You are on page 1of 11

Sensors and Actuators B 230 (2016) 581–591

Contents lists available at ScienceDirect

Sensors and Actuators B: Chemical


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

Experimental and theoretical studies of gold nanoparticle decorated


zinc oxide nanoflakes with exposed {1 0 1̄ 0} facets for butylamine
sensing
Yusuf Valentino Kaneti a,b , Xiao Zhang a , Minsu Liu b , David Yu b , Yuan Yuan c ,
Leigh Aldous c , Xuchuan Jiang b,∗
a
School of Materials Science and Engineering, University of New South Wales, Sydney, NSW 2052, Australia
b
Department of Chemical Engineering, Monash University, Melbourne, VIC 3800, Australia
c
School of Chemistry, University of New South Wales, Sydney, NSW 2052, Australia

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

Article history: The exposed surface facets play an important role in determining the gas-sensing performance of nanos-
Received 14 September 2015 tructured materials. This study reports the facile hydrothermal synthesis of zinc oxide nanoflakes with
Received in revised form 17 February 2016 exposed {1 0 1̄ 0} facets, as confirmed by the high resolution transmission electron microscopy (HRTEM)
Accepted 19 February 2016
and the corresponding selected area electron diffraction (SAED) analysis. The gas-sensing properties of
Available online 22 February 2016
the ZnO nanoflake sensor were investigated toward toxic n-butylamine, an important marker compound
in food and medical industries. The pure ZnO nanoflake sensor exhibits a response of 23.9–50 ppm of n-
Keywords:
butylamine at an optimum operating temperature of 300 ◦ C. Density Functional Theory (DFT) simulations
Organic amine
Gas-sensing
were used to study the adsorption behavior of n-butylamine on the ZnO(1 0 1̄ 0) surface. The results show
Density functional theory simulation that n-butylamine chemically adsorb on the ZnO(1 0 1̄ 0) surface through the formation of a bond between
Gold nanoparticles the nitrogen atom of the n-butylamine (C4 H11 N) and the surface Zn atom of ZnO. To further improve
Zinc oxide nanostructures the gas-sensing properties, the as-prepared ZnO nanoflakes were subsequently loaded with three dif-
ferent quantities of Au (1.37, 2.82, and 5.41 wt% Au). The gas-sensing measurements indicate that the
Au nanoparticle-decorated ZnO nanoflakes display superior sensing performance to non-modified ZnO
nanoflakes by exhibiting 4–6 times higher response and an improved selectivity toward n-butylamine
gas, along a decreased optimum operating temperature of 240 ◦ C. Moreover, the response and recovery
properties of the ZnO nanoflake sensor are improved by a factor of 1.5–2.5 depending on the Au loading.
The enhanced sensing performance of the Au nanoparticle-decorated ZnO nanoflakes to n-butylamine
gas can be attributed to the excellent catalytic activity of Au nanoparticles (NPs) which promotes a
greater adsorption of oxygen molecules on the surface of ZnO and the presence of multiple electron
depletion layers, specifically at the surface of ZnO and at the ZnO/Au interface, which greatly increases
their conductivity upon exposure to the gas.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction ing of harmful gases, because of the low manufacturing cost, easy
processing, and reliable performance. Semiconducting gas sensors
Detection and monitoring of hazardous gases in food, chemical usually consist of metal oxide materials, such as zinc oxide (ZnO),
and manufacturing industries have received increasing atten- tin dioxide (SnO2 ), iron oxide (Fe2 O3 ), tungsten oxide (WO3 ), and
tion recently due to the growing awareness about public health indium oxide (In2 O3 ) [2].
safety [1]. Many studies have been carried out to develop high- Among these oxides, ZnO, is an important n-type semiconduc-
performance gas sensors with excellent response, selectivity and tor with a band gap of Eg = 3.3 eV, and has been widely applied for
stability. Among the various forms of gas sensors, semiconducting different applications including gas sensors [3,4], photocatalysts
(chemiresistive) gas sensors are particularly attractive for the sens- [5,6] solar cells [7,8], and lithium-ion batteries [9]. ZnO exhibits
excellent thermal and chemical stability and high electrical conduc-
tivity, which makes it highly attractive for gas-sensing application
∗ Corresponding author. [10]. However, pure ZnO sensors typically suffer from low selectiv-
E-mail address: xuchuan.jiang@monash.edu (X. Jiang). ity and high optimum operating temperatures of ≥300 ◦ C, which

http://dx.doi.org/10.1016/j.snb.2016.02.091
0925-4005/© 2016 Elsevier B.V. All rights reserved.
582 Y.V. Kaneti et al. / Sensors and Actuators B 230 (2016) 581–591

may impede their broader applications [11,12]. To reduce such Table 1


Summary of the synthesis conditions used for the preparation of Au-decorated ZnO
problems, many efforts have been devoted in this area aiming
samples.
to controlling particle size and morphology [13], maximizing the
exposure of active crystal facets [14,15], and introducing transition Samples V (HAuCl4 ·3H2 O) (␮L) V (NaBH4 ) (␮L) Au loading (wt %—ICP)
metal dopants [11,15–19], metal oxide additives [1,17] or noble ZnO NF/Au (1) 203 121 1.37
metal sensitizers [11,17,20–23]. ZnO NF/Au (2) 508 303 2.82
Organic amines (e.g., n-butylamine) are important marker com- ZnO NF/Au (3) 1016 606 5.41

