, respectively (at 300°C) [7]. Also, a sol-gel derived 4 at% La-SnO2 thin film
revealed good responses of ~3.3 and 2.1 to 10 and 1 ppm ethanol, respectively (at 350°C) [4].
Moreover, 3 mol% La-doped SnO2 nanoparticles prepared by co-precipitation gave a fair
response of 3.3 to 100 ppm acetone at 300°C with a short response time and good selectivity
against ethanol, ammonia and liquid petroleum gas (LPG) [8]. Furthermore, 5 wt% La-doped
SnO2 fabricated by a hydrothermal process exhibited a high response of 213 and 898 to100 and
500 ppm ethanol, respectively (at 300°C) [5]. In another work, a Y-doped SnO2 nanostructures
3
 

prepared by a thermal evaporation process showed a good responses of 15.8 to 400 ppm acetone
at 210°C [9]. Additionally, electrospun Eu-doped SnO2 nanofibers displayed a high response of
127 to 300 ppm acetone at 280°C [10]. Similarly, graphene loaded Ni-doped SnO2 nanoparticles
synthesized by flame spray pyrolysis (FSP) and electrolytic exfoliation offered a high response
of 169.7 to 200 ppm acetone at 350°C [11]. Regarding other advanced materials, Cu/Zn oxide
multilayer structures produced by sputtering/oxidation processes demonstrated an optimal
response of 34% to 100 ppm ethanol at 360°C [12]. Similarly, porous Cu, Cd, Zn and Ni Ferrites
prepared by a sol-gel selfcombustion technique displayed quite similar responses in the range of

PT
13-46 % to 150 ppm acetone and 150 ppm ethanol at optimal working temperatures of 300, 350,
350 and 250°C [13,14]. Likewise, sol-gel-prepared Ni1-x-Cox Ferrites with x =0.25 yielded
relatively high responses of 61 and 56 % to 150 ppm acetone and 150 ppm ethanol at an optimal
working temperature of 250°C [15]. Besides, Yb Ferrite nanoparticles synthesized by a sol-gel

RI
method gave a decent response of 9.5 to 3 ppm acetone at 230°C [16]. Moreover, ZnO–Fe3O4
with inverse opal structures fabricated by the sol-gel technique with opal templates provided a

SC
response of 2 to a very low acetone concentration of 0.8 ppm at 485°C [17]. Lastly, polyaniline
(Pani)-Mg Ferrite composites offered responses of 45 and 31 % to 150 ppm acetone and ethanol
at a relatively low operating temperature of 200°C [18].
It can be seen that the gas response and selectivity of SnO2 based sensors are still relatively high

U
compared with some other advanced materials and are dependent considerably on doping as well
as production method. In particular, La doping enhances sensitivity and selectivity towards either
N
acetone or ethanol depending on the sensor’s preparation condition. It is thus interesting to
explore the gas-sensing behaviors of La-doped SnO2 materials synthesized by another innovative
A
method. FSP is an appealing nanopowder production method that can produce metal-
doped/loaded metal oxide nanoparticles with large specific surface area in a single high-
M

temperature step. It has successfully been applied to produce a variety of unloaded and metal-
loaded/doped metal oxide nanoparticles that exhibit high gas-sensing performances [11].
However, there is still no report of FSP-made La2O3-doped SnO2 nanoparticles for gas-sensing
D

applications. In this work, La2O3-doped SnO2 nanoparticles are prepared by FSP with varying
doping concentrations for the first time and systematically characterized for gas-sensing
TE

performances in air quality monitoring applications towards acetone, sulfur dioxide (SO2),
hydrogen sulfide (H2S), nitrogen dioxide (NO2), benzene (C6H6), toluene (C7H8), xylene (C8H10)
and formaldehyde (CH2O).
EP

<H1>1. Experimental
<H2>1.1. Synthesis of La-doped SnO2 nanoparticles
Undoped SnO2 and La2O3-doped SnO2 nanoparticles with 0.1–2 wt% La were produced by the
CC

FSP method whose details of experimental setup were reported previously [11]. The precursor
solution for the synthesis was prepared by dissolving appropriate amounts (0.50 M) of tin (II) 2-
ethylhexanoate (Aldrich, 95%) and lanthanum (III) acetylacetonate (Aldrich, 95%) in ethanol
(Aldrich, 99%). The precursor mixture was delivered into a nozzle at a constant feeding rate of
A

5 mL/min through the inner reactor capillary via a syringe pump and dispersed in a flame
supported by the methane/oxygen gas mixture under the 5/5 condition. Precursor droplets were
vaporized and combusted to form nanoparticles via nucleation, condensation, coagulation,
coalescence and La2O3-doping on SnO2 nanoparticles. It was observed that the flame height was
∼8–10 cm and the flame color turned form yellowish to orange, implying the decrease of flame
temperature with increasing La precursor concentration. The flame-made undoped
SnO2 nanopowder was labeled as P-0 whereas SnO2 nanoparticles doped with 0.1, 0.2, 0.5, 1 and
2 wt% La were indicated as P-0.1La, P-0.2La, P-0.5La, P-1La and P-2La, respectively.
4
 

<H2>1.2. Preparation of sensing films


Sensing films were fabricated from synthesized powders by spin-coating of powder paste. The
flame-made nanoparticles (60 mg) were mixed into an organic paste comprising ethyl cellulose
(Fluka, 30–70 mPas) and -terpineol (Aldrich, 90%) as a binder and a solvent, respectively.
They were crushed in a mortar for 30 min to form a homogeneous paste, which was then spin-
coated on Al2O3 substrates (0.40 cm × 0.55 cm × 0.04 cm) equipped with interdigitated Au
electrodes. The resulting sensing films were annealed in an oven at 150°C for 1 h and at 450°C
for 2 h with a heating rate of 1°C/min for binder removal prior to sensing test. The sensors

PT
fabricated from P-0 and P-0.1La to P-2La powder samples were designated as S-0 and S-0.1La
to S-2La, respectively.
<H2>1.3. Structural characterizations
X-ray diffraction (XRD) patterns of nanoparticles and sensing films were measured to analyze

RI
phase composition by a commercial X-ray diffractometer (Rigaku, TTRAXIII) using CuKα
radiation detected at a scanning speed of 3°/min. The structural morphologies of nanoparticles

SC
and sensing films were examined by high-resolution transmission electron microscopy (HR-
TEM, TEM: JSM-2100, JEOL) and scanning electron microscopy (SEM: JSM-6335F, JEOL).
The elemental composition was evaluated by energy-dispersive X-ray spectroscopy (EDX) in
SEM and scanning transmission electron microscopy (STEM) modes. The specific surface area

U
of nanoparticles was determined by nitrogen absorption with Brunauer–Emmett–Teller (BET)
analysis (Micromeritics Tristar 3000) after degassing sample at 120°C for 2 h. Moreover, the
N
surface chemical compositions and oxidation states of nanoparticles as well as sensing films
were analyzed by X-ray photoelectron spectroscopy (XPS: AXIS UltraDLD, Kratos analytical,
A
UK).
<H2>1.4. Gas-sensing measurement
M

The gas-sensing performances of sensing films were characterized towards C3H6O, SO2, H2S,
NO2, C6H6, C7H8, C8H10 and CH2O at operating temperatures ranging from 150 to 400°C under
atmospheric conditions by the standard flow through technique in a sealed stainless steel
D

chamber. A flux of synthetic dry air was flowed to mix at a standard T-junction with the desired
concentration of pollutants dispersed in synthetic air at a constant total flow rate of 2 L/min. The
TE

gas flow mixing rates were precisely manipulated using computer controlled multi-channel mass
flow controllers )Brook Instruments model 5850E). These gases were selected for testing since
they are important indoor and outdoor air pollutants and the tested acetone concentration was in
EP

the range of 0.1–400 ppm covering the nominal application range in ambient (0.9–200 ppm).
Due to the limitation of mass flow control range, the acetone sensing measurements were divided
into high and low concentration ranges of 6–400 and 0.1–4 ppm, respectively. In the high
CC

concentration range, acetone and other VOCs including C6H6, C7H8, C8H10 and CH2O were
tested by bubbling through their pure liquid sources with dry air and their concentrations were
calibrated by commercial gas detectors (E Instruments International, LLC) while calibrated
standard gases were used for the test of C3H6O at low concentrations and other gases including
A

