Chen2017 - Metodo Hydrothermal-2

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Sensors and Actuators B 238 (2017) 491–500

Contents lists available at ScienceDirect

Sensors and Actuators B: Chemical


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

Acetone sensing performances based on nanoporous TiO2 synthesized


by a facile hydrothermal method
Nan Chen a , Yuxiu Li b , Dongyang Deng b , Xu Liu b , Xinxin Xing b , Xuechun Xiao b,c ,
Yude Wang a,c,∗
a
Department of Physics, Yunnan University, 650091 Kunming, People’s Republic of China
b
School of Materials Science and Engineering, Yunnan University, 650091 Kunming, People’s Republic of China
c
Yunnan Province Key Lab of Micro-Nano Materials and Technology, Yunnan University, 650091 Kunming, People’s Republic of China

a r t i c l e i n f o a b s t r a c t

Article history: Nanoporous titanium dioxide was synthesized by a hydrothermal method without using of surfactant or
Received 15 April 2016 template. The structure, morphology, surface chemical states and specific surface area were characterized
Received in revised form 6 July 2016 by X-ray diffraction (XRD), transmission electron microscopy (TEM), X-ray photoelectron spectroscopy
Accepted 19 July 2016
(XPS), and N2 adsorption-desorption isotherms, respectively. The as-synthesized products are anatase-
Available online 20 July 2016
TiO2 with small grain size (about 12.27 nm) and high surface area (147.17 m2 g−1 ). The as-synthesized
porous TiO2 powder was used to fabricate indirect-heating gas sensor whose gas-sensing characteristics
Keywords:
toward acetone were investigated. At its optimal operation temperature, the sensor possesses a good
Nanoporous TiO2
Hydrothermal method
sensitivity, selectivity, linear dependence, low detection limitation, and response/recovery, repeatability
Surfactant-free as well as long-term stability. Especially for the high sensitivity and fast response/recovery, its response
Gas sensor reaches 25.97 for 500 ppm acetone, which is several times higher that of the reported TiO2 -based sensors.
Gas response The response and recovery times are only 13 and 8 s, respectively. Those values demonstrate the potential
Acetone of using as-synthesized TiO2 for acetone gas detection, particularly in the dynamic monitoring. Apart from
these, the mechanism related to the advanced properties was also investigated and presented.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction effumability also increases the risk. On the other hand, acetone was
known to be the final product for other ketone bodies’ metabolism
With the development of the world’s economy, issues of envi- [3]. In blood, acetone may be freely transported through alveolus
ronment become the subject of widespread concern. Volatile walls and easily mix with alveolus air. Hence, the acetone concen-
organic compounds (VOCs) are primary sources of environmental tration in the breath exhaled by diabetics is proportional to the
pollutants and considered seriously harmful to human body. For acetone content in the blood. Thus its concentration in blood may
example, acetone (CH3 COCH3 ), a widely used chemical reagent in be used as a measure of ketoacidosis. Therefore, fabricating acetone
industrial processes, can anesthesia human central nervous system, gas sensor with high sensitivity possesses important meaning for
which would cause a series of negative influence. When the concen- the field of occupational safety and human health. And the key for
tration is higher than 173 ppm, it can cause damages to eyes, noses, the reliable gas sensor is the sensing material.
and central nervous system [1]. Human’s exposure to high levels of As an important, wide-energy-gap (Eg = 3.0–3.4 eV) semicon-
acetone may cause respiratory irritation, mood swings, and nausea. ductor, titanium dioxide (TiO2 ) has been intensively studied as a
Likewise, breathing high levels of acetone (in industrial area) can key material for fundamental research and technological appli-
cause respiratory tract irritation, dizziness, and loss of strength [2]. cations in the fields of semiconductors, lithium-ion batteries
Meanwhile, acetone gas also shows extreme flammability. Explo- [4–7], photocatalytic decomposition [8–10] and solar cell [11–13],
sion or flash fire may occur with volume content between 2.5% and because of its good chemical stability, non-toxicity, abundance and
12.8% at a temperature higher than its flash point (−20 ◦ C). The high low cost [14]. It is also reported that TiO2 , as a semiconductor, also
shows certain gas sensing character, especially for acetone. How-
ever, the sensitivity, even for acetone, is still un-ideal. For instance,
∗ Corresponding author at: Department of Physics, Yunnan University, 650091 Bhowmik et al. [15] reported the response of the sensor based on
Kunming, People’s Republic of China. TiO2 nanotubes only reaches 3.35 toward 1000 ppm acetone. Deng
E-mail addresses: ydwang@ynu.edu.cn, wangyd99@163.com (Y. Wang). et al. [16] reported that the response of the gas sensor fabricated

http://dx.doi.org/10.1016/j.snb.2016.07.094
0925-4005/© 2016 Elsevier B.V. All rights reserved.
492 N. Chen et al. / Sensors and Actuators B 238 (2017) 491–500

