You are on page 1of 10

Fabrication and characterization of SnO2

nanorods for room temperature gas sensors


Cite as: AIP Advances 8, 095219 (2018); https://doi.org/10.1063/1.5050991
Submitted: 04 August 2018 . Accepted: 04 September 2018 . Published Online: 20 September 2018

Amrit P. Sharma, Pashupati Dhakal, Dhiren K. Pradhan, et al.

ARTICLES YOU MAY BE INTERESTED IN

Resistive gas sensors based on metal-oxide nanowires


Journal of Applied Physics 126, 241102 (2019); https://doi.org/10.1063/1.5118805

Structural and photoluminescence properties of tin oxide and tin oxide: C core–shell and
alloy nanoparticles synthesised using gas phase technique
AIP Advances 6, 095321 (2016); https://doi.org/10.1063/1.4964313

Gas sensing in 2D materials


Applied Physics Reviews 4, 021304 (2017); https://doi.org/10.1063/1.4983310

AIP Advances 8, 095219 (2018); https://doi.org/10.1063/1.5050991 8, 095219

© 2018 Author(s).
AIP ADVANCES 8, 095219 (2018)

Fabrication and characterization of SnO2 nanorods


for room temperature gas sensors
Amrit P. Sharma,1,a Pashupati Dhakal,2 Dhiren K. Pradhan,3
Makhes K. Behera,1 Bo Xiao,1 and Messaoud Bahoura1
1 Centerfor Materials Research, Norfolk State University, Norfolk, Virginia 23504, USA
2 Jefferson
Lab, Newport News, Virginia 23606, USA
3 Geophysical Laboratory, Carnegie Institution for Science, Washington, D.C. 20015, USA

(Received 4 August 2018; accepted 4 September 2018; published online 20 September 2018)

Highly sensitive large-scale tin oxide (SnO2 ) nanostructures were grown on a glass
substrate by thermal evaporation of a mixture of anhydrous tin (II) chloride (SnCl2 )
and zinc chloride (ZnCl2 ) powders at 550◦ C in air. We demonstrate a single cell
vapor deposition system to precisely control nanostructural morphology of SnO2
by changing the weight ratio of SnCl2 and ZnCl2 and growth temperature. The
morphology and structural property of as-grown nanostructures were characterized
using scanning electron microscopy (SEM) and X-ray diffraction (XRD). The SEM
images revealed that the SnO2 nanostructures with different densities, sizes, and
shapes can be achieved by adjusting the weight ratio of SnCl2 and ZnCl2 . A thin
film gas sensor based on SnO2 nanostructures with diameter ∼20 nm and length
∼100 nm showed ∼85% sensitivity and 53 seconds of response time, whereas the
nanorods with diameter ∼100 nm and length ∼ 1µm showed ∼50% sensitivity with
198 seconds response time. The nanostructured material with small size and shape
showed better sensitivity on sensing at room temperature compared to previously
reported SnO2 based sensors. © 2018 Author(s). All article content, except where
otherwise noted, is licensed under a Creative Commons Attribution (CC BY) license
(http://creativecommons.org/licenses/by/4.0/). https://doi.org/10.1063/1.5050991

I. INTRODUCTION
With an increase in the advancements in the modern sensing technology, as the numbers of fab-
rication and manufacturing techniques are innovated, there is also an increase in various negative
impacts such as an increase in contaminants, harmful gases, and energy consumption. There-
fore, it makes it imperative that research efforts need to be focused not only on improving the
modern world technology but also needs to be focused on mitigating the harmful impacts of the
technological and industrial advancements. Being able to accurately sense the harmful gases at a
lower operating temperature and voltage is one of the most important steps associated with the
solutions.
Metal oxide semiconductors such as SnO2 , TiO2 , ZnO, WO2 , In2 O3 , and CuO have attracted
significant attention in detecting toxic, harmful, and combustible gases.1–8 Among these materials,
SnO2 , an n-type semiconductor with wide band gap (3.62 eV), is one of the most popular materi-
als being used in sensing number of applications from health and safety to energy efficiency and
emission control9 because of high sensitivity, high crystallinity, thermal/chemical stability, tunable
morphology, simple design, and low cost.4,6,7 Their main drawbacks remains the low selectivity and
relatively high operating temperature, which limits their long-term stability, use in hazardous and
explosive environments, possibility to reduce their power consumption and direct implementation
into electronic circuit.10,11

a
Corresponding author. E-mail: a.p.sharma@spartans.nsu.edu (Amrit P Sharma)