pounds for quality control in food industries and medical diagnosis


[24]. n-Butylamine is also frequently used as a vulcanizing accel-
butylamine gas (along with other gases e.g. ethanol, n-butanol, and
erator and reaction initiator in the polymer industries, a chemical
acetone) and ZnO(1 0 1̄ 0) surface of the nanoflakes during the gas-
intermediate in the production of emulsifying agents, rubber chem-
sensing process.
icals, tanning agents and special soaps. Furthermore, it is also
utilized in the manufacture of pharmaceuticals, dyes, insecticides
and textiles. n-Butylamine is toxic and is easily absorbed through 2. Experimental
skin. Direct exposure of n-butylamine vapor can cause eye, skin
and upper respiratory tract irritation [25]. Therefore, it is nec- 2.1. Chemicals
essary to develop sensor material(s) to detect n-butylamine gas
with excellent response, selectivity, and stability. To date, atmo- Zinc chloride (ZnCl2 , 99%), sodium hydroxide (NaOH, 99%),
spheric n-butylamine has been quantified by isotachophoresis, and gold(III) chloride trihydrate (HAuCl4 .3H2 O, 99.99%), sodium boro-
by high pressure liquid chromatography (HPLC), however, these hydride (NaBH4 ), n-butylamine (C4 H11 N, 99%) acetone (C3 H6 O,
techniques require expensive equipments [26]. To date, there were 99.9%), n-butanol (C4 H10 O, 99.8%), ethanol (C2 H6 O, 99.8%),
few limited reports on the use of solid state semiconductor nano- methanol (CH4 O, 99.9%), and heptane (C7 H14 , 99%) were purchased
sensors for detection of n-butylamine, with only sensors based on from Sigma–Aldrich and used as received without further purifica-
WO3 , V2 O5 and AgVx Oy nanostructures have been reported so far tion. All the chemicals were of analytical grade. Ultra-pure water
[27–29]. The reported responses of these sensors to n-butylamine, was used throughout the experiments.
however, are relatively low and require further improvement.
A clear understanding of the sensing mechanism of ZnO nano- 2.2. Synthesis of porous ZnO nanoflakes
sensors is crucial for ensuring the successful application of these
sensors in real life monitoring applications. Theoretical simulation The porous ZnO nanoflakes were prepared according to our pre-
studies have been carried out to investigate the adsorption of oxi- vious report with minor modifications [2]. In a typical procedure,
dizing (e.g., NOx , SOx ) and reducing gases (CO, NH3 , ethanol) on a mixture solution was first prepared by mixing a 40 mL solu-
different ZnO surfaces [30–33]. The DFT study by Breedon et al. tion of 0.15 M sodium hydroxide (NaOH) and a 2 mL solution of
[30,34] found that NO2 and NO interacted weakly with ZnO(1 0 1̄ 0) 0.3 M ZnCl2 . Next, the white-colored suspension was subsequently
and (2 1̄ 1̄ 0) crystal without generating any significant surface dis- transferred into a 50 mL Teflon-lined stainless-steel autoclave and
tortions. Prades et al. [32] expanded the study further by examining heated at 150 ◦ C for 16 h and allowed to cool to room temperature
the adsorption of NO2 on ZnO(1 0 1̄ 0) and (1 1 2̄ 0) crystal planes naturally. The resulting precipitate was collected by centrifugation,
with 12.5% O vacancy and found that NO2 was strongly adsorbed washed thoroughly with deionized water and ethanol several times
on the surface Zn atoms, in the presence of oxygen vacancies. More- and finally dried at 60 ◦ C for 5 h before further use.
over, the DFT investigation by An et al. on the adsorption of several
reducing and oxidizing gases on defect-free, single-walled ZnO 2.3. Synthesis of Au nanoparticle-decorated ZnO nanoflakes
nanotube reveal that O2 and H2 molecules were physisorbed on the
sidewall of the ZnO nanotube while CO, NH3 , and NO2 were molec- Briefly, three steps are involved in the synthesis process. Firstly,
ularly chemisorbed. Despite some successes, theoretical simulation 30 mg of ZnO nanoflakes was well-sonicated in 10 mL of deionized
studies of the adsorption of organic amine (e.g., n-butylamine) on water for 20 min. Secondly, varying quantities of a 0.01 M solu-
the ZnO surface were not yet reported. Hence, there exists a lack tion of HAuCl4 ·3H2 O were added into the ZnO suspension under
of clear understanding of the atomic scale interactions that occur magnetic stirring to achieve ZnO/Au nanocomposites with three
between organic amine gas molecules and ZnO surface during the different Au loading (labeled as ZnO NF/Au (1), ZnO NF/Au (2), and
gas-sensing process, particularly in terms of: (i) the effect of the ZnO NF/Au (3)). Finally, different amounts (depending on the Au
interaction on the bond length and bond angle of the amine gas loading) of freshly made 0.05 M NaBH4 solution (Au3+ :BH4 − = 1:3)
molecules, and (ii) the adsorption mechanism of the amine gas were added into the ZnO suspension and allowed to react for 15 min
molecules on the ZnO surface. under stirring. The purplish products were thoroughly washed with
This study reports the facile hydrothermal synthesis of ZnO water and ethanol several times until the supernatant was clear
nanoflakes with {1 0 1̄ 0} facets under mild hydrothermal condi- and then dried in an oven at 60 ◦ C for 5 h. The details regarding the
tions and the subsequent modification with different Au contents quantities of HAuCl4 and NaBH4 solutions used during the synthe-
(1.37, 2.82, and 5.41 wt% Au) for gas-sensing application. The sis process can be found in Table 1. The final Au contents in the
morphology and composition of the as-prepared ZnO/Au nanocom- samples of ZnO NF/Au (1), ZnO NF/Au (2), and ZnO NF/Au (3) were
posites were characterized by various analytical techniques, determined by ICP-AES to be 1.37, 2.82, and 5.41 wt%, respectively.
including transmission electron microscopy (TEM), high resolu-
tion TEM (HRTEM), X-ray diffraction (XRD), X-ray photoelectron 2.4. Characterization
spectroscopy (XPS), and inductively coupled plasma atomic emis-
sion spectroscopy (ICP-AES). The effect of the different Au loading The phase composition and purity of the synthesized ZnO
on the sensing properties of ZnO nanoflakes toward n-butylamine nanoflakes were examined using Phillips X’pert Multipurpose
and other gases and the important parameters such as response, X-ray Diffraction System (MPD) equipped with graphite mono-
selectivity and stability were evaluated. Additionally, density func- chromatized Cu-K␣ radiation ( = 1.54 Å) in the 2 range of 20–70◦ .
tional theory (DFT) simulations were also conducted for achieving The particle morphologies were observed on a FEI Nova NanoSEM
a clear atomic-scale understanding of the interaction between n- 230 field emission scanning electron microscope (SEM). TEM
Y.V. Kaneti et al. / Sensors and Actuators B 230 (2016) 581–591 583

images were recorded with a Tecnai G2 20 transmission elec-


tron microscope, operated at an accelerating voltage of 200 kV.
HRTEM images were recorded on a Phillips CM200 field emission
gun transmission electron microscope with an accelerating voltage
of 200 kV. X-ray photoelectron spectroscopy (XPS) surface analy-
sis of the samples was carried out using an ESCALAB250Xi X-ray
photoelectron spectrometer, which utilizes Al-K␣ radiation as the
excitation source. ICP-AES measurements were carried out to quan-
tify the amount of Au in the ZnO/Au samples using Optima7300DV-
ICP-OES PerkinElmer, with detection limits of ∼0.05 mg/L.