H2S, NO2 and SO2. The sensing films were heated to different operating temperatures using an
external NiCr heater with a regulated DC power managed by a proportional-integral temperature
controller. The sensor currents at an applied voltage of 10 V were continuously monitored to
obtain sensor response curves. The sensors were exposed to a gas sample for ~10 min at each gas
concentration and the synthetic dry air was then resumed for 25 min. Additional current-voltage
measurements were also carried out before and after acetone exposure over the voltage range of -
100100 V with a step of 2 V and a step time of 2 min. The sensor response (S) is defined as the
resistance ratio, Ra/Rg [4], where Ra is the baseline of resistance in dry air and Rg is the stabilized
5
 

change in resistance in the reducing gas including C3H6O, H2S, SO2, C6H6, C7H8, C8H10 and
CH2O. The response definition is reversed for oxidizing gas such as NO2. The response time (tres)
is defined as the time required until 90% of the response signal is reached while the recovery
time (trec) denotes the time needed until 90% of the original baseline signal is recovered.
<H1>2. Results and discussion
<H2>2.1. Structural properties of FSP made powders and sensing films
XRD patterns of flame-spray-made undoped SnO2 (P-0) and 0.1−2 wt% La-doped SnO2 (P-
0.1La to P-2La) nanopowders are shown Fig. 1(a). It is apparent that XRD peaks of all samples

PT
are sharp but exhibit some broadening, indicating polycrystalline structure with nanometer-scale
crystallite sizes. In addition, all peak positions are fitted well with those of the tetragonal SnO2
cassiterite-tetragonal phase (JCPDS file no. 41-1445), displaying (1 1 0), (1 0 1), (2 0 0), (2 1 1),
and (2 2 0) dominant planes. Moreover, peaks of secondary La or La2O3 phases cannot be

RI
detected even at the highest La content due to the formation of solid solution of La2O3 in SnO2
matrix or very low diffraction intensity of La2O3 species. Nevertheless, the La content seems to

SC
have some influence on peak broadening and the crystallite size. This effect can be clearly seen
from the average crystallite size of SnO2 nanoparticles estimated from Scherrer's equation using
the data of the first three dominant planes as shown in the inset of Fig. 1 (a). As the La content
increases from 0 to 0.5 wt%, the crystallite size of SnO2 decreases considerably from 7.5 nm to

U
5.3 nm but then slightly increases to 6.5 nm as the La content increases further to 2.0 wt%. The
results indicate that La dopants inhibit grain growth of SnO2nanocrystal but the grain growth
N
inhibition effect reduces at higher La doping levels. La dopants may limit grain growth via
solute-induced drag effect [19], which may be weakened at high La contents due to the
A
agglomeration of dopant species. Fig. 1(b) shows the XRD patterns of sensing films made of
SnO2 doped with 0−2 wt % La (S-0 to S-2La) on the Au/Al2O3 substrate after annealing and
M

sensing test. It confirms that the sensing films have similar crystallinity to their corresponding
powders displaying only the peaks of the SnO2 cassiterite phase, which are relatively broad and
low in intensity compared with those of Al2O3 and Au phases from the substrates.
D

The presence of La dopants in the sensing film (S-2La) was firstly verified by EDX together with
the top-view surface morphology by SEM as illustrated in Fig. 2. It demonstrates round
TE

nanoparticles with fine diameters (< 50 nm) distributed quite uniformly on the film. The
corresponding EDX maps of Sn, La and O elements also confirm uniform distribution of these
species over the imaged area. Additionally, the density of La is correspondingly lower than those
EP

of Sn and O in accordance with the relative composition values. Nevertheless, the mapping
resolution of the employed SEM instrument is insufficient to indicate if La species are decorated
on or distributed in the particles. The respective full-area EDX spectrum affirms the presence of
CC

all expected elements of sensing films with insignificant contributions of other elements from
contaminations and substrate materials, implying large film thickness and high film purity. The
Sn, La and O compositions are calculated from the spectrum to be 77.79 wt% (33.87 at%), 1.98
wt% (0.74 at%) and 20.23 wt% (65.39 at%), respectively, which are in good agreement with the
A

intended La content (2 wt%) and the stoichiometric values of Sn (31 at%) and O (67 at%).
Fig. 3 demonstrates typical bright-field (BF) and corresponding high-resolution (HR)-

TEM images of FSP-made undoped SnO2 (a, b), 0.5 wt% LaSnO2 (c, d) and 2 wt% LaSnO2 (e,

f). From the BF-TEM images (Fig.3 (a, c, e)), undoped and La2O3-doped SnO2 nanoparticles are

mainly spheroidal and occasionally polyhedral with varying diameters in the range of 5–20 nm.
6
 

In addition, the particle sizes of undoped SnO2 are observed to be notably larger than those of

La-doped ones. The corresponding SAED patterns (inset) indicate that they are mainly

polycrystalline with primary diffraction rings that can be assigned to (1 1 0), (1 0 1), (2 0 0), (2 1

1) and (1 1 2) planes of the tetragonal-cassiterite SnO2 phase in agreement with the observed

PT
XRD patterns (Fig. 1). The related HR-TEM images (Fig. 3 (b, d, f)) display lattice fringes on

various nanoparticles whose d-spacings correspond to diffracted planes of the tetragonal SnO2

RI
phase. Moreover, the sizes of SnO2 nanoparticles incline to decrease while the secondary phases

SC
of La or La2O3 cannot be observed, suggesting a solid solution of La species in SnO2 lattice. The

experimental results agree well with other studies of La-incorporated SnO2 materials [5–9].

U
According to the Hume-Rothery rule, La3+ ions cannot substitute Sn4+ ions in lattice because the

N
difference between the effective ionic radii of La3+ and Sn4+ (0.103 and 0.069 nm) is much larger
A
than the limit of 15%. Thus, the obtained structure cannot be the La-substituted SnO2 solid
M

solution but the solid solution of La2O3 crystallites within SnO2 nanocrystalline matrix. As a

result, these materials are identified as La2O3-doped SnO2. However, the La2O3 crystallites
D

cannot still be observed because they may be very small and exhibit too low phase contrast to be
TE

resolved by the non-aberration-corrected TEM instrument.


EP

The detailed distribution of La species on individual SnO2 nanoparticles (P-2La) was further
assessed by EDX mapping in STEM mode as illustrated in Fig. 4. The STEM image displays
group of round and oval nanoparticles with 7–15 nm in diameters loosely aggregated on the C-
CC

coated Cu grid. It should be noted that the electron beam is scanned in the STEM mode, resulting
in a reduced resolution. Thus, the STEM image cannot be as sharp as those of the high-resolution
TEM mode (previously illustrated in Fig. 3). The STEM mode is necessary for the high-
resolution EDS mapping, which gives much more precise distributions of La species on SnO2
A

nanoparticles than the EDS mapping based on scanning electron microscopy (SEM) presented
earlier in Fig. 2. The respective EDX maps of Sn, La and O elements in Fig. 4 show the details of
their distribution over the scanned area. It is seen that Sn, O and La species are quite evenly
dispersed only on the regions containing nanoparticles. Moreover, the high-resolution EDX maps
reveal that La species are distributed within individual particles with a relatively low density
compared with Sn and O ones in accordance with its low content of 2 wt%. It can be noticed that
some O signals are additionally found in the void regions (the white areas in the STEM image),
which should arise from some oxygen species adsorbed on the C/Cu grid. The corresponding
7
 

full-area EDX spectrum asserts the presence of all expected elements of La/SnO2 nanoparticles
with additional contributions of C and Cu from the TEM grid. Hence, the results support the
hypothesis that La elements form La2O3 crystallites distributed within SnO2 particles.
The BET specific surface area (SSABET, left axis)) and average BET particle diameter (dBET, right
axis) of La2O3-doped SnO2 nanoparticles (P-0 to P-2La) as a function of La content are shown in
Fig. 5. It is seen that SSABET increases considerably from 113.5 to 152.2 m2/g while the particle
diameter decreases substantially from 7.6 to 5.7 nm as the La content increases from 0 to 0.1
wt%. As the La doping level increases further to 0.5 wt%, the surface area and particle diameter

PT
slightly fluctuate and remain approximately the same. In contrast, SSABET decreases to 119.1
m2/g while dBET increases to 7.2 nm as the La content increases further to 2 wt%. It can be
observed that the dependence of dBET on the La content is similar to that of the average crystallite
size obtained from the XRD data (Inset of Fig. 1(a)). Thus, the particle and crystallite size tend to

RI
reduce with increasing La content, indicating that the inclusion of La species leads to less
coalescence of condensed SnO2 atoms during FSP synthesis due possibly to the lower flame