from TiO2 nanofibers can not reach 3 toward 50 ppm acetone. From 10◦ to 90◦ (2␪) in steps of 0.01◦ . Transmission electron microscopy
the mentioned limitation of the concentration, the reported sensi- (TEM) measurement was performed on a Zeiss EM 912  instru-
tivity obviously can not meet the practical applications. Hence, a ment at an acceleration voltage of 120 kV, while high-resolution
further design for higher sensitivity is urgently needed. transmission electron microscopy (HRTEM) characterization was
As we know, the gas sensing of metal oxide sensors generally done using JEOL JEM-2100 Electron Microscope (with an acceler-
are based on the catalytic reaction between the target gas and the ation voltage of 200 kV). The samples for TEM were prepared by
adsorbed oxygen on the surface of the sensing materials [17]. That is dispersing the final dry samples in ethanol, and this dispersing was
to say, the gas sensing property highly relies on the exposed surface then dropped on carbon–copper grids covered by an amorphous
area. In another words, a high specific surface area could provide carbon film. The nitrogen adsorption isotherm was measured at
more active sites and consequent higher sensitivity. From this point 77.3 K with a Micromeritics ASAP 2010 automated sorption ana-
of view, fabricating porous TiO2 could be a good choice to enhanc- lyzer. Prior to the measurement, the sample was degassed at 300 ◦ C
ing the sensitivity toward acetone for practical use. Pang et al. [18] for 6 h under a vacuum. X-ray photoelectron spectroscopy (XPS)
prepared porous TiO2 nanobelts by an alkali hydrothermal process was carried out at room temperature in an ESCALAB 250 system.
with a temperature of 200 ◦ C for 72 h, then immersing the prod- During XPS analysis, an Al K␣ X-ray beam was adopted as the
uct in HCl aqueous for 72 h and finally annealing at 600 ◦ C for 3 h. excitation source and the vacuum pressure of the instrument cham-
The waste sludge templated TiO2 with a surface area of 130 m2 g−1 ber was 1 × 10−7 Pa as read on the panel. Measured spectra were
was synthesized through a calcination process [19]. Yu et al. [20] decomposed into Gaussian components by a least-square fitting
synthesized porous TiO2 films with the surface area ranged from method. Bonding energy was calibrated with reference to the C1s
35 to 43 m2 g−1 by a complicated procedure with template of sil- peak (284.6 eV).
ica. Obviously, these processes were multistep, long-playing and
power-wasting. For practicing purpose, the economy effect also
2.3. Preparation and test of gas sensor
should be taken into consideration. In addition, Lechuga et al. [21]
reported that anionic and non-ionic surfactant had an acute toxicity
The fabrication of indirect-heating structure sensor was
to aquatic organisms. Hence, a simple, convenient and surfactant-
described in the literature [22,23]. TiO2 oxide was mixed with
free method to porous TiO2 is urgent to appear.
deionized water to form paste, and then coated onto the outside
In this paper, nanoporous TiO2 with a high surface area of
of an alumina tube (4 mm in length, 1.2 mm in external diameter,
147.17 m2 g−1 was obtained by a simple hydrothermal method
and 0.8 mm in internal diameter) with a pair of Au electrodes and
without using of any surfactant or template. The nanoporous TiO2
platinum wires installed at each end. The thickness of the sensitive
was then used as a sensing material for indirect heating sensor. Its
body, which was dried and calcined in air at 400 ◦ C for 2 h, was about
acetone gas sensing properties were measured, and a remarkable
0.5 mm. A Ni–Cr alloy wire crossing the alumina tube was used as a
enhanced sensitivity toward acetone that the reported TiO2 -based
resistor to ensure both substrate heating and temperature control.
sensor is gained. To get further understanding on the related mech-
In order to improve their stability and repeatability, the gas sen-
anism including formation and enhanced acetone gas sensing,
sor was aged at operating temperature 320 ◦ C for 120 h in air. The
X-ray diffraction (XRD), transmission electron microscopy (TEM),
sensor’s resistance was measured by using a conventional circuit in
X-ray photoelectron spectroscopy (XPS), and the Brunner-Emmett-
which the element was connected with an external resistor in series
Teller (BET) were carried out.
at a circuit voltage of 5 V. Then, the sensor was well connected to
a bakelite base through platinum wires to perform electrical mea-
2. Experimental details
surements using a WS-30A system, which structure was shown
in Fig. S1. During the testing process, the needful amounts of the
2.1. Preparation of nanoporous TiO2
target substance were injected into the chamber by a microinjec-
tor when the resistance of the sensor was stable. The liquid was
All the chemical reagents used in the experiments were obtained
evaporated quickly in the chamber (18 L in volume). The electrical
from commercial sources as guaranteed-grade reagents and used
response of the sensor was measured with an automatic test sys-
without further purification.
tem, controlled by a personal computer. The gas response ␤ was
Nanoporous TiO2 was prepared by a simple low temperature
defined as the ratio of the electrical resistance in air (Ra ) to that in
hydrothermal method. The reaction can be described as follows:
gas (Rg ), namely ␤ = Ra /Rg [24,25]. In addition, the response time
TiOSO4 + 2H2 O → TiO(OH)2 + H2 SO4 (1) was defined as the time required for the gas response to reach 90%
of the final equilibrium value after a test gas was injected, and the
dehydrate
TiO(OH)2 → TiO2 + H2 O (2) recovery time was the time needed for gas response to decrease by
90% after the gas sensor was exposed in air again.
In a typical synthesized experiment, 4.899 g titanyl sulfate was
added to 50 mL deionized water with stirring until a homogenous
solution was obtained. Then the solution was transferred into a 3. Results and discussion
Teflon-lined stainless steel autoclave with a capacity of 80 mL and
reacted under hydrothermal conditions at a temperature of 180 ◦ C The structural features of the as-synthesized nanoporous TiO2
for 4 h. The autoclave was cooled down to room temperature in particles were analyzed by XRD. The initial assignment was fur-
a standard atmosphere. The resulting products were centrifuged, ther confirmed by the refinement of the diffraction pattern with
and the white precipitates were thoroughly washed with deionized the Rietveld method. The experimental pattern, together with
water and dried at 60 ◦ C. the calculated pattern obtained from the Rietveld refinement and
difference profile are shown in Fig. 1. The structural parame-
2.2. Characterization of as-synthesized nanoporous TiO2 ters calculated from the Rietveld profile refinement are presented
in Table 1. The difference curve between the calculated and
X-Ray diffraction (XRD, Rigaku D/MAX-3B powder diffractome- experimental XRD patterns reveals an excellent agreement. The
ter) with a copper target and K␣1 radiation (␭ = 1.54056 Å) was used experimental diffraction peaks can be perfectly indexed to anatase
for the phase identification, where the diffracted X-ray intensities TiO2 (JCPDS No. 21-1272), space group: I41 /amd (141). The results
were recorded as a function of 2␪. The sample was scanned from indicated the high purity of the obtained product. The volume-
N. Chen et al. / Sensors and Actuators B 238 (2017) 491–500 493