2158-3226/2018/8(9)/095219/9 8, 095219-1 © Author(s) 2018


095219-2 Sharma et al. AIP Advances 8, 095219 (2018)

Under ambient conditions, there is significant oxygen adsorption on the surface of SnO2 and
each oxygen atom accepts an electron from its surface to complete the bond, thereby decreasing the
electron density and increasing resistance. The oxygen only adsorbs onto the surface and thus the
electrons are only removed to a certain depth from the surface known as Debye length, typically on
the order of 2-100 nm.9 Those abundant oxygen vacancies on the surface of SnO2 makes it electrically
and chemically active and hence is weakly conductive. When SnO2 is exposed to the reducing agent
such as ethanol, its active surface with oxygen adsorbates reacts with the gas and enable the electrons
to the active material, causing an increase in electronic conductivity, whereas the effect is reversed
for oxidizing agents.12–14 The SnO2 based gas sensors are used in sensing alcohol, H2 , O2 , H2 S, NO2 ,
and CO.1 Alcohol sensors are widely used to detect ethanol levels on breath tests for safety.15,16
Recently, a lot of research has been concentrated on improving sensitivity, selectivity, stability,
response time, cost and reducing operating temperature of the gas sensors. The sensing mechanism
in SnO2 is based on the chemisorption reaction that takes place at the surface of the metal oxide, so
enhancing surface area of the sensitive materials leads to more sites for adsorption of target gases.17,18
In most cases, gas sensors are sensitive only at high temperature (above 150◦ C) but the operating
temperature can be greatly reduced with sensor geometry and with elaborately designing nanostruc-
tured sensing materials.12 The strength of the Vander Walls bond between the material and the gas
is the most important factor which determines the operating temperature of the sensors.9 Moreover,
some volatile organic compounds (such as ethanol and acetone) have a high vapor pressure even at
room temperature. Therefore, their detection at room temperature open a lot of potential applications
in room temperature gas sensing. Most materials behave differently at nanoscale compared to their
bulk counterparts.19,20 It is well known that the particle size and morphology of the materials have
great influence in their unusual thermodynamic and kinetic behaviors and play important role to form
nano-sized domain morphologies. Therefore, the synthesis of nanostructure with controlled size and
morphologies may open new opportunities for exploring materials’ chemical and physical proper-
ties. Nanostructured materials with small size and shape have large surface area to volume ratio such
that more atoms residing on the materials’ surface sites to facilitate surface reactions.21 Gas sensors
operating at high temperatures may cause coalescence and structural changes22 due to a thermally
induced grain growth, which decreases the sensor’s stability and life time. Hence, operating at room
temperatures can inhibit structural changes, reduce power consumption, and enable safer detection
of flammable or explosive gases. The room temperature operation of the gas sensor removes the
requirement of heating elements, thereby gaining practical flexibility to realize cost-effective and
miniature devices. In this view, more attention should be paid to design and synthesize material to
sense at room temperature.
Sensing performance depends on the diffusion of gas throughout the pores of the sensing layers
which can be improved by engineering the surface properties (e.g. porosity, shape, surface area to
volume ratio, etc.).23 Majority of the research has been focused on controlling SnO2 nanostructure
as well as fabrication techniques to increase its performance. It has been reported that the perfor-
mance of SnO2 nanomaterials are greatly affected by their morphology, structure and surface area.24
Therefore, the nanostructured materials with different dimensionality and shape would offer good
opportunities to explore new physical and chemical properties. Guo et al.25 synthesized 3D SnO2
microstructures assembled by porous nano-sheets through hydrothermal process, exhibited good
response-recovery performance, high sensitivity, excellent long term stability but working at high tem-
perature 300◦ C for ethanol. Similarly, Zn-doped SnO2 microporous hollow spheres by Wang et al.26
and Zn-doped SnO2 nanostructure by Zhao et al.24 reported their operating temperature at 240◦ C.
Moreover, J. Xu et al.27 exhibited the SnO2 -based nanorods display much better gas sensing proper-
ties than the hollow spheres, which is due to their higher rate of electron transfer and larger surface
area for adsorption and diffusion. Recently, Meng et al.28 demonstrated the enhanced performance
of Au/SnO2 /RGO nanocomposites towards 100 ppm of ethanol at 200◦ C with high reproducibility.
A controlled growth of the SnO2 nanostructure will enable the films to be fine-tuned for different accu-
racy, operating temperature and sensitivity. Here, we demonstrate the large-scale synthesis of highly
reactive SnO2 nanostructures, engineering of their surface properties by varying the weight ratio of
SnCl2 and ZnCl2 and their utilization in the fabrication of high performance room temperature ethanol
sensors.
095219-3 Sharma et al. AIP Advances 8, 095219 (2018)