SC
synthesis temperature as a result of La precursor addition [20].
The surface chemical compositions and oxidation states of existing elements in 2 wt% La2O3-
doped SnO2 nanopowder (P-2La) as well as the sensing film (S-2La) after sensing test are
evaluated from the XPS data as illustrated in Fig. 6. It should be noted that only the XPS spectra

U
of 2 wt% La-doped SnO2 nanopowder are presented because La 3d signals of SnO2 nanopowders
with lower La concentrations are unclear or absent because of the limited resolution of XPS
N
techniques. The survey scan spectra (Fig. 6(a)) demonstrate the appearance of Sn 3d, La 3d and
O 1s from La-doped SnO2 materials and C 1s from organic surface contaminations. By omitting
A
the carbon contamination (Inset table in Fig. 6(a)), the La contents of P-2La and S-2La are
estimated to be 2.54 and 2.34 wt%, respectively. The results are in good agreement with the
M

EDX data and the intended La-doping level. Regarding Sn elemental states, the tin 3d core level
of P-2La (Fig. 6(b)) exhibits one doublet pair, Sn 3d5/2:Sn 3d3/2, at binding energies of
486.7:495.1 eV while the corresponding peaks of S-2La can be decomposed into the main and
D

tiny minor doublet pairs (Sn 3d5/2:Sn 3d3/2) at 486.7:495.1eV and 487.7:496.1 eV, respectively.
The main pair can be assigned to the highest oxidation state of Sn4+O while the minor pair of S-
TE

2La is corresponding to the Sn4+OH state due to water adsorption from humidity [20,21]. For
the La element (Fig. 6(c)), the La 3d doublet pair (La 3d5/2:La3d3/2) can be deconvoluted into
three pairs, comprising a main pair and two minor pairs. P-2La exhibits the main doublet pair (La
EP

3d5/2:La 3d3/2) at 834.6:851.5 eV and the minor doublet pairs at 833.7:850.2 eV and 835.8:852.6
eV, respectively, while the corresponding pairs of S-2La are slightly shifted to 834.6:851.5 eV,
835.8:852.6 eV and 833.5:850.4 eV, respectively. The main doublet pair can be assigned to La3+
CC

of La2O3 while the minor doublet pairs may be attributed to La3+ of La-OOH and La(OH)3
associated with water adsorption [22–24]. Thus, the elemental XPS data confirm the presence of
La2O3 species on the SnO2 support. In the case of oxygen element, O 1s core level of P-2La (Fig.
6(d)) can be split into four contributions centered at 530.5 (main peak), 531.7, 532.6 and 533.6
A

eV while those of S-2La locate at 530.5 (main peak), 531.4, 532.3 and 533.4 eV, respectively.
The main component of O 1s can be ascribed to lattice oxygen (O2) of La2O3-doped SnO2 lattice
on the outermost surface while the minor peaks located at 531.4–531.7, 532.3–532.6 and 533.4–
533.6 eV may be assigned to the chemisorbed oxygen species, hydroxyl groups due to water
adsorption and loosely bound oxygen species, respectively [25–27].
Figs. 7(a)–(f) demonstrate cross-sectional SEM micrographs of S-0, S-0.1La, S-0.2La, S-0.5La,
S-1La and S-2La sensing films fabricated by spin-coating on Al2O3 substrates equipped with Au
interdigitated electrodes after annealing and sensing test, respectively. It is seen that the films
8
 

contain densely agglomerated particles but with some pores and rough asperities. In addition, all
sensing films have similar thicknesses in the range of 12–16 µm, demonstrating good
reproducibility of binder/powder mixing and spin-coating methods [11,20]. Therefore, FSP
together with paste mixing and spin-coating techniques are practical for the fabrication of thick
film metal oxide gas sensors.
<H2>2.2. Gas-sensing properties
The gas-sensing properties of SnO2 sensors with different La contents were evaluated in terms of
sensor responses toward various VOCs and environmental gases including C3H6O, H2S, NO2,

PT
SO2, C6H6, C7H8, C8H10 and CH2O at their nominal concentrations and operating temperature of
350C as illustrated in Fig. 8. It is seen that the undoped SnO2 sensor exhibits relatively high
response to C3H6O and SO2 and low responses to other gases. Hence, the undoped SnO2 sensor
displays good selectivity towards C3H6O and SO2. With La2O3 doping at low La contents

RI
(0.10.2 wt%), the responses to C6H6, C7H8, and CH2O rise significantly and the responses to
C3H6O, C8H10 and SO2 increase moderately while the response to H2S weakly increases and the

SC
response to NO2 slightly decreases, resulting in high selectivity towards C6H6. At the moderate
La content of 0.5 wt%, the C3H6O response rises greatly while the responses to H2S and NO2
slightly increase but the responses to SO2, C6H6, C7H8, C8H10 and CH2O drop moderately.
Consequently, the La2O3-doped SnO2 sensor with 0.5 wt% La becomes highly selective to

U
C3H6O. However, the C3H6O response reduces considerably and the responses to other gases
degrade moderately as the La content increases further to 1 and 2 wt%. Thus, the SnO2 sensors
N
with high La contents show relatively low C3H6O selectivity compared with the sensor with the
optimal La content of 0.5 wt%. According to the results, the optimal La content of 0.5 wt% leads
A
to particularly significant improvement of acetone-sensing performances at the operating
temperature of 350C. Thus, only detailed results for acetone will be further elaborated.
M

The effect of operating temperature ranging from 150 to 400°C on the responses to 400 ppm
acetone (C3H6O) of SnO2 nanoparticles with different La contents is shown in Fig. 9. It can be
seen that the acetone response of each sensor initially rises to an optimal value at an optimal
D

operating temperature before decreasing as the temperature increases further. In addition, the
optimal temperature of SnO2 nanoparticles with low La contents (00.2 wt%) is around 300°C
TE

while those with higher La contents (0.51 wt%) exhibit a relatively high optimal working
temperature of 350°C. Moreover, the acetone response is seen to elevate significantly as the La
content increases up to 0.5 wt% but then becomes considerably lower as the La concentration
EP

increases further at all operating temperatures. Especially, the 0.5 wt % La-doped SnO2 sensor
(S-0.5La) displays the highest response of ~3,626 to 400 ppm acetone at the optimal working
temperature of 350°C. The roles of operating temperature and La2O3 doping will be further
CC

discussed in the subsequent section.


The changes in resistances of SnO2 sensing films with different La concentrations (S-0 to S-2La)
under exposure to various C3H6O concentrations in the high concentration range (6–400 ppm) at
the optimal operating temperature are demonstrated at Fig. 10(a). It is seen that the baseline
A

resistance of SnO2 sensor increases substantially by around two orders of magnitude as the La
concentration increases from 0 to 2 wt%. The elevation of resistance with increasing La content
will be described based on La2O3-SnO2 heterostructures to be discussed in the next section. After
subjected to acetone vapor, the resistances of all sensors reduce abruptly, confirming the typical
n-type semiconducting behavior towards a reducing gas. In addition, the resistance change of
SnO2 sensor enhances significantly as the La content increases up to 0.5 wt% before declining
considerably when the La content increases further to 2 wt%. At low acetone concentrations
(620 ppm), the noise is observed during acetone exposure. It could arise from the large
9
 

fluctuation of acetone concentration due to the bubbling process at very low bubbling flow rates
(< 1 cm3/min) when the fluctuation cannot be sufficiently stabilized by the gas mixing process.
To verify the cause of noise problem and attain lower acetone concentrations, addition gas
sensing measurements were carried out at low acetone concentrations of 0.1–4 ppm using a
standard gas bottle containing 50 ppm acetone balanced in dry air. The corresponding changes in
resistances of SnO2 sensing films under exposure to various C3H6O concentrations in the low
range (0.1–4 ppm) are displayed in Fig. 10(b). It can be seen that the resistances upon exposure
to acetone derived from the acetone standard gas cylinder contain negligible noise at acetone

PT
concentrations low than 6 ppm, in contrast to the results from the bubbling test. The results prove
that the observed noise in Fig. 10(a) is indeed due to the use of bubbling process and the sensors
are stable after acetone exposure. In addition, the trend of resistance changes of SnO2 sensors
with different La contents follows the same tendency as the results in the high acetone

RI
concentration range in Fig. 10(a) with 0.5 wt% being the optimal La content. The results confirm
the repeatability of sensor responses and show that the La2O3-doped SnO2 sensors can detect

SC
acetone at a concentration as low as 0.1 ppm.
The related acetone-sensing characteristics in terms of sensor response (solid line, left axis) and
response time (dash line, right axis) of S-0 to S-2La as a function of acetone concentration in the
range of 0.1–400 ppm are displayed Fig. 10(c). It can be seen that the sensor response and