tern. Fig. 2(c) and (d) are the morphologies of as-synthesized TiO2
after calcined at 400 ◦ C for 2 h, while based sensor aged at operating
temperature 320 ◦ C for 120 h in air are shown in Fig. 2(e) and (f).
It can be seen that the morphologies of the samples after calcina-
tion and “aging” have no obviously variations compared with the
untreated sample. Fig. 2(d) and (f) show the nanostructures with
an interplanar spacing of about 0.352 nm of the samples after cal-
cination and “aging”, respectively, which is close to the value of the
(101) lattice planes of anatase TiO2 .
Based on the above results and analysis, a probable mecha-
nism of nanoporous TiO2 formation was proposed in Fig. 3. At this
temperature, TiO2+ ions from TiOSO4 aqueous solution begin to
hydrolyze and TiO(OH)2 crystalline cores are formed. Then the crys-
talline cores grow to a certain dimension. At the same time, due to
the special pressure and the temperature of hydrothermal method,
the TiO(OH)2 nanocrystal dehydrate and company with an irregu-
lar contraction, leading to the formation of the nanoporous TiO2 .
However, in a homogenous solution under hydrothermal condi-
tions, the forming mechanism of nanoporous TiO2 is very complex
and difficult to understand. However, it is very important to control
the porous size and surface area of the nanoporous TiO2 . Therefore,
Fig. 1. X-ray diffraction analysis of as-synthesized TiO2 . The experimental data is the further research work will be focus on the forming mechanism
shown in red, the calculated patterns in black, and the difference curves in blue. of nanoporous TiO2 to us.
The short vertical bars in green represent the positions of the Bragg reflections. (For
Nitrogen adsorption-desorption isotherms measurements were
interpretation of the references to colour in this figure legend, the reader is referred
to the web version of this article.) performed to detect the surface adsorption properties of sam-
ple. The N2 adsorption-desorption isotherms are shown in Fig. 4.
The isotherm can be categorized as type IV with small hystere-
Table 1
Structure data and refinement parameters for nanoporous TiO2 calculated by sis loops observed at a relative pressure of 0.45–0.95. The specific
Rietveld refinement of the experimental XRD powder pattern. surface area estimated from the BET method was 147.17 m2 g−1 ,
showing the porous structure of the sample. The specific surface
Space group I41 /amd (141)
area of as-synthesized nanoporous TiO2 is higher than the values
Lattice parameter
reported in other reported TiO2 , such as rutile TiO2 (90.7 m2 g−1 )
a (Å) 3.7946
c (Å) 9.4998
[6], TiO2 microspheres (67 m2 g−1 ) [8], spherical mesoporous TiO2
Unit Cell Volume (Å3 ) 136.787 (92.57 m2 g−1 ) [9], organized mesoporous TiO2 films (107.9 m2 g−1 )
[13], TiO2 layer (71.4 m2 g−1 ) [16], mesoporous TiO2 nanoparticles
Ti
x 0
(88.5 m2 g−1 ) [26], anatase TiO2 nanosheets (113.1 m2 g−1 ) [27],
y 0 TiO2 nanoplates (107.7 m2 g−1 ) [28]. Inset of Fig. 4 illustrates its
z 0 corresponding pore size distributions obtained from desorption
O branches. It can be concluded from the pore size distributions that
x 1 as-synthesized TiO2 have pores with diameter of about 3.59 nm.
y 0 In virtue of the porous structure and high surface area, the as-
z 0.2113 synthesized nanoporous TiO2 would provide more active sites to
Average apparent size (nm) 12.27
Rwp (%) 34.96
get in touch with acetone gas. In addition, the surface area of as-
Rp (%) 26.44 synthesized TiO2 after calcination and “aging” are 143.36 m2 g−1
and 141.12 m2 g−1 , respectively. It seems that the surface area of the
as-synthesized TiO2 have no obvious change after heat treatment.
weighted average crystallite size calculated from the Rietveld The surface/near surface chemical compositions of as-
profile refinement is 12.27 nm. Moreover, in section 2.3, we men- synthesized TiO2 were analyzed by XPS, as shown in Fig. 5.
tioned that the thickness of the sensitive body was dried and Apart from the C1 s peak at 284.6 eV, only the peaks related to Ti
calcined in air at 400 ◦ C for 2 h. Hence, the structural features of and O are observed, indicating a good purity of the as-synthesized
the sample after calcination were also analyzed by XRD, which was samples. The high-resolution XPS spectrum of O1 s is shown in
shown in Fig. S2. One can see that the experimental data are in Fig. 5(a). One can find that there are two kinds of oxygen in the
agreement with anatase TiO2 (JCPDS No. 21-1272), indicating that surface: the peak centers at 530.26 eV indexes lattice oxygen
the structural features have no changes before and after calcination. (Olattice ) while the peak at 532.21 eV presents adsorbed oxygen
The morphologies of as-synthesized TiO2 , the as-synthesized (Oads ) [29,30]. Olattice is attributed to the oxygen ions in the
TiO2 calcined at 400 ◦ C for 2 h and “aged” at 320 ◦ C for 120 h crystal lattice which are is thought to be pretty stable and has no
were performed by TEM, as shown in Fig. 2. The as-synthesized contribution to the gas response, meanwhile, Oads is attributed
nanoporous TiO2 was easily found from Fig. 2(a). One can observe to the absorbed oxygen ions, which has a very important role in
that the as-synthesized TiO2 has a rather uniform analogous shape the gas sensing property [31]. Through calculating the area of Oads
as well as size. To get further insight into the atomic order of the and Olattice emission lines, the concentrations of Olattice and Oads to
nanoporous TiO2 , high-resolution images were recorded. As shown O 1s are estimated to be 37.75% and 62.25%, respectively. While
in Fig. 2(b), it indicates that the nanostructures is structurally uni- to Degussa P25, the concentrations of Olattice and Oads to O 1 s
form with an interplanar spacing of about 0.352 nm, which is close are estimated to be 75.78% and 24.22%, respectively (Fig. S3). The
to the value of the (101) lattice planes of anatase TiO2 . The crystal- Oads of as-synthesized TiO2 is 2.57 times higher than that of P25,
lite size of the nanoporous TiO2 particles is around 12 nm, which is demonstrating the adsorbed oxygen is abundant to as-synthesized
in good agreement with the particle size obtained from the XRD pat- nanoporous TiO2 . In Fig. 5(b), Ti 2p spectrum reveals the existence
494 N. Chen et al. / Sensors and Actuators B 238 (2017) 491–500