2.5. Gas sensor fabrication and measurements

The gas sensors were fabricated according to the procedures


reported in our previous studies. Basically, the Au-decorated
ZnO samples were first mixed and ground with the binder Fig. 1. XRD patterns of (a) pure ZnO nanoflakes, and ZnO nanoflakes loaded with
polyvinylidene fluoride (PVDF) in agate mortar. Then, 1-methyl- (b) 1.37 wt% Au (labeled as ZnO NF/Au (1)), (c) 2.82 wt% Au (labeled as ZnO NF/Au
2-pyrrolidone was added into the mixture to form white slurry, (2)), and (d) 5.41 wt% Au (labeled as ZnO NF/Au (3)).
which was then coated on a ceramic tube with previously printed
Au electrodes and Pt conducting wires. The ceramic tube was where E(A/surface) is the total energy of the gas molecules adsorp-
subsequently sintered at 350 ◦ C for 3 h to remove the polymeric
tion system in the equilibrium state on ZnO(1 0 1̄ 0) surface; EA and
binder and to improve the stability of the sensor. The gas-sensing
Esurface are the total energies of the free gas molecules and clean
properties of the Au nanoparticle-decorated ZnO nanoflakes were
ZnO(1 0 1̄ 0) surface, respectively. With this definition, a negative
evaluated using a computer-controlled WS-30A gas-sensing mea-
value corresponds to the stable adsorption on ZnO(1 0 1̄ 0) surface.
surement system (Weisheng Electronics Co., Ltd., Henan, China), as
schematically shown in Fig. S1.
3. Results and discussion
Prior to the measurement, a Ni–Cr resistor was inserted into
the ceramic tube as a heater, which enables control of the oper-
3.1. Morphology and composition
ating temperature by adjusting the heating voltage (Vheating ). A
reference resistor was placed in series with the sensor to form a
The overall quantity of Au in the three Au-decorated ZnO
complete measurement circuit. The probe gas such as n-butylamine
nanoflake samples (labeled as ZnO NF/Au (1), ZnO NF/Au (2) and
was injected into the testing chamber using a micro-syringe. The
ZnO NF/Au (3)) were initially determined using ICP-AES. As shown
response (S) is defined as the ratio of the resistances measured in
in Table 1, the Au contents in samples ZnO NF/Au (1), (2), and (3)
air (Ra ) and in the tested gas atmosphere (Rg ): S = Ra /Rg . The output
are 1.37, 2.82 and 5.41 wt%, respectively. The phase composition of
voltage was set at 5 V and the gas-sensing measurements were car-
the ZnO nanoflakes with or without metallic decoration was char-
ried out at a relative humidity of 55–65%, with a reference resistor
acterized using XRD. Fig. 1 displays the XRD patterns of the pure
of 100 k.
and Au-decorated ZnO products. All of the diffraction peaks can be
well indexed to the diffraction peaks of hexagonal wurtzite ZnO
2.6. Density functional theory (DFT) simulations (JCPDS No. 36-1451). In comparison, all the XRD patterns of the
Au-decorated ZnO products show the presence of diffraction peaks
DFT simulations were conducted using the commercial soft- of ZnO along with additional peaks at 2 = 38.4, 44.0, and 66.4◦ ,
ware: Materials Studio (Version 4.3, Accelrys Inc., 2007) with the which corresponds well to the (1 1 1), (2 0 0), and (2 2 0) peaks of
Dmol3 module. GGA (generalized gradient approximation) was cubic gold (JCPDS No. 04-0784). The elemental EDS map of the Au
employed for the exchange-correlation. In this work, PBE (Perdew, nanoparticle-decorated ZnO nanoflakes shown in Fig. S2 clearly
Burke, and Emzerhof) exchange-correlation functional coupled highlights the presence of Zn, O, and Au elements. Furthermore,
with the precise numerical basis set DNP (double numerical plus it also shows the good dispersion of the Au NPs on the surface of
polarization) and a basis file of 4.4 was used. The core electrons the ZnO nanoflakes.
were treated with DFT semi-core pseudo potentials (DSPPs) with a Fig. 2a displays the SEM image of the pure ZnO nanoflakes, show-
global orbital cut off of 4.4 Å. The surface simulations were per- ing diameters of 400–800 nm. The TEM image of an individual ZnO
formed using a slab model under periodic boundary conditions nanoflake (Fig. 2b) clearly indicates the presence of pores with sizes
(Scheme 1). A supercell with a surface thickness of ∼9 Å and a vac- ranging from 2–20 nm. Such porous structure may be beneficial for
uum space of 20 Å was created from the ZnO lattice structure in increasing the accessible surface area and active sites, available for
which the bottom layers (∼5 Å) were fixed and the top layers were diffusing gas molecules during the sensing process [13]. BET surface
relaxed to simulate the surface properties. Then, the optimized gas area measurements of the pure and Au-modified ZnO nanoflakes
molecules were positioned ∼2.0 Å above the adsorption surface. are shown in Fig. S3. The specific surface areas (S.S.A) of the pure
The gas molecules investigated include n-butylamine, methanol, ZnO NF, ZnO NF/Au (1), ZnO NF/Au (2), and ZnO NF/Au (3) samples
ethanol, n-butanol, acetone and heptane. were measured to be 8.56, 9.39, 9.98, and 7.90 m2 /g, respectively.
The adsorption energy can be used to describe the variation Furthermore, the pore sizes of the four samples were found to vary
of total energy of each species before and after adsorption. The between 2–20 nm. Additionally, no major change was observed in
signs and values of adsorption energy can indicate the possibility of the average pore size of the ZnO nanoflakes before and after surface
adsorption. In the present work, the adsorption energy (Eads ) was decoration by Au nanoparticles.
calculated according to the equation [15]: The HRTEM image of a single ZnO nanoflake shows the presence
of distinct lattice fringes with an interplanar spacing of 0.261 nm,
E ads = E (A/surface) −(E A + E surface ) corresponding to the d-spacing of ZnO(0 0 0 1) crystal plane (Fig. 2c)
[2,4]. The selected area electron diffraction (SAED) pattern of the
584 Y.V. Kaneti et al. / Sensors and Actuators B 230 (2016) 581–591

Scheme 1. (a) Side view (top) and (b) top view (bottom) of the ZnO (1 0 1̄ 0) crystal surface (red and blue atoms represent O and Zn, respectively); (c) Schematic illustration
of the crystal faces in ZnO nanoflakes; (d) Free n-butylamine molecule (blue, grey, and white atoms represent N, C and H, respectively). (For interpretation of the references
to color in this figure legend, the reader is referred to the web version of this article.)

Fig. 2. (a) SEM, (b) TEM and (c) HRTEM images of pure ZnO nanoflakes; (c) SEM, (d) TEM and (e) HRTEM images of Au nanoparticle-decorated ZnO nanoflakes.

ZnO nanoflakes is consistent with the diffraction pattern of the become rough following the Au modification, indicating the suc-
[1 0 1̄ 0] zone axis of ZnO (inset of Fig. 2c). This provides evidence cessful deposition of Au nanoparticles.
that the bottom and the top surfaces of the ZnO nanoflakes expose The HRTEM image of a single Au-decorated ZnO nanoflake
the ±{1 0 1 0} planes [15,35]. Fig. 2d shows a typical SEM image (Fig. 2f) taken from the selected area in Fig. 2e shows the crystal
of the Au nanoparticle-decorated ZnO nanoflakes (sample ZnO lattices of ZnO(0 0 0 1) along with additional lattice fringes with a
NF/Au (2)). It can be observed that the surface of the nanoflakes measured d-spacing of ∼0.238 nm, indexed to the Au(1 1 1) plane
Y.V. Kaneti et al. / Sensors and Actuators B 230 (2016) 581–591 585

Fig. 3. TEM images of ZnO nanoflakes loaded with (a) 0 wt% Au (ZnO NF), (b) 1.37 wt% Au (ZnO NF/Au (1)), (c) 2.82 wt% Au (ZnO NF/Au (2)), and (d) 5.41 wt% Au (ZnO NF/Au
(3)).