FIG. 1. The schematic of the samples synthesis.

II. EXPERIMENTAL DETAILS


A. Synthesis of large-scale tin oxide nanostructures
The schematic diagram of experimental set up for the synthesis of highly sensitive SnO2 nanos-
tructure is shown in Fig. 1. In a typical experiment, SnO2 nanostructures were synthesized by low
temperature vapor-solid (VS) process.19,29 The source materials, powders of anhydrous SnCl2 (99%
purity, Alfa Aesar) and ZnCl2 (99.95% purity, Alfa Aesar) were mixed in different ratios but kept
the total weight of the mixture the same as 1 gram. We use off-the-shelf glass slides as the sub-
strates for the deposition. A general degrease cleaning process (acetone, isopropanol and deionized
water) was performed before loading the substrate in the growth cell, where the mixture of the source
powder was filled in an alumina ceramic cup. The distance between the source and substrate is
∼4 cm. The growth cell (Fig. 1) is then transferred into a commercial muffle furnace. The furnace
was heated gradually in presence of air at a rate of 12◦ C/min and then held at 550◦ C for 10 min-
utes before cooling down at the rate of 20◦ C/min. Finally, a grey product was grown on the glass
substrate.

B. Device fabrication
Two electrodes separated by ∼5 mm was made on the surface of the SnO2 nanostructures using
silver paste and copper wires were attached to the electrodes as illustrated in Fig. 2. The fabricated
sensor was placed inside a quartz tube of length 40 cm and diameter 3 cm. Fig. 3 sketches out the
diagram of the experimental set up. An ultra-high purity argon gas was flowing through a round
bottom flask containing 75 ml of ethanol at temperature ∼50◦ C in atmospheric pressure. The argon
gas was used for two purposes; to clear out gas and products from the quartz tube, and to send ethanol
vapors over the sensor as a carrier gas. A fixed low concentration of test gas that is the mixture of

FIG. 2. Schematic diagram of gas sensor fabrication process.


095219-4 Sharma et al. AIP Advances 8, 095219 (2018)

FIG. 3. Gas sensing experimental set up.

ethanol vapors and argon was introduced in the sensing chamber via mass flow controller (MKS-
MFC). Oxygen gas was introduced for the full recovery of the sensor before the next sensing cycle.
A steady current of 10 µA was supplied from Keithley 6220 and the voltage was monitored using
Keithley 2182A nanovoltmeter during the experiments. All measurements were carried out at room
temperature and the response of each sensor was obtained from the resistance versus time plots. The
response of the sensor (η) is defined as:
Ra − Rg
η= × 100% (1)
Ra
where Ra and Rg are the resistance of the sensor in the presence of the argon gas and test gas,
respectively.