U
response time improve substantially as the La content increases from 0 to 0.5 wt% but then
rapidly degrade when the La level additionally increases to 2 wt%. Specifically, the optimal
N
La2O3-doped SnO2 sensor (S-0.5La) offers a very high acetone response (S = 3,626 at 400 ppm)
and a short response time (tres= 2.8 s) compared with S-1La (S = 1,019, tres= 3.3 s), S-2La (S =
A
332, tres= 3.4 s), S-0.2La (S = 18.1, tres= 2.9 s), S-0.1La (S = 16.3, tres= 3.8 s) and S-0 (S = 8.2,
tres= 4.9 s). The achieved response value of 3,626 is significantly higher than those of other
M

SnO2-based and Ferrite-based acetone sensors, which offered sensor response values of <170 to
acetone in the same range of concentrations as listed in Table 1 [11]. Regarding the response
dependency on the acetone concentration, the responses of all sensors increase monotonically
D

with increasing acetone concentration according to the well-known modified power law
equations as listed in the inset labels. It is also seen that the exponent values of SnO2 sensors
TE

with low La contents (00.2 wt%) are slightly above 0.5, implying that the oxygen species
interacting with acetone molecules are mostly O2. In contrast, the exponent values of sensors
with moderate and high La levels (0.52 wt%) are close to 1, identifying O as the main reacting
EP

oxygen species. At the lowest acetone concentration of 0.1 ppm, the optimal sensor still offers a
good response of ∼1.2, corresponding to the low acetone detection limit of 0.08 ppm (predicted
at the response of 1.1 based on the power-law equation). The attained detection limit is still
CC

better than other highly sensitive acetone sensors including porous Yb-Ferrite films and ZnO–
Fe3O4 inverse opal structures, which offered similar detection limits at the expense of high
material cost and high working temperature, respectively [6–11,16,17]. Therefore, the optimal
La2O3-doped SnO2 sensor has a great potential for air pollution monitoring and breath analysis of
A

acetone.
<H2>2.3. Sensing mechanism
From the results, the sensor resistance increases monotonically with increasing La2O3 doping
concentration while the sensor response to acetone of SnO2 nanoparticles is optimally enhanced
at the moderate La content of 0.5 wt%. The roles of La2O3dopant on sensor characteristics may
be explained on the basis of the heterojunction (La2O3SnO2) and catalytic as well as specific
surface area and grain size effects. From the BET and XRD results, the specific surface area
increases while the particle diameter and grain size reduce considerably due to La2O3 doping at
10
 

low and moderate La contents of 0.10.5 wt%. Therefore, the observed gas-sensing enhancement
could be partly attributed to the enhancement of surface area and the reduction of structural sizes
due to the influence of La2O3 dopant as an effective grain growth inhibitor [28]. However, the
observed changes of specific surface area and structure sizes are only within a factor of 2, which
are still much smaller than the large increment of acetone response. Thus, additional mechanisms
relating to heterojunction and catalytic effects of La2O3 dopant should be the main contribution
to the observed gas response enhancement.
The role of La2O3 dopant on resistance and acetone-sensing mechanisms of SnO2 nanoparticles

PT
will be explained based on the proposed models as depicted in Fig. 11. For the undoped SnO2
nanoparticles (Fig. 11(a)), the resistance is controlled by the number of electrons created by the
thermal activation of intrinsic defects (oxygen vacancies) as well as chemisorbed oxygen species

RI
( O2 , O , O2 ). The oxygen chemisorption process abstracts electrons from the conduction band of
SnO2, inducing surface electron depletion regions, lowering the Fermi level and raising the
resistance [29]. Nevertheless, the undoped SnO2 material still exhibits relatively low electrical

SC
resistance compared with several other metal oxides due to its high level of oxygen vacancies
[3–7]. From TEM and STEM data (Figs. 3–4), the sizes of SnO2 particles are as small as 10 nm
and the width of depletion regions and Debye length could be on the order of tens of nm. It is

U
thus possible that the FSP-made SnO2 nanoparticles are completely depleted in air before gas
exposure. However, the experimental results (referred to the response to oxidizing NO2 gas in the
N
inset of Fig. 8) have proven that these nanoparticles are still quite far from the fully depleted
condition in air at the optimal working temperature of 350C because the sensors still exhibits a
A
considerable n-type NO2 response and the n-type NO2 response can increase further at higher
NO2 concentrations. The substantial increase of resistance upon NO2 exposure indicates that
M

adsorption can take a substantial amount of remaining electrons in conduction band,


leading to the expansion of depletion regions. To clarify this further, the depletion width of FSP-
made undoped SnO2 nanoparticles is estimated to be 3.3 nm by the method based on the
D

Schottky barrier model [30] using the previously reported resistance data of FSP-made undoped
and Ni-doped SnO2 nanoparticles [11]. The value is comparable with other reported depletion
TE

width and Debye length of 3 nm for undoped SnO2, further confirming that the SnO2
nanoparticles are not fully depleted since the depletion width is smaller than the particle radius
[31,32]. With La2O3 doping, La2O3 crystallites should be formed and dispersed in and on SnO2
nanoparticles as represented in Figs. 11(b)(d). La2O3 is known to exhibit p-type
EP

semiconductivity and p-n heterojunctions of La2O3SnO2 will thus be formed upon La2O3 doping
according the energy band diagram as depicted in Fig. 12. In Fig. 12 (a), the energy band
positions of La2O3 and SnO2 at the point of zero charge before forming the heterocontact are
CC

theoretically calculated based on the Pearson absolute electronegativity method using the
electronic parameters and equations as listed in Table 2 [33–35]. After the equalization of Fermi
level, energy bands bend due to energy difference, resulting in the formation of the potential
A

barriers for electrons and holes at the p-n heterojunction and inducing additional depletion
regions as projected in Fig. 11(b)(d). In this band structure, holes from p-type La2O3 are
blocked by the discontinuous potential barrier at the interface while electrons will migrate into
internal SnO2 regions due to potential fields. Thus, the electrical conduction will be dominated
by the n-type SnO2 layer. At low and moderate La contents (0.10.5 wt%), the sizes of La2O3
crystallites will be small since aggregation can hardly occur at a low material density and the
number of La2O3 crystallites will increase substantially with increasing doping concentration
(Figs. 11(b)(c)). Consequently, the depletion regions will expand significantly and the electrical
11
 

resistance of La2O3-doped SnO2 nanoparticles increases rapidly with increasing La content (Fig.
10(a)). At higher La contents (12 wt%), La2O3 crystallites are likely to aggregate into larger
ones, resulting in a lower number of crystallites, less rapid expansion of depletion region and less
increase of resistance with increasing La content (Fig. 10(a)). In addition, the formation of La2O3
crystals involves the creation of oxygen vacancies during the transformation from LaO+ to La2O3
on the SnO2 matrix according to the chemical reactions represented in Kröger-Vink notation as
[36]:
La3  H 2O( g )  LaO  2H  (1)

PT
1
Oox  Vo  2e   O2 (2)
2

RI
2 LaO   Oox  La2O3  Vo (3)

SC
where H+, Oox and Vo are protons, neutral oxygen atoms in lattice and doubly ionized oxygen
vacancies, respectively. The process generates new oxygen vacancies and form La3+−Vo

U
complexes that will lead to lower conductivity of La2O3 crystallites.
To experimentally verify the presence of La2O3−SnO2 (p-n) heterojunctions and investigate the
N
related conduction mechanisms, current-voltage (I-V) characteristics of La2O3-doped SnO2
nanoparticles with different La contents before and after exposure to 4 ppm acetone were
A
measured in the voltage range of ±100 V at 350C as illustrated in Fig. 13(a) and (b),
respectively. In addition, the magnitudes of current-voltage values for forward and reverse bias
M

regions are plotted in log-log scales order to better observe the degree of linearity as displayed in
the upper and lower insets, respectively. Before acetone exposure in air (Fig. 13(a)), undoped
SnO2 exhibits a linear and symmetric I-V characteristic under forward and reverse biases. The
D

linear I-V characteristic of undoped SnO2 nanoparticles confirms the typical ohmic behavior and
the carrier-scattering-limited current conduction mechanism of the n-type SnO2 semiconductor
TE