Fig. 2. TEM (a) and HRTEM (b) images of nanoporous TiO2 ; (c) and (d) are the images of nanoporous TiO2 after calcined at 400 ◦ C for 2 h, and (e) and (f) are for images of
nanoporous TiO2 and aging at 320 ◦ C for 120 h, respectively.

of two peaks of Ti 2p1/2 and Ti 2p3/2 at the position of 464.7 eV and TiO2 materials was analyzed at working temperatures ranging
459.1 eV. The distance between these two peaks is equal to 5.6 eV, from 260 ◦ C to 400 ◦ C at acetone ambient with a concentration
which is in good agreement with the energy reported for TiO2 : the of 500 ppm, while gas response of the sensor fabricated with
values correspond to the 2p binding energy of Ti(IV) ions [4]. Degussa P25 was also analyzed at working temperatures ranging
To evaluate the potential applicability of the fabricated gas from 260 ◦ C to 400 ◦ C at acetone ambient with a concentration of
sensor for acetone detection, some fundamental gas sensing 1000 ppm. As shown in Fig. 6. It is obviously seen that the gas
parameters were investigated. The operating temperature is one response value of the as-synthesized nanoporous TiO2 based sensor
of the most important factors for the gas sensor, which highly increases up to 370 ◦ C and reaches the highest, then decreases over
determined the nature of the sensing materials and the gas-sensing 370 ◦ C. Why this case happened can explain as following. Obviously,
process between the gas and the surface of materials. Normally, the at a low operating temperature, the low gas response (␤ = Ra /Rg )
first approach is to select the optimum temperature of gas sen- can be obtained for the target gas acetone do not have enough
sor. The gas response of the sensor fabricated with nanoporous thermal energy to react with the surface electron of as-synthesized
N. Chen et al. / Sensors and Actuators B 238 (2017) 491–500 495

Fig. 3. Schematic illustrations for the growth process of the as-synthesized nanoporous TiO2.

Fig. 4. Nitrogen adsorption-desorption isotherms of as-synthesized nanoporous


TiO2 , and the inset is the corresponding pore size distributions.

nanoporous TiO2 , which leads to a low response. With the oper-


ating temperature increasing, the thermal energy obtained is high
enough to overcome the activation energy of the surface reaction
[32,33]. Moreover, the reduction in gas response after the maxi-
mum is due to the low gas adsorption ability of the gas molecule
at high temperature causes the low utilization rate of the sens-
ing material, which is the reason for the reduction in gas response
[34–36]. What’s more, response of the sensor fabricated by Degussa
P25 is obviously inferior to the nanoporous TiO2 , while its response
only can reach 3.87 for 1000 ppm acetone at the temperature of
400 ◦ C. Therefore, 370 ◦ C was chosen as the optimum tempera-
ture to the sensor based on as-synthesized nanoporous TiO2 , and
monitored different gases.
The gas responses toward different gas concentration from Fig. 5. The high-resolution XPS survey spectra of O 1s (a) and Ti2p (b) of as-
20 ppm to 1500 ppm at the optimum temperature are shown in synthesized TiO2 .
Fig. 7. As shown in Fig. 7(a), the response value of all sensors
increases along with the increased concentration of VOCs (acetone, tone, methanol, formaldehyde, ethanol, isopropanol and n-butanol,
methanol, formaldehyde, ethanol, isopropanol and n-butanol). The their response values are 25.97, 9.86, 3.53, 11.19, 5.35 and 7.56
straight lines are calibration curves and experimental data which respectively which confirms the nanoporous TiO2 based sensor has
were fitted as: comparatively favorable selectivity to acetone. In the front, we have
ˇ = mC gas + k (3) mentioned that Oads has a very important role in the gas sensing
property. When our sensor exposes to air, the adsorbates such as
where m is the gas response coefficient, Cgas is the gas concentra- − , O− , O2− are formed by capturing electrons from the conduc-
O2ads ads ads
tion, and k is a constant. The correlative coefficients R2 are higher tance band [37], which builds a space-charge region on the surface
than 0.95 indicating a good linear dependence. Fig. 7(b) shows the of the metal oxide grains, resulting in an electron-depleted surface
gas response value of the sensor toward different targeted gases layer. As gases reach the surface of the material, they absorb energy
with the same concentration of 500 ppm at 370 ◦ C. While to ace- and become activated. Then, they react with the particles absorbed
496 N. Chen et al. / Sensors and Actuators B 238 (2017) 491–500

Fig. 6. Gas responses of the sensor based on (a) as-synthesized TiO2 at different
operating temperature towards 500 ppm acetone, and (b) P25 at different operating
temperature towards 1000 ppm acetone.

on the material surface, following the reaction rate described by


the Arrhenius equation [38]:

E
r = C exp(− ) (4)
kT
where r is the reaction rate constant, C is the pre-exponential factor,
E is the activation energy, k is the Boltzmann constant, and T is the
temperature.
The reaction rate constant (r) increases with a decreasing acti-
vation energy (E). In other words, the rate of response becomes
higher by lowering the activation energy (E) at the same operat-
ing temperature. Based on the above discussion, at the optimum
temperature of 370 ◦ C, acetone gas perhaps has a suitable activa-
tion energy and reaction rate constant increases, resulting in the
high response. While to other gases, their activation energies may
be higher than acetone, leading to the low reaction rate constant.
Or 370 ◦ C is a high temperature that gases disintegrated when they
reached the surface of the sensor.
Fig. 8 displays a dynamic response-recovery curve of the
nanoporous TiO2 toward different acetone concentration from
20 ppm to 1500 ppm. A good response-recovery property can be
observed. The gas response keeps almost constant with only small
Fig. 7. The gas-sensing properties of the sensor fabricated from nanoporous TiO2
fluctuations when it reaches its dynamic balance in both air and at an optimum temperature of 370 ◦ C. (a) Variation of gas response to different gas
acetone; meanwhile, the gas response also rapidly increases and (acetone, methanol, formaldehyde, ethanol, isopropanol, n-butanol) concentrations
decreases in response and recovery situations, respectively. More from 20 to 1500 ppm, (b) the response of gas sensor to various gases (500 ppm) at
specifically, the response and recovery time toward 500 ppm ace- 370 ◦ C.