III. RESULTS AND DISCUSSION


A. Growth mechanism
There are two well-accepted mechanism for the growth of 1-D nanostructure: the vapor-liquid-
solid (VLS) and vapor-solid (VS) mechanism. In our study, no metal catalyst was used, and no
liquid droplet was found at the tip of the nanorods. Therefore, the growth mechanism is VS process.
During the vapor reaction, the ZnCl2 and SnCl2 powders were vaporized to form their corresponding
vapors (equations (2) and (3)). The SnCl2 vapor reacted with H2 O vapor and O2 to form SnO2
vapor (equation (4)). However, the ZnCl2 powders were only vaporized to form ZnCl2 vapors and
act as an interspace separator between SnO2 nuclei (Fig. 4a), it did not react with H2 O because
Sn2+ hydrolyzes easily compared with Zn2+ during the reaction.22 The chemical reactions can be
formulated as follows:
ZnCl2 (S) → ZnCl2 (g) (2)
SnCl2 (S) → SnCl2 (g) (3)
2SnCl2 (g) + 2H2 O + O2 → 2SnO2 + 4HCl (4)
The vapors of SnO2 were collected on the glass substrate in the form of distinct spherical nanoparticles
and they become the preferred site for deposition (Fig. 4b). The supersaturation of these nanoparticles
helps to grow the nanorods because they serve as seeds for the growth of the nanorods.
As a hydrolytic agent, ZnCl2 , its concentration controls the hydrolysis rate of Sn2+ and formation
of water vapor ions (OH- ).30 Both the concentration of Sn2+ and OH- ions play an important role in
the formation of different size/shape SnO2 nanorods. When the weight ratio of SnCl2 and ZnCl2 is
1:1, the nucleation rate of SnO2 is less because of the formation of limited concentration of OH- ions
095219-5 Sharma et al. AIP Advances 8, 095219 (2018)

FIG. 4. Schematic diagram of the growth mechanism of SnO2 nanorods by VS process.

which restrains the hydrolysis of the Sn2+ . As a result, bigger sized nanorods are formed (Fig. 4c). As
the weight ratio changes to 1:3, keeping the total amount of the mixture fixed at 1 g, the concentration
of OH- ion is higher due to higher amount of ZnCl2 . The higher concentration of OH- assists in the
hydrolysis of Sn2+ and hence many SnO2 nanoparticles with smaller sizes (Fig. 4d) are formed. On
the other hand, at weight ratio 1:4, even though the concentration of OH- is higher, the concentration
of Sn2+ is lower because of lower amount of SnCl2 , as the total amount of the mixture is fixed
(1 g). The lower concentration of Sn2+ decreases the nucleation rate of SnO2 nanoparticles. As a
result, larger size of SnO2 nanorods with different shape (Fig. 4e) are formed. This growth mechanism
indicates that the appropriate weight ratio of SnCl2 and ZnCl2 is 1:3 for the growth of ultra-thin SnO2
nanorods.
B. Structural properties
Surface morphology and structural properties of as-grown nanostructures were characterized
using FESEM as shown in Fig. 5(a–c). The control over the SnO2 nanostructures with different
densities, sizes, and shapes was achieved by adjusting the weight ratio of SnCl2 and ZnCl2 . At
weight ratio 1:1, the shapes of the structures were found to be nanorods with diameter ∼100 nm

FIG. 5. FESEM images of SnO2 nanostructures with different ratio of SnCl2 and ZnCl2 (a) 1:1, (b) 1:3 and (c) 1:4.
095219-6 Sharma et al. AIP Advances 8, 095219 (2018)

and length ∼1 µm (Fig. 5a). As the weight ratio increases to 1:3, significant reduction in dimensions
(shape and size) of the resultant nanostructures was observed with diameter and length ∼20 nm and
∼100 nm, respectively, as shown in Fig. 5(b). With a further increase in weight of ZnCl2 , the shape
and the size of the resulting nanostructures increases, and the diameter and length were found to
be ∼50 nm and ∼1 µm, respectively, (Fig. 5c). The FESEM images reveal that SnO2 nanorods with
different sizes, shapes, and densities can be realized by the weight ratio of SnCl2 and ZnCl2 , which
further confirms the proposed growth mechanism. This result indicates that the appropriate ratio of
SnCl2 and ZnCl2 is 1:3 for the growth of ultra-thin SnO2 nanorods which have a larger surface area
to volume ratio, high specific surface area, and more surface active sites that play a significant role
in enhancing the functionalities of the fabricated sensing layer.12,21,31
Figure 6 shows XRD patterns of SnO2 on a glass substrate with different weight ratios of SnCl2
and ZnCl2 along with one peak of SnO on 1:1 ratio due to deficiency of oxygen. This reveals that all
the samples have a strong peak on the (101) plane. The enhanced (002) peak indicates that the crystal
grow along the [001] direction and the nanorods are well crystallized.32