[8]. With La2O3 doping, the I-V curves becomes increasingly asymmetric and nonlinear
especially under the forward bias as the La content increases from 0.1 to 2 wt%. In addition, the
current at a given applied voltage decreases substantially and displays more rectifying
characteristics with increasing La content. It has been demonstrated that the asymmetric
EP

rectifying I-V behaviors can signify the presence of p-n heterojunctions in metal oxide
composites [33,37], supporting the formation of La2O3-SnO2 heterojunctions in the La2O3-doped
SnO2 nanoparticles. In addition, the decrease of current with increasing La content also
CC

corresponds to the increasing amount of depletion regions induced around the p-n
heterojunctions as depicted in Figs. 11(b)(d). Under forward bias (the upper inset of Fig. 13(a)),
the slopes of I-V plots in log-log scale of La2O3-doped SnO2 nanoparticles are constant and
A

approximately equal to 1 at low voltages and then tend to increase at higher voltages (40100
V), indicating changes of conduction mechanisms of La2O3-doped SnO2 nanoparticles under
high field conditions. Additionally, the slopes at high voltages increase more rapidly and
significantly to a value of greater than 2 with increasing La content, suggesting the transition to a
space-charge limited current conduction mechanism [33], which can arise due to additional space
charges in the depletion regions, which become more dominant with increasing La2O3 doping
level. In the reverse bias regime (the lower inset of Fig. 13(a)), the slopes of I-V plots in log-log
scale are also around 1 at low voltages but later decrease at higher voltages (50100 V), which
12
 

may be ascribed to the lower reverse saturation current from some p-n junctions along the
applied field [37], which are prevailing at higher La2O3 doping levels.
Upon exposure to acetone, C3H6O molecules will interact with the chemisorbed oxygen

species ( O  ,O 2 ) according to the reactions [9]:

C3 H 6O( g )  8O (ads)  3CO2  3H 2O  8e  (4)

PT

C3 H 6O( g )  8O 2(ads)  3CO2  3H 2O  16e  (5)

RI
The reactions result in the release of trapped electrons at oxygen species into the conduction
band of SnO2, leading to higher electron carrier concentration, recession of depletion regions and
lower sensor resistance. From the I-V curves of La2O3-doped SnO2 nanoparticles upon exposure

SC
to 4 ppm acetone in Fig. 13(b), the I-V characteristics are apparently more linear with higher
slope values compared with the corresponding I-V curves in air (Fig. 13(a)), confirming the
reduction of sensor resistance and indicating a change of conduction mechanism of La2O3-doped

U
SnO2 nanoparticles. The more linear I-V behaviors upon acetone exposure dictate that the main
conduction mechanism of La2O3-doped SnO2 nanoparticles becomes limited by the scattering of
N
majority charge carriers or electrons, which were released from preadsorbed oxygen species via
the reducing reaction with acetone molecules. This ohmic n-type conduction process is
A
predominant over the conduction across p-n heterojunctions even at high voltages because
electrons released from oxygen species will be mostly transferred to the n-type SnO2
M

semiconductor [38]. From Fig. 13(b), only La2O3-doped SnO2 nanoparticles with 2 wt% La
exhibit a partial rectifying behavior of p-n heterojunctions under the forward bias (the upper inset
of Fig. 13 (b)). The p-n hetero junction effect can partially be observed in this specific case
D

because there are relatively large amount of p-n junctions while the concentration of released
electrons from gas interaction is relatively small.
TE

The influence of La doping level on the gas response can thus be further considered based on the
heterostructures of La2O3 crystallites embedded within SnO2 matrix. The presence of
La2O3SnO2 heterojunctions can considerably affect the availability and type of oxygen species
as well as the reducing reaction rate with acetone. Without La2O3 doping, the acetone reducing
EP

reactions on SnO2 surface require a high activation energy and hence can occur only at a few
active adsorption sites, leading to a low acetone response as exemplified in Fig. 11(a). With
La2O3 doping, La2O3SnO2 heterojunctions may enhance the acetone reaction rate via the
CC

enhancement of chemisorbed peroxide ions that will further dissociate into oxygen ions (O)
and transfer to the surface oxygen vacancies of SnO2 due to the strong surface basicity of La2O3
sites [36]. This is in accordance with the results from power-law analysis that there is more
contribution of O species in the responsive reaction with acetone (Fig. 10(b)). Furthermore,
A

La2O3 can stimulate H-abstraction chemical reactions, which will lower the reaction energy for
the oxidation of hydrated carbon species including acetone and ethanol. Thus, the basic
properties of La2O3 at the heterojunctions may promote the sensitivity and selectivity towards
hydrated carbon molecules via surface dehydrogenation [36]. Furthermore, the results in this
study demonstrate superior response enhancement of acetone over ethanol, which is similar to
some studies of La-SnO2 sensors but different from some others [3–7]. A possible explanation
for the selectivity effect is that the La2O3 crystallites on SnO2 surface produced by FSP process
13
 

may orient in some particular crystallographic orientations that exhibit suitable basic properties
for abstraction of the ketone group of acetone. Consequently, the reducing reactions can take
place particularly around La2O3 surface sites, leading to enhanced acetone responses as
illustrated in Figs. 11(b)(d). At a low La content (Fig. 11(b)), the response enhancement is still
not large due low number of La2O3 surface sites. When the La content rises to a moderate level
(Fig. 11(c)), the number of La2O3 surface sites increases substantially, resulting in a greatly
enhanced acetone response. However, La2O3 crystallites will aggregate into larger ones when the
La content becomes too high (Fig. 11(d)), leading to less amount of La2O3 surface sites and a

PT
less enhanced acetone response.
Regarding the effect of operating temperature, the SnO2 nanoparticles with low La contents
(0.2 wt%) exhibit the optimal operating temperature of 300°C while those with higher La

RI
contents display a slightly higher optimal operating temperature of 350°C but with a much
higher acetone response. The initial increase of response with increasing temperature is well
known to be ascribed to increasing oxygen chemisorption and reducing reaction rates due to

SC
higher thermal energy while the reduction in response at high temperature is typically due to the
difficulty in exothermic vapor adsorption, leading to the dominance of chemical desorption [5].
For undoped SnO2, O2 ions exist on surface at the temperature below 160°C before

U
transforming into Oand O2 ones at higher temperatures while O2 and O  / O 2  ions desorb from
the surface as the temperature goes below 150 and 560°C, respectively [39]. Thus, the acetone
N
response of undoped SnO2 is optimized at around 300°C when the adsorption/desorption rate
ratios of oxygen species (O and O2) are optimal. With the low La contents, the optimal
A
operating temperature is still the same, indicating insignificant change of the
adsorption/desorption behaviors of oxygen species. At the moderate and high La contents
M

(0.5wt%), the optimal operating temperature increases slightly to 350°C due possibly to the
transition of dominant oxygen species from O2 to O (according to the power-law analysis),
which may be attributed to the strong surface basicity of La2O3 [36].
D

<H1>3. Conclusions
In conclusion, FSP-made La2O3-doped SnO2 nanoparticles with 0−2 wt% La contents were
TE

systemically studied for acetone-sensing applications. Structural characterization results showed


that SnO2 nanostructures exhibited mainly the spheroidal morphology with a polycrystalline
tetragonal SnO2 phase and La2O3 crystallites should form a solid solution with the SnO2 lattice.
EP

Gas-sensing results demonstrated that the SnO2 sensing film with the optimal La content of 0.5
wt% exhibited a very high response of ∼3,626 toward 400 ppm acetone with a short response
time within a few seconds at the optimal operating temperature of 350°C. The enhancement of
CC

acetone response by La2O3 doping was primarily explained based on the formation of
La2O3SnO2 heterojunctions, which could be experimentally dictated from I-V characteristic
data. Furthermore, the optimal sensor displayed high acetone selectivity against SO2, H2S, NO2,
C6H6, C7H8, C8H10 and CH2O. Therefore, the FSP-made La2O3-doped SnO2 sensors are potential
A

for responsive and selective detections of acetone.


<ACK>Acknowledgements

The authors gratefully acknowledge the financial support from Thailand Graduate Institute of
Science and Technology (TGIST) (SCA-CO-2558-1039-TH, TG-44-10-58-018M), National
Science and Technology Development Agency (NSTDA), National Research University (NRU)
Project under the Office of the Higher Education Commission (CHE), Ministry of Education,
National Research Council of Thailand (NRCT), Thailand Research Fund (TRF: RSA6080014,
14
 

IRG5780013), Center of Excellence (CoE), Graduate School, Materials Science Research


Center, Department of Physics and Materials Science, Faculty of Science, Chiang Mai
University. The special thanks should be given to the National Electronics and Computer
Technology Center (NECTEC), Pathumthani, Thailand for the sensor facilities.
<REF>References

<BIBL>
[1] A.W. Boots, J.B.N. van B. Joep, W.D. Jan, S. Agnieszka, F.W. Emile, J. van S. Frederik,;1;

PT
The versatile use of exhaled volatile organic compounds in human health and disease, J. Breath
Res. 6 (2012) 27108. <DOI>10.1088/1752-7155/6/2/027108.</DOI>
[2] A. Caron, N. Redon, F. Thevenet, B. Hanoune, P. Coddeville,;1; Performances and
limitations of electronic gas sensors to investigate an indoor air quality event, Build. Environ.