[36]. These results confirm the successful deposition of Au NPs on peaks at the binding energies of 83.5 eV and 85.4 eV, indexed to Au
the surface of ZnO nanoflakes. Fig. 3 shows the TEM images of ZnO 4f7/2 and Au 4f5/2 peaks, respectively. The binding energies of the
nanoflakes loaded with three different amounts of Au. It is obvi- Au4f peaks in the Au nanoparticle-decorated ZnO nanoflakes are
ous that with increasing Au loading, a greater amount of Au NPs is much lower than those found in bulk metallic Au (84.0 eV for Au
deposited on the surface of the ZnO nanoflakes. The average size of 4f7/2 while 87.67 eV for Au 4f5/2 ) [37,38]. The decrease in binding
the Au NPs in samples ZnO NF/Au (1), ZnO NF/Au (2), and ZnO NF/Au energies is caused by the larger electronegativity of Au compared
(3) was measured to be 7.18, 8.08, and 8.68 nm, respectively, from to Zn atoms in ZnO, which promotes a charge transfer from ZnO
a measurement of 120 Au NPs (Fig. S4). This suggests that there is a to Au. In addition to the two Au 4f peaks, a weak shoulder peak
slight increase in the size of the deposited Au NPs with increasing is also noted at 88.68 eV, corresponding to the Zn 3p3/2 peak from
Au loading. ZnO; whereas the other shoulder peak at 91.68 eV can be indexed
XPS technique was employed to identify the surface compo- to the Zn 3p1/2 peak. The surface compositional analysis of the three
sition of Au nanoparticle-decorated ZnO nanoflakes using sample ZnO/Au samples is summarized in Table S1. By comparing the ICP
ZnO NF/Au (2) as an example. As shown in Fig. S5a, the Zn 2p XPS and XPS results in Tables 1 and S1, respectively, it can be deduced
peak of the pure ZnO nanoflakes at 1021.22 eV can be assigned to that the majority of the Au NPs were decorated on the surface of
Zn 2p3/2 , while in the ZnO/Au nanoflakes, the position of this Zn the ZnO nanoflakes.
2p3/2 peak shifted by 0.13 eV to 1021.35, suggesting that a charge
transfer occurs between Au and ZnO. In addition, the O 1s spec-
trum shown in Fig. 4c is asymmetric, suggesting the presence of 3.2. Gas-sensing performance
several oxygen species. Deconvolution of the O 1s peak shows the
existence of two peaks centered at 530.2 eV (OI ) and 531.9 eV (OII ), Fig. 5a displays the responses of sensors made from
attributed to the O2− ions in the wurtzite ZnO and O− and O2− ions pure ZnO nanoflakes (ZnO NF) and Au nanoparticle-decorated
in the oxygen deficient regions mainly caused by oxygen vacancies. ZnO nanoflakes (ZnO NF/Au(1), (2) and (3)) toward 50 ppm
The high binding energy oxygen species, OIII centered at 532.8 eV of n-butylamine with increasing operating temperature from
belong to the chemisorbed oxygen species caused by the surface 150–330 ◦ C. The pure ZnO NF sensor exhibits an optimum operating
hydroxyl groups (O H bond) [37]. temperature of 300 with S = 23.9 toward 50 ppm of n-butylamine.
Fig. 4d shows the Au 4f XPS spectrum of the Au nanoparticle- In comparison, the three ZnO NF/Au sensors exhibit an identical
decorated ZnO nanoflakes. This figure reveals the presence of two optimum operating temperature of 240 ◦ C, which is 60 ◦ C lower
than that of the pure ZnO NF sensor. Interestingly, even at a low
586 Y.V. Kaneti et al. / Sensors and Actuators B 230 (2016) 581–591

Fig. 4. (a) XPS survey spectra and XPS high resolution spectra of (b) Zn2p, (c) O1s and (d) Au4f of the Au nanoparticle-decorated ZnO nanoflakes (sample ZnO NF/Au (2)—with
2.82 wt% Au).

Fig. 5. (a) Response vs. operating temperature curves of pure ZnO nanoflakes (NF) and ZnO NF/Au sensors toward 50 ppm of n-butylamine, (b) dynamic response-recovery
transients of ZnO NF/Au sensors toward different concentration of n-butylamine at 240 ◦ C, (c) response vs. concentration curves of pure ZnO NF and ZnO NF/Au sensors at
240 ◦ C, and (d) selectivity test results of pure ZnO NF and ZnO NF/Au sensors at 240 ◦ C.
Y.V. Kaneti et al. / Sensors and Actuators B 230 (2016) 581–591 587

Table 2
Comparison of the n-butylamine sensing performance of the as-prepared Au nanoparticle-decorated ZnO nanoflakes against previously reported metal oxide nanosensors.

Morphologies Response (S = Ra /Rg ) Concentration (ppm) Operating temp. (◦ C) Refs.


a
ZnO/Au nanoflakes 73.2 50 240 This work
Ag2 V4 O11 nanobelts 3.5 50 260 [27]
VOx@15% Ag nanobelts 2.3 50 260 [27]
Ag0.35 V2 O5 nanobelts 2.1 50 260 [27]
V2 O5 nanofibers 0.42 10 22 [28]
a
Contains 2.82 wt% Au.