tone are 13 and 8 s, respectively. According to the previous analysis,


the high surface area makes a contribution to the quick response and the service length were determined. To verify the stability of
and recovery time, and permits the nanoporous TiO2 based gas the sensor, the gas responses toward 500 ppm acetone over 28 days
sensor to exhibit the outstanding adsorption and desorption per- were tested at its optimal temperature. As shown in Fig. 10, the
formance. gas response evolution shows that the responses only have a small
In addition, repeatability is also one important parameter which fluctuation, which is below 3.71% of its initial value. This illustrates
can be used to evaluate the reliability of a fabricated sensor. To ver- a good stability of the gas sensor.
ify the repeatability of the sensor, the gas response evolutions in 7 A brief summary of the sensing performances of various TiO2
cycles were tested toward 500 ppm acetone at 370 ◦ C. Fig. 9 illus- nanostructures based gas sensor toward acetone are shown in
trates the reproducibility of the sensor based on the as-prepared Table 2. As it has been introduced, the gas responses of TiO2 or
porous TiO2 , revealing that the sensor maintains its initial response some of its composites toward acetone are poor [39–48]. Moreover,
amplitude without a clear decrease upon 7 cycles successive sens- most of the reported sensors show a bad response/recovery prop-
ing tests at 500 ppm of acetone. erty, which seldom reported accurate values. For example, Rella
In practical applications, the long-term stability of a gas sensor et al. [39] reported the response/recovery time for TiO2 nanopar-
has attained much attention for which the reliability of gas sensors ticles towards 200 ppm acetone are 240 s and 180 s, respectively.
N. Chen et al. / Sensors and Actuators B 238 (2017) 491–500 497

Table 2
Comparison of varied titanium oxide nanostructures in acetone sensing performances.

Materials Operating temperature (◦ C) ˇ Response/Recovery time (s) Reference

Nanoporous 370 2.22@20 ppm 18/9 This


TiO2 4.93@50 ppm 16/13 work
7.44@100 ppm 14/10
11.3@200 ppm 14/6
25.97@500 ppm 13/8
TiO2 nanoparticles 400 7.5@200 ppm 240/180 [39]
Surface-coarsened TiO2 nanobelts 350 4.6@300 ppm – [40]
TiO2 nanobelts 2.9@300 ppm –
TiO2 nanorods 500 20@500 ppm 14/8 [41]
TiO2 400 0.75@100 ppm – [42]
TiO2 -V2 O5 300 9.5@100 ppm
Pure TiO2 5.6@400 ppm 147/
1 at% Nb/TiO2 8.1@400 ppm 9/ [43]
3 at% Nb/TiO2 400 13@400 ppm 33/
5 at% Nb/TiO2 8.0@400 ppm 126/
TiO2 350 0.5@100 ppm – [44]
TiO2 -WO3 (L) 12@100 ppm
TiO2 thin film 300 0.35@25 ppm – [45]
TiO2 /NiO 99/1 0.45@25 ppm
TiO2 /NiO 95/5 0.70@25 ppm
TiO2 /NiO 90/10 15.0@25 ppm
TiO2 thin film 150 1.15@50 ppm 14/22 [46]
TiO2 nanotubes 150 2.08@103 ppm 21/38 [47]
Ni-TiO2 nanotubes 100 5.56@103 ppm 31/25
Pd-TiO2 nanotubes 100 33.3@103 ppm 45/18
TiO2 nanorod 400 5.82@200 ppm – [48]
NiO-TiO2 nanorod 9.33@200 ppm

Fig. 8. Dynamic response/recovery curve of the sensor based on porous TiO2 to Fig. 9. The reproducibility of nanoporous TiO2 to acetone with a concentration of
different acetone concentration from 20 to 1500 ppm at the operating temperature 500 ppm by 7 cycles at operating temperature of 370 ◦ C.
of 370 ◦ C.

in Fig. 11, which is similar as our reported work [50]. In air atmo-
The response time for 5% Nb doped TiO2 nanoparticles towards sphere, oxygen adsorbs onto the surface of nanoporous TiO2 , and
400 ppm acetone is about 126 s [43]. Bhattacharyya et al. [47] electron transfers from conduction band to the oxygen molecules
reported the response time for Pd modified TiO2 nanotubes towards to form various kinds of oxygen ions with different valence states
1000 ppm acetone is 45 s. The slow response and recovery time nat- − , O− , O2− ), which leads to the formation of a thick space-
(O2ads ads ads
urally cannot meet the need of real-time dynamic measurement. charge layer and a consequent high resistance of the sensor. This
But the as-synthesized nanoporous TiO2 only needs 13 and 8 s to process can be described by the following equations [50–52]:
response and recovery with 500 ppm acetone. Furthermore, the fast
response/recovery time might result from the high operation tem- O2gas ↔ O2ads (5)
perature and surface area, which also is higher than the reported

values. O2ads + e− ↔ O2ads (6)
TiO2 is a typical n-type semiconductor, which gas sensing per- − −
formance depends on the change of surface occupation. According O2ads + e− ↔ 2Oads (7)
to Wolkenstein s model for semiconductors [49], we propose an −
analogous model for the nanoporous TiO2 , as schematically shown Oads + e− ↔ Oads
2−
(8)
498 N. Chen et al. / Sensors and Actuators B 238 (2017) 491–500

4. Conclusions

In conclusion, nanoporous TiO2 was successfully synthesized


by a facile hydrothermal method without using of surfactant or
template. Owing to the small size, as-synthesized nanoporous TiO2
possesses a high specific surface area (147.17 m2 g−1 ). It is found
that as-synthesized nanoporous TiO2 has a high concentration of
Oads , which can attribute to enhance the gas response of TiO2 to ace-
tone. The sensor fabricated from nanoporous TiO2 also processes
excellent response/recovery time, good linear dependence, certain
selectivity as well as repeatability and long-term stability at an
optimum temperature of 370 ◦ C, illustrating the potential of using
nanoporous TiO2 for acetone.

Acknowledgement

This work was supported by National Natural Science Founda-


tion of China (No. 51262029).