RI
107 (2016) 19–28. <DOI>10.1016/j.buildenv.2016.07.006.</DOI>
[3] A. Mirzaei, S.G. Leonardi, G. Neri,;1; Detection of hazardous volatile organic compounds

SC
(VOCs) by metal oxide nanostructures-based gas sensors: A review, Ceram. Int. 42 (2016)
15119–15141. <DOI>10.1016/j.ceramint.2016.06.145.</DOI>
[4] S. Kundu, A. Choudhury, S.M. Mursalin, M. Narjinary, R. Manna,;1; Synthesis,
characterization and low concentration ethanol sensing performance of sol-gel derived La(III)

U
doped tin oxide, J. Mater.Sci. Mater. Electron. 26 (2015) 6252–6260. <DOI>10.1007/s10854-
015-3211-0.</DOI>
N
[5] S. Shi, Y. Liu, Y. Chen, J. Zhang, Y. Wang, T. Wang,;1; Ultrahigh ethanol response of SnO2
nanorods at low working temperature arising from La2O3 loading, Sens. Actuators B Chem. 140
A
(2009) 426–431. <DOI>10.1016/j.snb.2009.04.058.</DOI>
[6] Y. Wu, H. Zhang, Y. Liu, W. Chen, J. Ma, S. Li, Z. Qin,;1; Synthesis and gas sensing
M

properties of single La-doped SnO2 Nanobelts, Sensors. 15 (2015) 14230–14240.


<DOI>10.3390/s150614230.</DOI>
[7] G.T. Ang, G.H. Toh, M. Zailani, A. Bakar, A.Z. Abdullah,;1; High sensitivity and fast
D

response SnO2 and La-SnO2 catalytic pellet sensors in detecting volatile organic compounds,
Process Saf. Environ. Prot. 89 (2010) 186–192. <DOI>10.1016/j.psep.2010.10.008.</DOI>
TE

[8] L.P. Chikhale, J.Y. Patil, A. V. Rajgure, F.I. Shaikh, I.S. Mulla, S.S. Suryavanshi,;1;
Structural, morphological and gas sensing properties of undoped and Lanthanum doped
nanocrystalline SnO2, Ceram. Int. 40 (2014) 2179–2186.
EP

<DOI>10.1016/j.ceramint.2013.07.136.</DOI>
[9] X. Li, Y. Liu, S. Li, J. Huang, Y. Wu, D. Yu,;1; The sensing properties of single Y-doped
SnO2 nanobelt device to acetone, Nanoscale Res. Lett. 11 (2016) 470. <DOI>10.1186/s11671-
CC

016-1685-1.</DOI>
[10] Z. Jiang, R. Zhao, B. Sun, G. Nie, H. Ji, J. Lei, C. Wang,;1; Highly sensitive acetone sensor
based on Eu-doped SnO2 electrospun nanofibers, Ceram. Int. 42 (2016) 15881–15888.
<DOI>10.1016/j.ceramint.2016.07.060.</DOI>
A

[11] S. Singkammo, A. Wisitsoraat, C. Sriprachuabwong, A. Tuantranont, S. Phanichphant, C.


Liewhiran,;1; Electrolytically exfoliated graphene-loaded flame-made Ni-doped SnO2 composite
film for acetone sensing, ACS Appl. Mater. Interfaces. 7 (2015) 3077–3092.
<DOI>10.1021/acsami.5b00161.</DOI>
[12] A.P. Rambu, F. Tudorache, I. Petrila, G.G. Rusu, V. Nica, M. Dobromir, S. Tascu,;1;
Combined effects of p–n heterojunctions and active surface areas in a composite material
dedicated to gas sensing applications, J. Mater. Sci. Mater. Electron. 26 (2015) 9837–9844.
<DOI>10.1007/s10854-015-3658-z.</DOI>
15
 

[13] N. Rezlescu, E. Rezlescu, F. Tudorache, P.D. Popa,;1; Gas sensing properties of porous Cu-,
Cd- and Zn- ferrites, Rom. Reports Phys. 61 (2009) 223–234.
https://www.scopus.com/inward/record.uri?eid=2-s2.0-
70350614100&partnerID=40&md5=ad1a564646061bfcc3bf2ae100c9f6cf.
[14] F. Tudorache, E. Rezlescu, P.D. Popa, N. Rezlescu,;1; Study of some simple ferrites as
reducing gas sensors, J. Optoelectron. Adv. Mater. 10 (2008) 1889–1893.
https://www.scopus.com/inward/record.uri?eid=2-s2.0-
49149101502&partnerID=40&md5=c09f1b0ea5c69c78078912e4f3dfaa7e.

PT
[15] F. Tudorache, P.D. Popa, M. Dobromir, F. Iacomi,;1; Studies on the structure and gas
sensing properties of nickel–cobalt ferrite thin films prepared by spin coating, Mater. Sci. Eng.
B. 178 (2013) 1334–1338. <DOI>10.1016/j.mseb.2013.03.019.</DOI>
[16] P. Zhang, H. Qin, W. Lv, H. Zhang, J. Hu,;1; Gas sensors based on ytterbium ferrites

RI
nanocrystalline powders for detecting acetone with low concentrations, Sensors Actuators B
Chem. 246 (2017) 9–19. <DOI>https://doi.org/10.1016/j.snb.2017.01.096.</DOI>

SC
[17] L. Zhang, B. Dong, L. Xu, X. Zhang, J. Chen, X. Sun, H. Xu, T. Zhang, X. Bai, S. Zhang,
H. Song,;1; Three-dimensional ordered ZnO–Fe3O4 inverse opal gas sensor toward trace
concentration acetone detection, Sensors Actuators B Chem. 252 (2017) 367–374.
<DOI>https://doi.org/10.1016/j.snb.2017.05.167.</DOI>

U
[18] F. Tudorache, M. Grigoraş,;1; Study of polyaniline - Iron oxides composites using for gas
detection, Optoelectron. Adv. Mater. Rapid Commun. 4 (2010) 43–47.
N
https://www.scopus.com/inward/record.uri?eid=2-s2.0-
77951998906&partnerID=40&md5=8b35ab79fd7852ea8829b2e947d38fc6.
A
[19] S.G. Kim, W.T. Kim, Y.B. Park,;1; Abnormal grain growth induced by boundary
segregation of solute atoms, Mater. Sci. Forum. 558–559 (2007) 1093–1099.
M

<DOI>10.4028/www.scientific.net/MSF.558-559.1093.</DOI>
[20] M.J. Height, L. Mädler, S.E. Pratsinis, F. Krumeich,;1; Nanorods of ZnO made by flame
spray pyrolysis, Chem. Mater. 18 (2006) 572–578. <DOI>10.1021/cm052163y.</DOI>
D

[21] C.D. Wagner, W.M. Riggs, D. L.E., J.F. Moulder, M. G.E.,;1; Handbook of X-ray
Photoelectron Spectroscopy, Perkin-Elmer Corp. 3 (1979) 84. <DOI>10.1093/ilj/1.1.251.</DOI>
TE

[22] M. Kwoka, N. Waczyńska, P. Kościelniak, M. Sitarz, J. Szuber,;1; X-ray photoelectron


spectroscopy and thermal desorption spectroscopy comparative studies of L-CVD SnO2 ultra
thin films, Thin Solid Films. 520 (2011) 913–917. <DOI>10.1016/j.tsf.2011.04.185.</DOI>
EP

[23] N. Sethulakshmi, A.N. Unnimaya, I.A. Al-Omari, S. Al-Harthi, S. Sagar, S. Thomas, G.