operating temperature of 140 ◦ C, the three ZnO NF/Au sensors Table 3


Summary of calculated bond distances and adsorption energy for n-butylamine (the
still exhibit a reasonably high response of 8.2–16 to 50 ppm of n-
numbering of atoms of the n-butylamine molecule is given in Scheme 1d).
butylamine, indicating their potential for low-temperature amine
sensing. Bond Gas phase ZnO
1 0 1̄ 0
The dynamic response-recovery transients of the pure ZnO NF
and ZnO NF/Au sensors are shown in Figs. S5 and 5 b, respectively. It Bond distance/angle N H1 (Å) 1.023 1.046
is clear that the output voltage of all the sensors increases when the N H2 (Å) 1.023 1.023
C H (Å) 1.102 1.102
n-butylamine gas is inserted, and subsequently decreases as the gas
C1 N (Å) 1.472 1.493
is removed. Fig. 5c displays the responses of pure ZnO NF and ZnO C1 C2 (Å) 1.526 1.527
NF/Au sensors toward different concentrations of n-butylamine at
H1 N C1 (◦ ) 106.0 105.9
240 ◦ C. It can be observed that the ZnO NF/Au sensors show 4–6
N C1 C2 (◦ ) 110.8 113.8
times higher response toward n-butylamine than pure ZnO NF sen-
sor, with a trend of ZnO NF/Au (2) [2.82 wt% Au] > ZnO NF/Au (1) NBA ZnZnO (Å) – 2.118
H1 OZnO (Å) – 1.849
[1.37 wt% Au] > ZnO NF/Au (3) [5.41 wt% Au] > ZnO NF [0 wt% Au].
In fact, the Au-decorated ZnO sensor, ZnO NF/Au (2), exhibits a Eadsorption (eV) – −7.30
response as high as 119.3 to 100 ppm of n-butylamine vapor. Fur-
thermore, even at a low concentration of 10 ppm, this particular
ZnO/Au sensor still exhibits a reasonably high response of 18.5, of the sensor. As the Au-decorated ZnO sensor, ZnO NF/Au (2),
indicating a possible detection limit in ppb level. From the above exhibit the highest response and best selectivity compared to the
analysis, it is evident that the Au modification greatly enhances the other ZnO NF/Au sensors, its reproducibility and stability were
sensing response of the ZnO nanoflakes toward n-butylamine; and evaluated toward n-butylamine vapor. As shown in Fig. 6c, the
that different Au contents can lead to different responses. response/recovery behavior of the ZnO NF/Au (2) sensor toward
Selectivity is another critical criterion in gas-sensing which 40 ppm of n-butylamine is highly reproducible following 8 straight
needs to be evaluated to ensure an accurate detection of a single cycles of testing at 240 ◦ C. Furthermore, the good stability of the
gas in a mixer with a higher response. Fig. 5d shows the selectivity ZnO NF/Au (2) sensor is clearly highlighted in Fig. 6d as the response
tests of the pure ZnO NF and three ZnO NF/Au sensors to different values remain very consistent during a 21-day test period. This
gases, including n-butanol, ethanol, acetone, methanol and hep- indicates its excellent potential as a sensitive, selective and sta-
tane at 240 ◦ C. It can be noted that under similar conditions, ZnO ble sensor material for the detection of toxic n-butylamine vapor.
NF/Au sensors show excellent selectivity to n-butylamine, with the Table 2 compares the n-butylamine sensing response of the fab-
response being 11–70 time higher than those to other gases. In com- ricated ZnO nanoflake/Au sensor (sample ZnO NF/Au (2)) against
parison, the response of the pure ZnO NF sensor to n-butylamine previously reported sensors based on metal oxide nanostructures.
is only 2–7 time higher than the response to other gases. This indi- It can be seen from Table 2 that the ZnO NF/Au (2) sensor (contains
cates that the Au decoration on ZnO nanoflakes can highly improve 2.82 wt% Au) show superior sensing performance to n-butylamine
the selectivity for accurate detection of n-butylamine. compared to Ag2 V4 O11 [27], Ag0.35 V2 O5 [27], VOx /Ag nanobelts
The response and recovery time values of pure ZnO NF and ZnO [27], and V2 O5 nanofibers [28].
NF/Au sensors are given in Fig. 6a and b, respectively. Here, the
response time ( res ) is defined as the time required for the conduc- 3.3. DFT simulation
tance variation of the sensor to reach 90% of its equilibrium value,
following exposure to n-butylamine vapor. On the contrary, the The HRTEM and SAED analysis in Fig. 2c reveal that the
recovery time ( rec ) is defined as the time needed by the sensor to top/bottom surface of the ZnO nanoflakes expose the ±(1 0 1̄ 0)
return to 10% of its original conductance in air, after removal of the plane. DFT simulations were used to simulate the adsorption
n-butylamine vapor. It can be observed from Fig. 6a that the  res behavior of n-butylamine along with other gas molecules such as
values of the pure ZnO NF sensor upon exposure to 10–100 ppm methanol, ethanol, acetone, and n-butanol, and heptane on the
of n-butylamine gas vary from 55–77 s, whereas the  res values ZnO(1 0 1̄ 0) surface. Scheme 1a and b shows the side and top view
of ZnO NF/Au (1), ZnO NF/Au (2), and ZnO NF/Au (3) sensors are of the ZnO(1 0 1̄ 0) surface, respectively. This surface was firstly
in the ranges of 23–57, 19–57, and 24–37 s, respectively. These optimized, and then the gas molecules were placed on the opti-
results indicate the improvement in the response speed of the ZnO mized ZnO(1 0 1̄ 0) surface during the simulations.
nanoflake sensor by a factor of 1.5–2.5 following the Au modifica- Table 3 shows the bond distances between various atoms,
tions. In comparison, the  rec values of the ZnO NF/Au (2) sensor the bond angles of the n-butylamine molecule before and after
upon exposure to 10–100 ppm range from 11–100 s, whereas the adsorption and the adsorption energies on ZnO(1 0 1̄ 0). The DFT
 rec values of ZnO NF/Au (1), ZnO NF/Au (2), and ZnO NF/Au (3) calculation performed for the free n-butylamine molecule shows
sensors are in the range of 11–46, 8–76, and 8–47 s, respectively that it exhibits bond distances of 1.023, 1.102, 1.526 and 1.472 Å,
(Fig. 6b). These results suggest that the decoration of Au NPs also corresponding to N H, C H, C C and C N bonds, respectively. The
significantly improves the recovery time of the ZnO NF sensor. DFT calculations indicate that the distance between H (C4 H11 N)
Reproducibility and stability are two important parameters and O on the ZnO(1 0 1̄ 0) surface is 1.849 Å (Table 3). Furthermore,
of a sensor material, crucial for ensuring long-term reliability from this table, it can be noted that the N H bond distance of the
588 Y.V. Kaneti et al. / Sensors and Actuators B 230 (2016) 581–591

Fig. 6. (a) Response and (b) recovery time values of pure ZnO NF and ZnO NF/Au sensors for different concentration of n-butylamine at 240 ◦ C. (c) Reproducibility and (d)
stability tests of the ZnO NF/Au (2) sensor (contains 2.82 wt% Au) toward 40 ppm of n-butylamine at 240 ◦ C.

C4 H11 N molecule stretches the furthest (by 0.23 Å) following its cases of n-butylamine, ethanol, acetone, and n-butanol. Hence, the
interaction with the surface Zn atom. Hence, as the n-butylamine responses of the pure ZnO sensor toward these two gases are
molecule approaches ZnO(1 0 1̄ 0) surface during the sensing pro- expected to be the lower than the other four gases. This is sup-
cess, the nitrogen atom of n-butylamine (C4 H11 N) forms a chemical ported by the gas-sensing results shown in Fig. 5d. At present, an
bond with the surface Zn atom. accurate DFT simulation study on the adsorption of these gases on
Alcohol molecules such as ethanol and n-butanol are also chemi- Au-modified ZnO(1 0 1̄ 0) surface is challenging as the Au NPs are
cally adsorbed on the ZnO(1 0 1 0) surface, however the mechanism randomly oriented on the ZnO surface (i.e., they do not follow a spe-
is slightly different to that of n-butylamine. The DFT simulation cific orientation). In this study, we have explained the role of Au in
results summarized in Table S2 reveal that the O H, C H, C O, and enhancing the gas-sensing performance of the ZnO nanoflakes in
C C bond distances in free ethanol molecule (prior to adsorption) terms of ‘electronic’ and ‘chemical’ sensitizations as described in
were determined to be 0.972, 1.101, 1.440, and 1.517 Å, respec- Section 3.4.
tively. In comparison, the O H, C H, C O, and C C bond distances
in free n-butanol molecule were calculated to be 0.972, 1.101, 1.439, 3.4. Sensing mechanism
and 1.520 Å, respectively, as shown in Table S3. These results are
in good agreement with previously reported theoretical [15,39,40] The sensing mechanism of an n-type semiconductor gas sensor
and experimental results [41]. As shown in Fig. 7b and c, the such as ZnO is largely based on the resistance change caused by
chemical adsorption of ethanol (C2 H6 O) or n-butanol (C4 H8 O) gas the adsorption of oxygen or gas molecules [3,15]. When the pure
molecule on the ZnO(1 0 1̄ 0) surface occurs through the formation ZnO sensor is exposed to air, oxygen molecules (O2 ) adsorb on the
of a (covalent) bond between the oxygen atom of ethanol/n-butanol surface of ZnO nanoflakes and become ionized to O− or O2− by cap-
molecule with the surface Zn atom of ZnO. turing free electron(s) from the conduction band of ZnO, according
Acetone also chemically adsorb on the ZnO(1 0 1̄ 0) surface by to the following equations [3,37]:
following a similar mechanism to ethanol and n-butanol, with the
oxygen atom of the acetone gas molecule being bonded to the sur- O2 (gas) ↔ O2 (adsorbed) (1)
face Zn atom of ZnO (Fig. 7d). However the adsorption energy of O2 (adsorbed) + e− ↔ O2 − (adsorbed) (2)
acetone (−4.9 eV) is less than those of ethanol (−7.15 eV) and n-
− − −
butanol (−7.24 eV), indicating a weaker adsorption, thus leading O2 (adsorbed) + e ↔ 2O (adsorbed) (3)
to a poorer response of the pure ZnO sensor to acetone than to − − 2−
O (adsorbed) + e ↔ O (adsorbed) (4)
ethanol or n-butanol. This is in accordance with the results depicted
in Fig. 5d which shows that the response of the ZnO nanoflakes is The above reactions lead to the formation of an electron
in the order of: acetone < ethanol < n-butanol. As for other gases depletion layer (L), which reduces the electron concentration.
such as heptane and methanol, they were found to be physisorbed However, as the ZnO nanoflake sensor is exposed to n-butylamine,
on the ZnO(1 0 1̄ 0) surface but not chemically adsorbed as in the the n-butylamine gas molecules react with the chemisorbed
Y.V. Kaneti et al. / Sensors and Actuators B 230 (2016) 581–591 589