C. Sensing application
The upper limit of resistance in our measurement was established by introducing pure argon gas
into the sensing tube. The samples grown at weight ratios of 1:1, 1:4 and 1:3 were labelled as sensor
1, sensor 2, and sensor 3 respectively. Time-dependent of change in resistance measurement for all
the samples are shown in Fig. 7. The results of the experiments, at room temperature, are summarized
in the Table I.
Carvajal et al.19 have reported a response of ∼42% in a sample with SnCl2 and ZnCl2 weight
ratio of 1:4 synthesized on Si substrate where the sensing response was measured at 250◦ C sensor
temperature. Our study showed a response of >70% with 1:4 ratio and the response was found to be as
high as 85% when the ratio was kept at 1:3 sensing at room temperature. The conduction mechanism
is solely dependent on nanostructures since the substrate was made of glass. The response curve was
fitted with (Rg /Ra ) ∼ e-t/τ to calculate the response time (τ) and found to be 53 seconds for sensor
3 and 198 seconds for sensor 1.
To demonstrate the reproducibility, the response of sensor 3 with 1:3 SnCl2 and ZnCl2 ratio for
different gas flow rates, is shown in Fig. 8. A fast response was observed with the first 5 sccm flow rate
as shown in Fig. 8(a). An increase in ethanol concentration provides more ethanol molecules to be
absorbed on the material surface, thereby favoring more electron transport kinetics which increases
its response. The response tends to reach its maximum value at 20 sccm and saturates thereafter. The
saturation can be attributed to a deficiency in ionosorbed oxygen species, which can support kinetic

FIG. 6. XRD patterns of SnO2 nanostructures on a glass substrate with different ratios of SnCl2 and ZnCl2 .
095219-7 Sharma et al. AIP Advances 8, 095219 (2018)

FIG. 7. The response of the ethanol with time for sensors 1-3 with 20 sccm flow rate of ethanol and argon.

TABLE I. Comparison of resistance measurements carried out on samples with different surface morphologies

Sample Weight ratio of SnCl2 and ZnCl2 Test gas Carrier gas Response (η) %

sensor 1 1:1 ethanol argon >50


sensor 2 1:4 ethanol argon >70
sensor 3 1:3 ethanol argon >85

mechanisms. For a given concentration of ethanol, the sensing mechanism is highly reproducible as
shown in Fig. 8(b).
SnO2 is n-type semiconductor with oxygen deficiency. The lattice oxygen is evaporated in the
form of gas, which makes the doubly ionized oxygen vacancy and trap electrons.20 The oxygen
ions absorbed on the surface of the SnO2 can forms a depletion layer, which leads to a decrease
in the electronic conductivity.33,34 Once exposed to reducing gas, the active substrate with oxygen
adsorbates will react with the gas and enable the electrons injected back into the active material,
causing an increase in electronic conductivity as illustrated in Fig. 9. The chemical mechanism
involved can be described by the following reactions:
O2 (gas) → O2 (adsorbed) (5)

O2 (adsorbed) + e− ↔ O2 (6)
− −
C2 H5 OH + 3O2 ↔ 3H2 O + 2CO2 + 3e (7)

FIG. 8. (a) The response as a function of time for different flow rates and (b) the reproducibility of the sensing application
for flow rate 20 sccm measured with sensor 3.
095219-8 Sharma et al. AIP Advances 8, 095219 (2018)

FIG. 9. Schematic model of oxygen ionosorption on SnO2 surface with ethanol exposure.