Srinivasan, M.R. Anantharaman,;1; On magnetic ordering in heavily sodium substituted hole
doped lanthanum manganites, J. Magn. Magn. Mater. 391 (2015) 75–82.
CC

<DOI>10.1016/j.jmmm.2015.04.092.</DOI>
[24] A. Machocki, T. Ioannides, B. Stasinska, W. Gac, G. Avgouropoulos, D. Delimaris, W.
Grzegorczyk, S. Pasieczna,;1; Manganese-lanthanum oxides modified with silver for the
catalytic combustion of methane, J. Catal. 227 (2004) 282–296.
A

<DOI>10.1016/j.jcat.2004.07.022.</DOI>
[25] S. Mickevičius, S. Grebinskij, V. Bondarenka, B. Vengalis, K.;1; Šliužiene, B.A. Orlowski,
V. Osinniy, W. Drube, Investigation of epitaxial LaNiO3-x thin films by high-energy XPS, J.
Alloy. Compd. 423 (2006) 107–111. <DOI>10.1016/j.jallcom.2005.12.038.</DOI>
[26] M.F. Sunding, K. Hadidi, S. Diplas, O.M. Løvvik, T.E. Norby, A.E. Gunnæs,;1; XPS
characterisation of in situ treated lanthanum oxide and hydroxide using tailored charge
referencing and peak fitting procedures, J. Electron Spectros. Relat. Phenomena. 184 (2011)
399–400. <DOI>10.1016/j.elspec.2011.04.002.</DOI>
16
 

[27] E. Baškys, V. Bondarenka, S. Grebinskij, M. Senulis, R. Sereika,;1; XPS study of sol-gel


produced lanthanum oxide thin films, Lith. J. Phys. 54 (2014) 120–124.
<DOI>10.3952/lithjphys.54208.</DOI>
[28] J.S. Wang, M.L. Zhou, Y.M. Wang, Z.R. Nie, J.X. Zhang, T.Y. Zuo,;1; Chemical stability
of La2O3 in La2O3-Mo cathode materials, Trans. Nonferrous Met. Soc. China. 11 (2001) 681–
683.
[29] M.M. Rahman, A. Jamal, S.B. Khan, M. Faisal,;1; Highly sensitive ethanol chemical sensor
based on Ni-doped SnO2 nanostructure materials, Biosens. Bioelectron. 28 (2011) 127–134.

PT
<DOI>10.1016/j.bios.2011.07.024.</DOI>
[30] J. Liu, X. Liu, Z. Zhai, G. Jin, Q. Jiang, Y. Zhao, C. Luo, L. Quan,;1; Evaluation of
depletion layer width and gas-sensing properties of antimony-doped tin oxide thin film sensors,
Sensors Actuators B Chem. 220 (2015) 1354–1360.

RI
<DOI>https://doi.org/10.1016/j.snb.2015.07.065.</DOI>
[31] N. Yamazoe,;1; New approaches for improving semiconductor gas sensors, Sens. Actuators

SC
B Chem. 5 (1991) 7–19. <DOI>http://dx.doi.org/10.1016/0925-4005(91)80213-4.</DOI>
[32] C. Xu, J. Tamaki, N. Miura, N. Yamazoe,;1; Grain size effects on gas sensitivity of porous
SnO2-based elements, Sens. Actuators B Chem. 3 (1991) 147–155.
<DOI>http://dx.doi.org/10.1016/0925-4005(91)80207-Z.</DOI>

U
[33] H. Tian, H. Fan, G. Dong, L. Ma, J. Ma,;1; NiO/ZnO p-n heterostructures and their gas
sensing properties for reduced operating temperature, RSC Adv. 6 (2016) 109091–109098.
<DOI>10.1039/C6RA19520B.</DOI> N
[34] X.-L. Ding, Z.-Y. Li, J.-H. Meng, Y.-X. Zhao, S.-G. He,;1; Density-functional global
A
optimization of (La2O3)n clusters, J. Chem. Phys. 137 (2012) 214311–214317.
<DOI>10.1063/1.4769282.</DOI>
M

[35] E.A. Floriano, L.V.A. Scalvi, M.J. Saeki, J.R. Sambrano,;1; Preparation of TiO2/SnO2 thin
films by sol-gel method and periodic B3LYP simulations, J. Phys. Chem. A. 118 (2014) 5857–
5865. <DOI>10.1021/jp411764t.</DOI>
D

[36] G. Zhang, S. Zhang, L. Yang, Z. Zou, D. Zeng, C. Xie,;1; La2O3-sensitized SnO2


nanocrystalline porous film gas sensors and sensing mechanism toward formaldehyde, Sens.
TE

Actuators B Chem. 188 (2013) 137–146. <DOI>10.1016/j.snb.2013.07.002.</DOI>


[37] H. Yoo, H. Kim, D. Kim, Rectifying and;1; NO gas sensing properties of an oxide
heterostructure with ZnO nanorods embedded in CuO thin film, Nanosci. Nanotechnol. Lett. 7
EP

(2015) 758–762.
http://www.ingentaconnect.com/content/asp/nnl/2015/00000007/00000009/art00014.
[38] L. Ma, H. Fan, H. Tian, J. Fang, X. Qian,;1; The n-ZnO/n-In2O3 heterojunction formed by a
CC

surface-modification and their potential barrier-control in methanal gas sensing, Sensors


Actuators B Chem. 222 (2016) 508–516.
<DOI>https://doi.org/10.1016/j.snb.2015.08.085.</DOI>
[39] N. Yamazoe, J. Fuchigami, M. Kishikawa, T. Seiyama,;1; Interactions of tin oxide surface
A

with O2, H2O and H2, Surf. Sci. 86 (1979) 335–344. <DOI>http://dx.doi.org/10.1016/0039-
6028(79)90411-4.</DOI>

</BIBL>
Biographies

<remove picture pageno 31>Nantikan Tammanoon received her B.Sc. (2nd honor) from
Chiang Mai University in 2015 on Materials Science. She successfully received the Young
17
 

Scientist and Technologist Program (YSTP-2014) Scholarship from the National Science and
Technology Development Agency (NSTDA), Thailand. She was currently a M.S. student in the
Division of Materials Science, Department of Physics and Materials Science at Chiang Mai
University until the present under financial support from Thailand Graduate Institute of Science
and Technology (TGIST) during 2015–2016. Her research program focuses on the
environmental gas sensor based on metal oxide and functionalized graphene.
<remove picture pageno 32>Anurat Wisitsoraat received his Ph.D., M.S. degrees from
Vanderbilt University, TN, U.S.A., and B. Eng. degree in Electrical Engineering from

PT
Chulalongkorn University, Bangkok, Thailand in 2002, 1997, and 1993, respectively. His
research interests include microelectronic fabrication, semiconductor devices, electronic and
optical thin film coating, gas sensors, and micro-electromechanical systems (MEMS).
<remove picture pageno 32>Ditsayut Phokharatkul is an assistant researcher at MEMS

RI
Laboratory in NECTEC, Thailand. He received his B.E. and M.E. degrees from Nagoya
University (Japan) in 2006 and 2008, respectively. His dissertation is development of high

SC
density horizontally aligned carbon nanotubes for CNT-FET application. His research interests
include microelectronic fabrication, semiconductor devices, carbon nanotubes, graphene and
sensors.
<remove picture pageno 32>Adisorn Tuantranont received the B.S. degree in Electrical

U
engineering from King Mongkut’s Institute of Technology Ladkrabang, Thailand, in 1995, and
the M.S. and Ph.D. degrees in electrical engineering from the University of Colorado at Boulder
N
in 2001. Since 2001, he has been the director of the Nanoelectronics and MEMS Laboratory,
National Electronic and Computer Technology Center (NECTEC), Pathumthani, Thailand. His
A
research interests are in the area of micro electro-mechanical systems (MEMS), nanoelectronics,
lab-on-a-chip technology and printed electronics. He has authored more than 50 refereed
M

journals, 150 proceedings, and holds five patents. He also received the Young Technologist
Award in 2004 from the Foundation for the Promotion of Science and Technology under the
patronage of H.M. the King, Thailand.
D

<remove picture pageno 33>Sukon Phanichphant is an Associate Professor in Chemistry at


Department of Chemistry, Faculty of Science, Chiang Mai University, since 1977. She is
TE

currently a senior researcher at the Materials Science Research Center, Faculty of Science,
Chiang Mai University. Her research interests include synthesis and characterization of
nanoparticles for use in medical and sensor applications as well as synthesis and characterization
EP

of conducting polymer for light-emitting devices.