Fig. 7. Slice view of the gas molecules (a) methanol, (b) ethanol, (c) n-butanol, (d) acetone, (e) heptane, (f) n-butylamine adsorbed on the ZnO (1 0 1̄ 0) crystal surface (red,
blue, grey and white atoms represent O, Zn, C and H, respectively). (For interpretation of the references to color in this figure legend, the reader is referred to the web version
of this article.)

oxygen species on the surface of the ZnO nanoflakes. This releases


the trapped electrons back into the conduction band of ZnO, which
increases the electron concentration, and ultimately decreases the
resistance of the ZnO sensor.
The sensing mechanism of the ZnO NF/Au sensor, however, is
quite different. This is because of the presence of Au NPs on the
surface of the ZnO nanoflakes, which can induce both chemical and
electronic sensitization. Chemical sensitization occurs due to the
excellent catalytic activity of Au NPs. Our previous simulation study
on Au-loaded SnO2 nanorods with {1 1 0} facets has shown that
the loading of Au NPs on the metal oxide surface can lower the
diffusivities of oxygen, attract the O O bond in oxygen molecules
and enable easier removal of H2 O (moisture) from the metal oxide
surface at lower temperatures (<250 ◦ C) [36]. These factors enable
O2 molecules to diffuse faster into surface vacancies and capture
electrons from the conduction band of ZnO nanoflakes to become
oxygen ions O− or O2− [1,42]. Consequently, this raises the quantity
of adsorbed oxygen and the oxygen molecule to ion conversion rate,
leading to a faster degree of electron depletion from the ZnO/Au
interface than that at the surface of pure ZnO nanoflakes. The n-
butylamine gas then reacts with the chemisorbed oxygen species
and releases the electrons back to the surface of ZnO/Au nanoflakes,
resulting in a greater change in electrical conductivity and therefore
an enhanced response toward n-butylamine.
Apart from chemical sensitization, the decoration of Au NPs
on the surface the ZnO nanoflakes may also induce electronic
sensitization, because of the differences in their work functions
((Au) = 5.1 eV, (ZnO) = 4.65 eV), as shown in Fig. 8 [1]. As Au has
a greater work function than ZnO (4.65 eV), a greater amount of
electrons will be transferred from the conduction band of ZnO to Au
than to the pure ZnO nanoflakes. This results in a greater band bend-
ing at the ZnO/Au interface with a broad depletion layer (Fig. 8b) Fig. 8. A schematic illustration of the energy band diagram of (a) pure ZnO and (b)
ZnO/Au (H and L in the diagram represent potential barrier and the depletion
[2,43]. The existence of this additional depletion layer at the ZnO/Au
layer, respectively).
590 Y.V. Kaneti et al. / Sensors and Actuators B 230 (2016) 581–591