The ability to control and manipulate the shape and size of a material offers to tailor its various
properties. As the size reduces to nanoscale dimension, the surface-to-volume ratio increases which
generates more reactivity due to a dominant surface-like behavior caused by a large fraction of
atoms at the surface. At this scale, the surface features, such as defects and disorders also become
quite distinct and play important role in determining the physical or chemical properties of the
nanomaterials. Absorbed molecules on the surface due to this defect and disorders in the lattice
plane changes its surface energy.35 As a result, at a nanoscale dimension, the surface area and surface
energy increases tremendously. Therefore, sensor 3 with small size and shape demonstrates its highest
sensitivity as compared to sensors 1 and 2, which can be attributed as follows:
Firstly, the sensitivity of a gas sensor is
S = A(C)N (8)
where A is constant, C is gas concentration.36 The value of N is either 1 or 1/2. N=1, if the size
of SnO2 is comparable to space-charge-layer (less than 20 nm) and N = 1/2, if its size is greater
than 20 nm.37,38 As the size of the nanorods are less than 20 nm, its sensitivity is linearly increased
with gas concentration (equation (8)) which makes it highly sensitive. Secondly, when the size of
the nanostructure is less than 20 nm, the conductivity is essentially controlled by the intercrystallite
conductivity.37 Only fewer charges acquired from the surface reactions will cause large changes of
conductivity of the whole structure. Therefore, the higher the crystallinity of the SnO2 , the higher
will be the sensitivity to ambient gas molecules when its particle size is small. Furthermore, when
the crystallite dimensions are driven below 20 nm, all atoms will be within a Debye length of the
surface and hence, oxygen vacancies will be created throughout the material which greatly enhance
its response. Hence, the sensitivity (response) of the gas sensors can be enhanced by the superior
control over morphology of the SnO2 nanostructures.

IV. CONCLUSIONS
We have fabricated SnO2 nanorod-based gas sensors on glass substrate which demonstrated a
performance with ∼85% response with respect the ethanol gas. The preferential growth direction is
determined by surface energy, whereas the morphology is determined by the growth kinetics and
precursor materials. The morphology of nanostructure can be controlled by the weight ratio of SnCl2
and ZnCl2 and the response depends on the morphology of the nanostructures. The SnO2 nanostructure
with 1:3 ratio showed the highest response with the response time 53 sec at room temperature as
the nanostructure with small shape and size are more reactive and hence more sensitive. This sensor
demonstrated superiority over previously reported SnO2 -based gas sensors in terms of its operation
at room temperature. The ability to be miniaturized gives the fabricated sensors the capability to be
integrated with the existing technologies such as cell phones or other portable electronic devices. The
sensors may be useful in low power sensing of several other toxic, flammable, or explosive gases
besides ethanol at room temperature.

ACKNOWLEDGMENTS
We would like to acknowledge Dr. Sangram K. Pradhan at Norfolk State University and Shankar
Karki from Old Dominion University, Norfolk VA for useful discussions. This work was supported
095219-9 Sharma et al. AIP Advances 8, 095219 (2018)

by NSF-CREST (CREAM) grant number HRD 1547771 and NSF-CREST (CNBMD) grant number
HRD 1036494.
1 S. M. Ingole, S. T. Navale, Y. H. Navale, D. K. Bandgar, F. J. Stadler, R. S. Mane, N. S. Ramgir, S. K. Gupta, D. K. Aswal,
and V. B. Patil, J. Colloid Interface Sci. 493, 162 (2017).
2 S. T. Navale, V. V. Jadhav, K. K. Tehare, R. U. R. Sagar, C. S. Biswas, M. Galluzzi, W. Liang, V. B. Patil, R. S. Mane, and

F. J. Stadler, Sensors Actuators, B Chem. 238, 1102 (2017).


3 N. Yamazoe and N. Miura, Sensors Actuators B. Chem. 20, 95 (1994).
4 S. Gubbala, V. Chakrapani, V. Kumar, and M. K. Sunkara, Adv. Funct. Mater. 18, 2411 (2008).
5 S. T. Navale, D. K. Bandgar, S. R. Nalage, G. D. Khuspe, M. A. Chougule, Y. D. Kolekar, S. Sen, and V. B. Patil, Ceram.

Int. 39, 6453 (2013).