<remove picture pageno 33>Visittapong Yordsri received his B. Eng. degree in Material
Engineering from Kasetsart University, Bangkok, Thailand in 2004. His research interests
CC

include nanomaterials characterization by transmission electron microscope. His current work is


about nanocarbon materials and characterizations. He is currently a senior technician of
transmission electron microscope laboratory.
A

<remove picture pageno 34>Chaikarn Liewhiran received his B.Sc. from Srinakharinwirot
University in 2002 on Physics, M.S. and Ph.D. degrees of Materials Science from the Chiang
Mai University in 2004 and 2006, respectively. He received the Associate Professor in 2013 and
was a lecturer in the Department of Physics and Materials Science at Chiang Mai University
until the present, He currently received the TRF-CHE Young Scopus Researcher Award in
Physical Science in 2014 from Thailand by the joining of Thailand Research Fund (TRF), Office
of the Higher Education Com-mission (CHE) and Scopus. His research program focuses on the
18
 

Nanoscience and Nanotechnology, the fundamentals of physical and chemical synthesis of metal
oxide and metal–ceramic nanoparticles and their applications in nanocomposites, and the
development of novel nanomaterials in selective bio- and chemical gas sensing for
environmental monitoring.

Figures & Captions

PT
<Figure>Fig. 1. XRD patterns of FSP-made La2O3-doped SnO2 with 0–2wt% La (a)
nanopowders (P-0 to P-2La) and corresponding sensing films (S-0 to S-2La) on Au/Al2O3
substrates after annealing and sensing test. Inset: Crystallite size versus La concentration of SnO2
nanopowders.

RI
SC
<Figure>Fig. 2. SEM image (top view), EDX maps and EDX spectrum of La2O3-doped SnO2

sensing films with 2 wt%La on Au/Al2O3 substrates after annealing and sensing test.

U
N
<Figure>Fig. 3. BF-TEM and HR-TEM images of (a, b) FSP-made SnO2 (P-0), (c, d)
A
La2O3/SnO2 with 0.5 wt% La (P-0.5La) and (e, f) La2O3/SnO2 with 2 wt% La (P-2La). Insets: the
corresponding SAED patterns.
M

<Figure>Fig. 4. STEM image, EDX maps and EDX spectrum of La2O3-doped SnO2
nanoparticles with 2 wt% La (P-2La).
D

<Figure>Fig. 5. Specific surface area and average particle diameter of 0–2 wt% La2O3-doped
TE

SnO2 nanoparticles (P-0 to P-2La) as a function of La content.

<Figure>Fig. 6. XPS spectra of La2O3-doped SnO2 nanopowder and sensing film with 2 wt% La
EP

(P-2La and S-2La): (a) survey scan (b) Sn 3d, (c) La 3d and (d) O 1s.
CC

<Figure>Fig. 7. SEM images (cross-sectional view) of (a) S-0, (b) S-0.1La, (c) S-0.2La, (d) S-
A

0.5La, (e) S-1La and (f) S-2La sensing films on Au/Al2O3 substrates after annealing and sensing

test.
19
 

<Figure>Fig. 8. The selectivity histogram of sensor responses to various gases including toxic

gases (5 ppm NO2, 10 ppm H2S, 500 ppm SO2), and VOCs (400 ppm C3H6O, C6H6, CH2O,

C7H8, C8H10) at the optimal operating temperature 350°C of La2O3-doped SnO2 sensors with

different La-contents.

PT
<Figure>Fig. 9. Histograms of sensor response to 400 ppm of C3H6O of FSP-made La2O3-doped

RI
SnO2 sensors with different La contents (S-0 to S-2La) at various operating temperatures ranging

SC
from 150 to 400°C.

<Figure>Fig. 10. The change in resistance of SnO2 sensing films with different La

U
concentrations (S-0 to S-2La) under exposure to various C3H6O concentrations in (a) the high
N
range (6–400 ppm) and (b) the low range (0.1–4 ppm) and (c) correlation of sensor response
A
(solid line, left axis) and response time (dash line, right axis) vs. acetone concentration of S-0 to
M

S-2La sensors at 350°C.


D

<Figure>Fig. 11. Acetone-sensing mechanisms of SnO2 nanoparticles with (a) no La2O3 doping
TE

and with La2O3 doping at (b) low, (c) moderate and (d) high concentrations.

<Figure>Fig. 12. The energy band diagram of La2O3/SnO2 (p/n) heterojunction (a) before
EP

contact and (b) after contact at thermal equilibrium.


CC

<Figure>Fig. 13. Current (I)-Voltage (V) characteristics of SnO2 sensing films with different La

concentrations (S-0 to S-2La) (a) in dry air and (b) under exposure to 4 ppm C3H6O. Insets: I vs.
A

V in log-log scale under forward bias (upper left) and reverse bias (lower right) at 350C.

<Table>Table 1. Gas-Sensing performances of SnO2 with various additives and other advanced

materials prepared by different methods toward VOCs.


20
 

Sensing materials Synthesis Gases Temp. Response, Conc. References

4 at% La-SnO2 Sol-gel Eth 350C ∼70 % , 10 ppm Kundu et al.[4]


∼53 % , 1 ppm
5 wt% La-SnO2 Hydrothermal Eth 200C 213, 100 ppm Shi et al.[5]
nanorods
898, 500 ppm
La-SnO2 nanobelts Thermal Act 2.28, 100 ppm Wu et al.[6]

PT
230C
evaporation Eth 3.75, 100 ppm
Ethanediol 8.76, 100 ppm

RI
5 at% La-SnO2 thin Sol–gel Act 300C 59, 500 ppm Ang et al.[7]
film Eth 57, 500 ppm
Methanol 55, 500 ppm

SC
3 mol% La-SnO2 Co-precipitation Act 300C 3.3, 100 ppm Chikhale et al. [8]
nanoparticles Eth, LPG Selectivity, Acetone
Ammonia

U
Y-doped SnO2 Thermal Act 210C 15.8, 400 ppm Li et al.[9]
nanobelts

Eu-doped SnO2
evaporation
Electrospinning Act
N
280C 127, 300 ppm Jiang et al.[10]
A
nanofibers
0.1 wt% Ni-SnO2-5 FSP / Act 350C 169.7, 200 ppm Singkammo et al.[11]
M

wt% graphene electrolytic


exfoliation
Cu/Zn oxides Sputtering/ Eth 350C 34%, 100 ppm Rambu et al.[12]
D

multilayer Annealing
Cu Ferrite Sol-gel Act, Eth 300C ∼46, 42 %, 150 ppm Rezlescu et al.,
TE

Cd Ferrite selfcombustion Act, Eth 350C ∼36, 43 %, 150 ppm Tudorache et


al.[13,14]
Zn Ferrite Act, Eth 350C ∼13, 37 %, 150 ppm
Ni Ferrite Act, Eth 250C ∼14, 17 %, 150 ppm
EP

Ni0.75Co0.25 Ferrites Sol-gel Act, Eth 250C ∼61, 56 %, 150 ppm Tudorache et al.[15]
Yb Ferrites Sol-gel Act 230C 9.5, 3 ppm Zhang et al.[16]
CC

ZnO–Fe3O4 inverse Sol-gel with Act 485C 2, 0.8 ppm Zhang et al.[17]
opal opal template
Pani–Mg Ferrites Sol-gel Act, Eth 200C ∼45, 31 %, 150 ppm Tudorache et al.[18]
A

0.5 wt% La-SnO2 FSP Act 350C 2,735, 200 ppm Present work
nanoparticles 3,626, 400 ppm
Act=Acetone; Eth=Ethanol, Response = Ra/Rg or 100(Ra-Rg)/Ra in % where Ra and Rg = Resistance in air and gas.

<Table>Table 2. Parameters for theoretical calculation of energy band positions of SnO2 and

La2O3.
21
 

Semicon- El. Elemental Ionization Elemental Semicon- Conduction Valence Energy


ductor electron energy, Ei electro- ductor band Energy, gap, Eg
affinity, (eV) negativit, electro- energy, Ec Ev (eV) (eV)
Eea (eV) Xe (eV) negativity, (eV)
X (eV)

SnO2 6.21 -0.14 3.56 3.7


Sn 1.11 7.34 4.23

PT
O 1.46 13.6 7.53
La2O3 5.23 -2.17 3.63 5.8
La 0.47 5.58 3.03

RI
Diff () 0.98 2.03 0.07 2.1

Note: Xe= (Ei +Eea)/2, X= geometric mean of Xe1 and Xe2 where Xe1 and Xe2 are Xe of the first and second elements.

SC
Ec = X – Ee – 1/2Eg, where Ee = Energy of free electrons on the hydrogen scale (4.5 eV). EV = Ec+Eg

TDENDOFDOCTD

U
N
A
M
D
TE
EP
CC
A

Das könnte Ihnen auch gefallen