interface in addition to that at the surface of the ZnO nanoflakes tralian Microscopy and Microanalysis Research Facilities (AMMRF).
lead to a greater change in resistance of ZnO/Au sensor following The authors would like to thank Dr. Rabeya Akter for the ICP anal-
the exposure to the n-butylamine gas. Ultimately, this promotes ysis, Dr. Jason Scott for the BET analysis, and Dr. Bill Gong for the
an enhanced response of the ZnO/Au sensor to n-butylamine com- XPS measurements.
pared to the pure ZnO sensor
Appendix A. Supplementary data
4. Conclusions
Supplementary data associated with this article can be found, in
This study has demonstrated the facile synthesis of ZnO the online version, at http://dx.doi.org/10.1016/j.snb.2016.02.091.
nanoflakes with exposed {1 0 1 0} facets through a hydrothermal
method and their subsequent modifications with Au NPs. The References
effects of the different Au loading on the crystal structure, surface
properties, and amine sensing performance of the ZnO nanoflakes [1] L. Wang, H. Dou, Z. Lou, T. Zhang, Encapsuled nanoreactors (Au@SnO2 ): a new
have been investigated. Several main findings of this study are sum- sensing material for chemical sensors, Nanoscale 5 (2013) 2686–2691.
[2] J. Zhang, X. Liu, S. Wu, B. Cao, S. Zheng, One-pot synthesis of Au-supported
marized below: ZnO nanoplates with enhanced gas sensor performance, Sens. Actuators B 169
(2012) 61–66.
• The pure ZnO nanoflakes exhibit pores with sizes of 2–10 nm and [3] Y.V. Kaneti, J. Yue, X. Jiang, A. Yu, Controllable synthesis of ZnO nanoflakes
with exposed (1 0 0) for enhanced gas sensing performance, J. Phys. Chem. C
the top/bottom surface of these nanoflakes expose the ±{1 0 1̄ 0} 117 (2013) 13153–13162.
plane. After loading the ZnO nanoflakes with 1.37, 2.82, and [4] S. Park, S. An, H. Ko, C. Jin, C. Lee, Synthesis of nanograined ZnO nanowires and
their enhanced gas sensing properties, ACS Appl. Mater. Interfaces 4 (2012)
5.41 wt%, Au NPs with average particle sizes of 7.18, 8.08, and
3650–3656.
8.68 nm, respectively, were successfully deposited on the surface [5] F. Lu, W. Cai, Y. Zhang, ZnO hierarchical micro/nanoarchitectures:
of the ZnO nanoflakes. solvothermal synthesis and structurally enhanced photocatalytic
• The Au nanoparticle-decorated ZnO nanoflakes exhibit a greatly performance, Adv. Funct. Mater. 18 (2008) 1047–1056.
[6] B. Li, Y. Wang, Facile synthesis and enhanced photocatalytic performance of
enhanced selectivity and 4–6 time higher response to n- flower-like ZnO hierarchical microstructures, J. Phys. Chem. C 114 (2009)
butylamine compared to the pure ZnO nanoflakes, with a faster 890–896.
response/recovery time by a factor of 1.5–2.5. Meanwhile, the [7] W.J.E. Beek, M.M. Wienk, R.A.J. Janssen, Efficient hybrid solar cells from zinc
oxide nanoparticles and a conjugated polymer, Adv. Mater. 16 (2004)
surface decoration of Au NPs is found to lower the optimum oper- 1009–1013.
ating temperature of the ZnO nanoflake sensor from 300 to 240 ◦ C. [8] W.J.E. Beek, M.M. Wienk, M. Kemerink, X. Yang, R.A.J. Janssen, Hybrid zinc
Furthermore, even at a low operating temperature of 140 ◦ C, the oxide conjugated polymer bulk heterojunction solar cells, J. Phys. Chem. B
109 (2005) 9505–9516.
ZnO/Au sensors still exhibit a reasonably high response of 8.2–16 [9] D. Bresser, F. Mueller, M. Fiedler, S. Krueger, R. Kloepsch, D. Baither, M. Winter,
to 50 ppm of n-butylamine, indicating their excellent low tem- E. Paillard, S. Passerini, Transition-metal-doped zinc oxide nanoparticles as a
perature amine-sensing properties. new lithium-ion anode material, Chem. Mater. 25 (2013) 4977–4985.
• The DFT simulation results provide new physical insights into [10] Y.V. Kaneti, J. Moriceau, M. Liu, Y. Yuan, Q. Zakaria, X. Jiang, A. Yu,
Hydrothermal synthesis of ternary ␣-Fe2 O3 –ZnO–Au nanocomposites with
the adsorption behaviors of different gases on the ZnO(1 0 1̄ 0) high gas-sensing performance, Sens. Actuators B 209 (2015) 889–897.
surface. n-Butylamine is found to chemically adsorb on the [11] S. Ghosh, C. RoyChaudhuri, R. Bhattacharya, H. Saha, N. Mukherjee,
Palladium–silver-activated ZnO surface: highly selective methane sensor at
ZnO(1 0 1̄ 0) surface through the formation of a bond between
reasonably low operating temperature, ACS Appl. Mater. Interfaces 6 (2014)
the nitrogen atom of the n-butylamine (C4 H11 N) and the surface 3879–3887.
Zn atom of ZnO. Other gases such as ethanol and n-butanol also [12] X. Li, X. Zhou, Y. Liu, P. Sun, K. Shimanoe, N. Yamazoe, G. Lu, Microwave
chemically adsorb on the ZnO(1 0 1̄ 0) surface, with the oxygen hydrothermal synthesis and gas sensing application of porous ZnO core–shell
microstructures, RSC Adv. 4 (2014) 32538–32543.
atoms being covalently bonded to the surface Zn atom of ZnO dur- [13] Z. Jing, J. Zhan, Fabrication and gas-sensing properties of porous ZnO
ing the interaction. However, heptane and methanol molecules nanoplates, Adv. Mater. 20 (2008) 4547–4551.
are only physically adsorbed on the ZnO(1 0 1̄ 0) surface. Hence [14] S. Tian, F. Yang, D. Zeng, C. Xie, Solution-processed gas sensors based on ZnO
nanorods array with an exposed (0 0 0 1) facet for enhanced gas-sensing
the responses of the ZnO nanoflake sensor toward these two gases properties, J. Phys. Chem. C 116 (2012) 10586–10591.
are lower than those to the other tested gases. [15] Y.V. Kaneti, Z. Zhang, J. Yue, Q.M.D. Zakaria, C. Chen, X. Jiang, A. Yu, Crystal
• The enhanced response of the Au nanoparticle-decorated ZnO plane-dependent gas-sensing properties of zinc oxide nanostructures:
experimental and theoretical studies, Phys. Chem. Chem. Phys. 16 (2014)
nanoflakes toward n-butylamine can be attributed to two main 11471–11480.
factors: (i) the high catalytic activity of Au NPs which increases [16] J.F. Chang, H.H. Kuo, I.C. Leu, M.H. Hon, The effects of thickness and operation
the quantity of adsorbed oxygen and the conversion rate of temperature on ZnO:Al thin film CO gas sensor, Sens. Actuators B 84 (2002)
258–264.
oxygen molecules to oxygen ions; and (ii) the existence of an [17] H. Gong, J.Q. Hu, J.H. Wang, C.H. Ong, F.R. Zhu, Nano-crystalline Cu-doped ZnO
additional electron depletion layer at the ZnO/Au interface in thin film gas sensor for CO, Sens. Actuators B 115 (2006) 247–251.
addition to that at the surface of the ZnO nanoflakes. The combi- [18] K. Zheng, L. Gu, D. Sun, X. Mo, G. Chen, The properties of ethanol gas sensor
based on Ti doped ZnO nanotetrapods, Mater. Sci. Eng. B 166 (2010) 104–107.
nation of these two factors promotes a higher change in resistance
[19] J. Haeng Yu, G. Man Choi, Electrical and CO gas sensing properties of
and therefore, a significantly improved response of the ZnO/Au ZnO–SnO2 composites, Sens. Actuators B 52 (1998) 251–256.
sensors toward n-butylamine. [20] F. Meng, N. Hou, Z. Jin, B. Sun, Z. Guo, L. Kong, X. Xiao, H. Wu, M. Li, J. Liu,
Ag-decorated ultra-thin porous single-crystalline ZnO nanosheets prepared
by sunlight induced solvent reduction and their highly sensitive detection of
Overall, the findings of this study will provide a better under- ethanol, Sens. Actuators B 209 (2015) 975–982.
standing of the atomic-scale interactions which occur between [21] Z. Yuan, X. Jiaqiang, X. Pengcheng, Z. Yongheng, C. Xuedong, Y. Weijun,
Decoration of ZnO nanowires with Pt nanoparticles and their improved gas
the ZnO surface and reducing amine/alcohol gases during the gas- sensing and photocatalytic performance, Nanotechnology 21 (2010) 285501.
sensing process. [22] Y. Zhang, Q. Xiang, J. Xu, P. Xu, Q. Pan, F. Li, Self-assemblies of Pd nanoparticles
on the surfaces of single crystal ZnO nanowires for chemical sensors with
enhanced performances, J. Mater. Chem. 19 (2009) 4701–4706.
Acknowledgments [23] S.T. Shishiyanu, T.S. Shishiyanu, O.I. Lupan, Sensing characteristics of
tin-doped ZnO thin films as NO2 gas sensor, Sens. Actuators B 107 (2005)
379–386.
We gratefully acknowledge the financial support of the Aus- [24] F. Paraguay D, M. Miki-Yoshida, J. Morales, J. Solis, W. Estrada L, Influence of
tralian Research Council (ARC DP1096185, FT0990942) projects. Al, In Cu, Fe and Sn dopants on the response of thin film ZnO gas sensor to
The authors also acknowledge access to the UNSW node of the Aus- ethanol vapour, Thin Solid Films 373 (2000) 137–140.
Y.V. Kaneti et al. / Sensors and Actuators B 230 (2016) 581–591 591