Appendix A. Supplementary data


Fig. 10. Dynamic response–recovery cycles of the as-synthesized gas sensor toward
500 ppm acetone gas at an operating temperature of 370 ◦ C. Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.snb.2016.07.094.
When the sensor is exposed to acetone, the reductive gas reacts
with the oxygen adsorbed on the sensor surface. Then the electrons
References
are released back to the conduction band of semiconductor, lead-
ing to a thinner space-charge layer and a lower potential barrier [1] Q.Q. Jia, H.M. Ji, Y. Zhang, Y.L. Chen, X.H. Sun, Z.G. Jin, Rapid and selective
[52]. This process results in a decrease in the resistance and can be detection of acetone using hierarchical ZnO gas sensor for hazardous odor
markers application, J. Hazard. Mater. 276 (2014) 262–270.
expressed by the following equations [41,53]:
[2] Acetone: Health Information Summary, New Hampshire, Department of
Environmental Service 2005.
CH3 COCH3 (gas) ↔ CH3 COCH3 (ads) (9)
[3] N. Makisimovich, V. Vorotyntsev, N. Nikitina, O. Kaskevich, P. Karabun, F.
Martynenko, Adsorption semiconductor sensor for diabetic ketoacidosis
CH3 COCH3 (gas) + 8O− → 3CO2 (gas) + 3H2 O(gas) + 8e− (10) diagnosis, Sens. Actuators B: Chem. 35–36 (1996) 419–421.
− − [4] Y.D. Wang, T. Chen, Q.Y. Mu, Electrochemical performance of W-doped
CH3 COCH3 (ads) + 4O2ads → 3CO2 (gas) + 3H2 O(gas) + 4e (11) anatase TiO2 nanoparticles as an electrode material for lithium-ion batteries,
J. Mater. Chem. 21 (2011) 6606–6613.
According to the analysis of the possible gas sensing mecha- [5] Y. Li, J.D. Luo, X.Y. Hu, X.F. Wang, J.C. Liang, K.F. Yu, Fabrication of TiO2 hollow
nism, one can find that there are abundant absorbed oxygen ions nanostructures and their application in Lithium ion batteries, J. Alloys
in the as-synthesized nanoporous TiO2 (in XPS section), which has Componds 651 (2015) 685–689.
[6] W.L. Wang, J.Y. Park, V.H. Nguyen, E.M. Jin, H.B. Gu, Hierarchical mesoporous
a great influence on the gas sensing properties, also indicates the rutile TiO2 /C composition nanospheres as lithium-ion battery anode
reasonability of the sensing mechanism. materials, Ceram. Int. 42 (2016) 598–606.
The high specific surface area (147.17 m2 g−1 ) and the high con- [7] Y. Cai, H.E. Wang, J. Jin, S.Z. Huang, Y. Yu, Y. Li, S.P. Feng, B.L. Su, Hierarchically
structured porous TiO2 spheres constructed by interconnected nanorods as
centration of Oads make a great contribution to the high response high performance anodes for lithium ion batteries, Chem. Eng. J. 281 (2015)
and fast response-recovery. As we introduced, only Oads plays a role 844–851.
in the in gas sensing performance, which attributes to the absorbed [8] L.Z. Yang, L.Z. Zhu, C. Liu, M. Fang, G.H. Liu, X.B. Yu, Synthesis and
photocatalytic property of porous TiO2 microspheres, Mater. Res. Bull. 43
oxygen ions and the gas sensing property is based on the reactions (2008) 806–810.
between gases and Oads . The high specific surface area brings a [9] B. Sun, G.W. Zhou, C.W. Shao, B. Jiang, J.L. Pang, Y. Zhang, Spherical
greater contact area between the gases and the materials, which mesoporous TiO2 fabricated by sodium dodecyl sulfate-assisted
hydrothermal treatment and its photocatalytic decomposition of
decreases the response and recovery time and assists in having
papermaking wastewater, Powder Technol. 256 (2014)
more Oads exposed to get in touch with gases. 118–125.

Fig. 11. A schematic diagram of the proposed reaction mechanism of the TiO2 -based sensor in air (a) and in acetone (b).
N. Chen et al. / Sensors and Actuators B 238 (2017) 491–500 499