6 Q. Zhao, Y. Gao, X. Bai, C. Wu, and Y. Xie, Eur. J. Inorg. Chem.1643 (2006).
7 J. F. McAleer, P. T. Moseley, J. O. W. Norris, D. E. Willams, P. Taylor, and B. C. Tofield, Mater. Chem. Phys. 17, 577

(1987).
8 O. A. Sahraei, A. Khodadadi, Y. Mortazavi, M. V. Naseh, and S. Mosadegh, Mater. Chem.185 (2009).
9 D. R. Miller, S. A. Akbar, and P. A. Morris, Sensors Actuators, B Chem. 204, 250 (2014).
10 M. Gardon, O. Monereo, S. Dosta, G. Vescio, A. Cirera, and J. M. Guilemany, Surf. Coatings Technol. 235, 848 (2013).
11 S. Zhan, D. Li, S. Liang, X. Chen, and X. Li, Sensors (Switzerland) 13, 4378 (2013).
12 J. Zhang, X. Liu, G. Neri, and N. Pinna, Adv. Mater. 28, 795 (2016).
13 D. E. Williams, Sensors Actuators, B Chem. 57, 1 (1999).
14 N. Barsan, D. Koziej, and U. Weimar, Sensors Actuators, B Chem. 121, 18 (2007).
15 M. Righettoni, A. Amann, and S. E. Pratsinis, Mater. Today 18, 163 (2015).
16 C. G. Carvajal, K. Kadri, G. Rutherford, R. Mundle, and A. K. Pradhan, ECS J. Solid State Sci. Technol. 4, S3038 (2015).
17 R. M. Walton, D. J. Dwyer, J. W. Schwank, and J. L. Gland, Appl. Surf. Sci. 125, 199 (1998).
18 M. I. Baraton and L. Merhari, Nanostructured Mater. 10, 699 (1998).
19 C. G. Carvajal, C. Snow-Davis, R. Mundle, and A. K. Pradhan, J. Electrochem. Soc. 161, B3151 (2014).
20 J. Pan, H. Shen, and S. Mathur, J. Nanotechnol. 2012.
21 A. Gurlo, Nanoscale 3, 154 (2011).
22 E. Comini, Anal. Chim. Acta 568, 28 (2006).
23 C. Li, M. Lv, J. Zuo, and X. Huang, Sensors 15, 3789 (2015).
24 Q. Zhao, D. Ju, X. Deng, J. Huang, B. Cao, and X. Xu, Sci. Rep. 5 (2015).
25 J. Guo, J. Zhang, D. Ju, H. Xu, and B. Cao, Powder Technol. 250, 40 (2013).
26 W. Wang, Y. Tian, X. Li, X. Wang, H. He, Y. Xu, and C. He, Appl. Surf. Sci. 261, 890 (2012).
27 J. Xu, D. Wang, L. Qin, W. Yu, and Q. Pan, Sensors Actuators, B Chem. 137, 490 (2009).
28 F. Meng, H. Zheng, Y. Chang, Y. Zhao, M. Li, C. Wang, Y. Sun, and J. Liu, IEEE Trans. Nanotechnol. 17, 212 (2018).
29 X. Wang, W. Liu, H. Yang, X. Li, N. Li, R. Shi, H. Zhao, and J. Yu, Acta Mater. 59, 1291 (2011).
30 D. F. Zhang, L. D. Sun, J. L. Yin, and C. H. Yan, Adv. Mater. 15, 1022 (2003).
31 X. Han, M. Jin, S. Xie, Q. Kuang, Z. Jiang, Y. Jiang, Z. Xie, and L. Zheng, Angew. Chemie - Int. Ed. 48, 9180 (2009).
32 X. Zhou, W. Fu, H. Yang, Y. Mu, J. Ma, L. Tian, B. Zhao, and M. Li, Mater. Lett. 93, 95 (2013).
33 Q. Zhao, D. Ju, and X. Deng, Sensors Actuators, B Chem. 15, 1025 (2017).
34 S. Basu, Y. H. Wang, C. Ghanshyam, and P. Kapur, Bull. Mater. Sci. 36, 521 (2013).
35 X. Ding, D. Zeng, and C. Xie, Sensors Actuators, B Chem. 149, 336 (2010).
36 G. Xi and J. Ye, Inorg. Chem. 49, 2302 (2010).
37 S. Das and V. Jayaraman, Prog. Mater. Sci. 66, 112 (2014).
38 A. Rothschild and Y. Komem, J. Appl. Phys. 95, 6374 (2004).

You might also like