[25] K. Kim, J.W. Lee, D. Shin, J.-Y. Choi, K.S. Shin, Organic isocyanide-adsorbed [42] A. Kolmakov, D.O. Klenov, Y. Lilach, S. Stemmer, M. Moskovits, Enhanced gas
gold nanostructure: a SERS sensory device for indirect peak-shift detection of sensing by individual SnO2 nanowires and nanobelts functionalized with Pd
volatile organic compounds, Analyst 137 (2012) 1930–1936. catalyst particles, Nano Lett. 5 (2005) 667–673.
[26] J. Panchompoo, L. Aldous, L. Xiao, R.G. Compton, Electroanalytical detection of [43] S.M. Majhi, P. Rai, Y.-T. Yu, Facile approach to synthesize Au@ZnO core–shell
n-butylamine at a nickel/carbon nanotube composite, Electroanalysis 22 nanoparticles and their application for highly sensitive and selective gas
(2010) 912–917. sensors, ACS Appl. Mater. Interfaces 7 (2015) 9462–9468.
[27] H. Fu, X. Yang, X. Jiang, A. Yu, Silver vanadate nanobelts: a highly sensitive
material towards organic amines, Sens. Actuators B 203 (2014) 705–711.
[28] I. Raible, M. Burghard, U. Schlecht, A. Yasuda, T. Vossmeyer, V2O5 nanofibres: Biographies
novel gas sensors with extremely high sensitivity and selectivity to amines,
Sens. Actuators B 106 (2005) 730–735.
[29] X. Han, X. Han, L. Li, C. Wang, Controlling the morphologies of WO3 particles Dr. Yusuf Valentino Kaneti has obtained his Ph.D. degree in July 2014 from the
and tuning the gas sensing properties, New J. Chem. 36 (2012) 2205–2208. University of New South Wales (UNSW), Australia. He is currently working as a post-
[30] M. Breedon, M.J.S. Spencer, I. Yarovsky, Adsorption of NO and NO2 on the doctoral fellow in the School of Materials Science & Engineering at UNSW, focusing
ZnO(2110) surface: a DFT study, Surf. Sci. 603 (2009) 3389–3399. on the synthesis and characterization of semiconductor metal oxide nanostruc-
[31] M.J.S. Spencer, K.W.J. Wong, I. Yarovsky, Density functional theory modelling tures/nanocomposites for energy and environmental applications. He has published
of and surfaces: structure, properties and adsorption of N2 O, Mater. Chem. 14 journal articles and 1 book chapter, with a H-index of 8.
Phys. 119 (2010) 505–514. Mr. Xiao Zhang received his undergraduate degree in materials science and engi-
[32] J.D. Prades, A. Cirera, J.R. Morante, Ab initio calculations of NO2 and SO2 neering from the University of New South Wales, Australia. His research interest
chemisorption onto non-polar ZnO surfaces, Sens. Actuators B 142 (2009) is on the synthesis and functional application of metal oxide nanostructures for
179–184. gas-sensing applications.
[33] W. An, X. Wu, X.C. Zeng, Adsorption of O2 , H2 , CO, NH3 , and NO2 on ZnO
nanotube: a density functional theory study, J. Phys. Chem. C 112 (2008) Mr. Minsu Liu is a Ph.D. candidate studying materials science and engineering at
5747–5755. Monash University, Australia. His research interest is on the synthesis of metal
[34] M. Breedon, M.J.S. Spencer, I. Yarovsky, Adsorption of NO2 on oxygen deficient oxide (especially for vanadium oxides) nanostructures for energy and environmen-
ZnO(2110) for gas sensing applications: a DFT study, J. Phys. Chem. C 114 tal applications.
(2010) 16603–16610.
[35] X.-G. Han, H.-Z. He, Q. Kuang, X. Zhou, X.-H. Zhang, T. Xu, Z.-X. Xie, L.-S. Zheng, Mr. David Yu is an M.Sc. candidate studying materials science and engineering at
Controlling morphologies and tuning the related properties of Monash University, Australia. His research interest is on the synthesis and applica-
nano/microstructured ZnO crystallites, J. Phys. Chem. C 113 (2008) 584–589. tion of magnetic nanoparticles for environmental and biomedical applications.
[36] Y.V. Kaneti, J. Yue, J. Moriceau, C. Chen, M. Liu, Y. Yuan, X. Jiang, A. Yu,
Ms. Yuan Yuan is a Ph.D. candidate studying materials science and engineering at the
Experimental and theoretical studies on noble metal decorated tin oxide
University of New South Wales, Australia. Her research interest is on the synthesis
flower-like nanorods with high ethanol sensing performance, Sens. Actuators
of shape-controlled noble metal nanostructures for catalytic applications.
B 219 (2015) 83–93.
[37] W.L. Ong, S. Natarajan, B. Kloostra, G.W. Ho, Metal nanoparticle-loaded Dr. Leigh Aldous is an ARC DECRA fellow and he works as a lecturer/researcher in
hierarchically assembled ZnO nanoflakes for enhanced photocatalytic the School of Chemistry at the University of New South Wales. Dr. Aldous’ research
performance, Nanoscale 5 (2013) 5568–5575. focuses upon electrochemistry in ionic liquids (for biomass conversion, hydro-
[38] H. You, R. Liu, C. Liang, S. Yang, F. Wang, X. Lu, B. Ding, Gold nanoparticle gen storage and unique physical characteristics), as well as electroanalysis using
doped hollow SnO2 supersymmetric nanostructures for improved nanocarbons. Dr. Aldous has published more than 67 journal articles and 2 patents
photocatalysis, J. Mater. Chem. A 1 (2013) 4097–4104. in the past 6 years with SCI citations over 1800 times, leading to a H-index of 26.
[39] N.I. Butkovskaya, Y. Zhao, D.W. Setser, Decomposition of chemically activated
ethanol, J. Phys. Chem. 98 (1994) 10779–10786. Prof. Xuchuan Jiang has fully devoted to the study on synthesis and self-assembly
[40] W. Wang, C. Zhu, Y. Cao, DFT study on pathways of steam reforming of of nanoparticles for energy and environmental applications since the award of his
ethanol under cold plasma conditions for hydrogen generation, Int. J. Ph.D. in 2001. His research focuses on developing new techniques for the synthesis
Hydrogen Energy 35 (2010) 1951–1956. of noble metals, metal oxides and their nanocomposites for energy, environmental
[41] G.F. Bauerfeldt, L.M.M. de Albuquerque, G. Arbilla, E.C. da Silva, Unimolecular and biomedical applications. He has published over 100 papers with SCI citations of
reactions on formaldehyde S0 PES, THEOCHEM 580 (2002) 147–160. over 4000 times, leading to a H-index of 29.

You might also like