[10] P. Esparza, T. Hernández, M.E. Borges, M.C. Álvarez-Galván, J.C. Ruiz-Morales, [37] S.R. Morrison, Selectivity in semiconductor gas sensors, Sens. Actuators B 12
J.L.G. Fierro, TiO2 modifications by hydrothermal treatment and doping to (1987) 425–440.
improve its photocatalytic behaviour under visible light, Catal. Today 210 [38] M.L. Zhang, Z.H. Yuan, J.P. Song, C. Zheng, Improvement and mechanism for
(2013) 135–141. the fast response of a Pt/TiO2 gas sensor, Sens. Actuators B: Chem. 148 (2010)
[11] P. Wang, F.L. He, J. Wang, H.G. Yu, L. Zhao, Graphene oxide nanosheets as an 87–92.
effective template for the synthesis of porous TiO2 film in dye-sensitized solar [39] R. Rella, J. Spadavecchia, M.G. Manera, S. Capone, A. Taurino, M. Martino, A.P.
cells, Appl. Surf. Sci. 358 (2015) 175–180. Caricato, T. Tunno, Acetone and ethanol solid-state gas sensors based on TiO2
[12] S.J. Kwon, H.B. Im, J.E. Nam, J.K. Kang, T.S. Hwang, K.B. Yi, Hydrothermal nanoparticles thin film deposited by matrix assisted pulsed laser evaporation,
synthesis of rutile–anatase TiO2 nanobranched arrays for efficient Sens. Actuators B: Chem. 127 (2007) 426–431.
dye-sensitized solar cells, Appl. Surf. Sci. 320 (2014) 487–493. [40] X.H. Wang, Y.H. Sang, D.Z. Wang, S.Z. Ji, H. Liu, Enhanced gas sensing property
[13] J.Y. Lim, C.S. Lee, J.M. Lee, J. Ahn, H.H. Cho, J.H. Kim, Amphiphilic block-graft of SnO2 nanoparticles by constructing the SnO2 -TiO2 nanobelt
copolymer templates for organized mesoporous TiO2 films in dye-sensitized heterostructure, J. Alloys Compounds 639 (2015) 571–576.
solar cells, J. Power Sources 301 (2016) 18–28. [41] H.Q. Bian, S.Y. Ma, A.M. Sun, X.L. Xu, G.J. Yang, J.M. Gao, Z.M. Zhang, H.B. Zhu,
[14] A.L. Linsebigler, G. Lu, J.T. Yates, Photocatalysis on TiO2 surfaces: principles Characterization and acetone gas sensing properties of electrospun TiO2
mechanisms, and selected results, Chem. Rev. 95 (1995) 735–758. nanorods, Superlattices Microstuct. 81 (2015) 107–113.
[15] B. Bhowmik, A. Hazra, K. Dutta, P. Bhattacharyya, Repeatability and stability of [42] M. Epifani, E. Comini, P. Siciliano, G. Faglia, J.R. Morante, Evidence of catalytic
room-temperature acetone sensor based on TiO2 nanotubes: influence of activation of anatase nanocrystals by vanadiumoxide surface layer: acetone
stoichiometry variation, IEEE Trans. Mater. Reliab. 14 (2014) 961–967. and ethanol sensing properties, Sens. Actuators B: Chem. 217 (2015) 193–197.
[16] J.A. Deng, L.L. Wang, Z. Lou, T. Zhang, Design of CuO-TiO2 heterostructure [43] S. Phanichphant, C. Liewhiran, K. Wetchakun, A. Wisitsoraat, A. Tuantranont,
nanofibers and their sensing performance, J. Mater. Chem. A 2 (2014) Flame-made Nb-doped TiO2 ethanol and acetone sensors, Sensors 11 (2011)
9030–9034. 472–484.
[17] Q. Li, N. Chen, X.X. Xing, X.C. Xiao, Y.D. Wang, I. Djerdj, NiO nanosheets [44] M. Epifani, E. Comini, R. Díaz, T. Andreu, A. Genç, J. Arbiol, P. Siciliano, G.
assembled into hollow microspheres for highly sensitive and fast-responding Faglia, J.R. Morante, Acetone sensing with TiO2 -WO3 nanocomposites: an
VOCs sensor, RSC Adv. 7 (2015) 80786–80792. example of response enhancement by inter-oxide cooperative effects,
[18] L.X. Pang, X.Y. Wang, X.D. Tang, Enhanced photocatalytic performance of Procedia Eng. 87 (2014) 803–806.
porous TiO2 nanobelts with Phase junctions, Solid State Sci. 39 (2015) [45] A. Wisitsoraat, A. Tuantranont, E. Comini, G. Sberveglieri, W. Wlodarski,
29–33. Characterization of n-type and p-type semiconductor gas sensors based on
[19] X.P. Wang, S.Q. Huang, N.W. Zhu, Z.Y. Lou, H.P. Yuan, Facial synthesis of NiOx doped TiO2 thin films, Thin Solid Films 517 (2009) 2775–2780.
porous TiO2 photocatalysts using waste sludge as the template, Appl. Surf. Sci. [46] B. Bhowmik, K. Dutta, A. Hazra, P. Bhattacharyya, Low temperature acetone
359 (2015) 917–922. detection by p-type nano-titania thin film: equivalent circuit model and
[20] H. Yu, L.Z. Wang, M. Dargusch, Low-temperature templated synthesis of sensing mechanism, Solid State Electron. 99 (2014) 84–92.
porous TiO2 single-crystals for solar cell applications, Sol. Energy 123 (2016) [47] P. Bhattacharyya, B. Bhowmik, H.J. Fecht, Operating temperature,
17–22. repeatability, and selectivity of TiO2 nanotube-based acetone sensor:
[21] M. Lechuga, M. Fernández-Serrano, E. Jurado, J. Núñez-Olea, F. Ríos, Acute influence of Pd and Ni nanoparticle modifications, IEEET Mater. Res. 15 (2015)
toxicity of anionic and non-ionic surfactants to aquatic organisms, Ecotox. 376–383.
Environ. Saf. 125 (2016) 1–8. [48] G.J. Sun, H. Kheel, S. Park, S. Lee, S.E. Park, C. Lee, Synthesis of TiO2 nanorods
[22] Y.D. Wang, C.L. Ma, X.H. Wu, X.D. Sun, H.D. Li, Mesostructured tin oxide as decorated with NiO nanoparticles and their acetone sensing properties,
sensitive material for C2 H5 OH sensor, Talanta 57 (2002) 875–882. Ceram. Int. 42 (2016) 1063–1069.
[23] Y.D. Wang, I. Djerdj, M. Amtonietti, B. Smarsly, Polymer-assisted generation of [49] T. Wolkenstein, Electronic Processes on Semiconductor Surfaces During
antimony-doped SnO2 nanoparticles with crystallinity for application in gas Chemisorption, Consultants Bureau, New York, 1991, pp. 35–182.
sensors, Small 4 (2008) 1656–1660. [50] M.H. Cao, Y.D. Wang, T. Chen, M. Antonietti, M. Niederberger, A high sensitive
[24] C.J. Dong, Q. Li, G. Chen, X.C. Xiao, Y.D. Wang, Enhanced formaldehyde sensing and fast-responding ethanol sensor based on CdIn2 O4 nanocrystals
performance of 3D hierarchical porous structure Pt-functionalized NiO via a synthesized by a nonaqueous sol-gel route, Chem. Mater. 20 (2008)
facile solution combustion synthesis, Sens. Actuators B: Chem. 220 (2015) 5781–5786.
171–179. [51] X. Liu, N. Chen, X.X. Xing, Y.X. Li, X.C. Xiao, Y.D. Wang, I. Djerdj, A
[25] D. Hu, B.Q. Han, S.J. Deng, Z.P. Feng, Y. Wang, J. Popovic, M. Nuskol, Y.D. Wang, high-performance n-butanol gas sensor based on ZnO nanoparticles
I. Djerdj, Novel mixed phase SnO2 nanorods assembled with SnO2 synthesized by a low-temperature solvothermal route, RSC Adv. 5 (2015)
nanocrystals for enhancing gas-sensing performance toward isopropanol Gas, 54372–54378.
J. Phys. Chem. C 118 (2014) 9832–9840. [52] X. Liu, N. Chen, B.Q. Han, X.C. Xiao, G. Chen, I. Djerdj, Y.D. Wang, Nanoparticle
[26] N. Wang, W.L. Fu, J. Zhang, X.R. Li, Q.H. Fang, Corrosion performance of cluster gas sensor: pt activated SnO2 nanoparticles for NH3 detection with
waterborne epoxy coatings containing polyethylenimine treated ultrahigh sensitivity, Nanoscale 7 (2015) 14872–14880.
mesoporous-TiO2 nanoparticles on mild steel, Prog. Org. Coat. 89 (2015) [53] T. Chen, H. Jung, J. Kim, Acetone vapor sensor made with
114–122. cellulose-TiO2 /MWCNTs hybrid nanocomposite, Proc. SPIE 8344 (2012)
[27] L. Ren, Y.Z. Li, J.T. Hou, J.L. Bai, M.Y. Mao, X.J. Zhao, N. Li, The pivotal effect of 83440B1-5.
the interaction between reactant and anatase TiO2 nanosheets with exposed
{0 0 1} facets on photocatalysis for the photocatalytic purification of VOCs,
Appl. Catal. B: Environ. 181 (2016) 625–634.
Biographies
[28] W. Guo, Q.Q. Feng, Y.F. Tao, L.J. Zheng, Z.Y. Han, J.M. Ma, Systematic
investigation on the gas-sensing performance of TiO2 nanoplate sensors for
enhanced detection on toxic gases, Mater. Res. Bull. 73 (2016)
Nan Chen received her B.S. degree in Department of Materials Science and Engi-
302–307.
neering from Yunnan University in 2014. She is currently a graduate student in
[29] M. Koudelka, A. Monnier, J. Sanchez, J. Augustynski, Correlation between the
Yunnan University and devotes to nanostructured functional materials and their
surface composition of Pt/TiO2 catalysts and their adsorption behavior in
applications in supercapacitors and gas sensors.
aqueous solutions, J. Mol. Catal. 25 (1984) 295–305.
[30] M.L. Zhang, Z.H. Yuan, T. Ning, J.P. Song, C. Zheng, Growth mechanism of Pt Yuxiu Li received her B.S. degree in Department of Materials Science and Engineer-
modified TiO2 thick film, Sens. Actuators B: Chem. 176 (2013) 723–728. ing from Yunnan University in 2014. She is currently a graduate student in Yunnan
[31] H. Aono, E. Traversa, M. Sakamoto, Y. Sadaoka, Crystallographic University and devotes to inorganic functional materials and their applications in
characterization and NO2 gas sensing property of LnFeO3 prepared by gas sensors.
thermal decomposition of Ln-Fe hexacyanocomplexes, Ln [Fe(CN)6 ]·nH2 O
Ln = La, Nd, Sm, Gd, and Dy, Sens. Actuators B: Chem. 94 (2003) 132–139. Dongyang Deng received her B.S. degree in Department of Materials Science and
[32] G.Y. Zhang, C.S. Li, F.Y. Cheng, J. Chen, ZnFe2 O4 tubes: synthesis and Engineering in 2014. She is currently a graduate student in Yunnan University and
application to gas sensors with high sensitivity and low-energy consumption, devotes to gas sensors and CO oxidation.
Sens. Actuators B: Chem 120 (2007) 403–410.
[33] B. Mondal, B. Basumatari, J. Das, C. Roychaudhury, H. Saha, N. Mukherjee, Xu Liu received his B.S. degree in Department of Materials Science and Engineer-
ZnO-SnO2 based composite type gas sensor for selective hydrogen sensing, ing from Yunnan University in 2015. He is currently a graduate student in Yunnan
Sens. Actuators B: Chem 194 (2014) 389–396. University and devotes to nanostructured metal oxides and their applications in gas
[34] P. Sun, X. Zhou, C. Wang, K. Shimanoe, G.Y. Lu, N. Yamazoeb, Hollow sensors and supercapacitors.
SnO2 /␣-Fe2 O3 spheres with a double-shell structure for gas sensors, J. Mater.
Xinxin Xing received her B.S. degree in Department of Materials Science and Engi-
Chem. A 2 (2014) 1302–1308.
neering from Yunnan University in 2013. She is currently a graduate student in
[35] Y.Q. Liang, Z.D. Cui, S.L. Zhu, Z.Y. Li, X.J. Yang, Y.J. Chen, J.M. Ma, Design of a
Yunnan University and devotes to porous materials and their applications in gas
highly sensitive ethanol sensor using a nano-coaxial p-Co3 O4 /n-TiO2
sensors.
heterojunction synthesized at low temperature, Nanoscale 5 (2013)
10916–10926. Xuechun Xiao received her B.E. degree in Department of High Polymer Material
[36] Y. Wang, Y.M. Wang, J.L. Cao, F.H. Kong, H.J. Xia, J. Zhang, B.L. Zhu, S.R. Wang, from China Textile University in 1997 and M.E. degree in Department of Materi-
S.H. Wu, Low-temperature H2 S sensors based on Ag-doped ␣-Fe2 O3 als Science and Engineering from DongHua University in 2004. Currently, she is a
nanoparticles, Sens. Actuators B: Chem 131 (2008) 183–189. teacher at the School of Materials Science and Engineering, Yunnan University. Her
500 N. Chen et al. / Sensors and Actuators B 238 (2017) 491–500

work is devoted to nanostructured functional materials and their applications in gas Institute of Metal and an Alexander von Humboldt fellow in Max-Planck-Institute
sensors. of Colloids and Interfaces, Germany, respectively. Currently, he is a professor at
the School of Physics and Astronomy, Yunnan University. His work is devoted to
Yude Wang obtained his M.S. degree in Physics Condensed State from Yunnan chemical and biochemical sensors, nanostructured functional materials and their
University in 1997 and Ph.D. in Materials Physics and Chemistry from Tsinghua applications.
University in 2003. From 2005–2008, he was a guest scientist in Max-Planck-

You might also like