Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

G Model

JIEC 3043 1–22

Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


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

1 Review

2 A review of recent developments in tin dioxide composites for gas


3 sensing application
4 Q1 J.P. Chenga,* , Jiao Wanga , Q.Q. Lia , H.G. Liub , Y. Lic,*
5 a
State Key Laboratory of Silicon Materials, School of Materials Science & Engineering, Zhejiang University, Hangzhou 310027, China
6 b
School of Materials Science and Engineering, Northwestern Polytechnical University, Xi'an 710072, China
7 c
Department of Metallurgical and Materials Engineering (MTE), The University of Alabama, Tuscaloosa, AL 35487, USA

A R T I C L E I N F O A B S T R A C T

Article history:
Received 14 June 2016 SnO2 has been extensively investigated and used to detect a variety of gases for practical application.
Received in revised form 11 August 2016 SnO2 nanomaterials with different morphologies and spatial assemblies have been fabricated in the last
Accepted 12 August 2016 few years in order to improve the gas sensing performances. Meanwhile, many reports on the fabrication
Available online xxx and gas sensing research using SnO2-based composites have been also published recently. In this work,
we reviewed the recent developments of conductivity type of gas sensors for various SnO2-based
Keywords: composites, including SnO2/inorganic metal oxide, SnO2/carbon nanomaterials, SnO2/noble metals,
Tin dioxide SnO2/polymer, and SnO2/other materials in the last five years. Most of reports demonstrated that using a
Composite
composite the properties of the gas sensing material could be greatly improved, such as high sensitivity,
Heterostructures
low working temperature, quick response, excellent stability or low detection limit. Each component had
Q2 Carbon nanomaterials
Noble metal its unique effect to influence the sensing properties of the composite. The possible development
Gas sensor directions were also discussed.
ã 2016 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

Contents

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Gas sensing mechanism of SnO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
SnO2 composite gas sensing materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
SnO2/inorganic metal oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Zero-dimensional heterojunctions or spherical heterostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
One-dimensional heterostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Two-dimensional heterostructured composite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Three-dimensional hierarchical heterostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
SnO2/carbon materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
CNTs/SnO2-based gas sensing materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Graphene/SnO2-based gas sensing materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
SnO2/noble metals sensing materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Pd/SnO2 sensing materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Pt/SnO2 sensing materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Au/SnO2 sensing materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Other metals and their mixture with SnO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
SnO2/polymer sensor materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Composite of SnO2/other materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Conclusions and outlooks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

* Corresponding authors.
E-mail addresses: chengjp@zju.edu.cn (J.P. Cheng), yuan.li1@northwestern.edu (Y. Li).

http://dx.doi.org/10.1016/j.jiec.2016.08.008
1226-086X/ã 2016 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

2 J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx

8 Introduction cost and high sensitivity. The sensing material is very important for 71

sensors and is usually deposited as a polycrystalline film or layer on 72


9 Gas sensing technology is receiving much attention in the a substrate with integrated electrodes and heating to make a gas 73
10 practical applications including industrial and academic fields. It sensor. Some typical important properties for a gas sensor are its 74
11 chiefly involves discriminating and monitoring the concentration sensitivity, operating temperature, selectivity and long-term 75
12 of gases in the industrial processes as well as in everyday life [1]. Its stability. 76
13 potential application areas can be industrial production, automo- Though many metal oxides are suitable for detecting gases, the 77
14 tive industry, medical application, indoor air quality supervision most widely used semiconductor gas sensing material is SnO2 with 78
15 and environmental monitoring, etc. [2,3]. Thus, various materials n-type conductivity, high adsorption capacity, highly chemical and 79
16 such as inorganic semiconductors, conjugated polymers and thermal stability in air. Its band gap is 3.6 eV (300 K). The 80
17 carbon nanomaterials have been explored to fabricate gas sensors mechanisms responsible for SnO2 gas sensing are not completely 81
18 that can be used to detect combustible, flammable and toxic gases, understood and subject to ongoing discussion. An empirical model 82
19 and/or oxygen depletion [4,5]. Among them, inorganic semicon- to explain the fundamental gas sensing mechanism is as follows. 83
20 ducting metal oxides are currently being intensively investigated According to this model, chemisorption of oxygen from the gas 84
21 because of their low cost, flexibility in production, easy usage and phase creates extrinsic surface acceptor states that immobilize 85
22 various gases validation. Many metal oxides are suitable for conduction band electrons from the near surface region of the n- 86
23 detecting combustible, reducing, or oxidizing gases by the type semiconductor. Thus a depletion layer is created at the 87
24 conductive measurements [6]. A lot of papers about metal oxide interfaces of the sensor material (e.g., SnO2) as a result of adsorbed 88
25 gas sensors have been published in recent years. oxygen from air under ambient conditions. The presence of other 89
26 Due to the high demand for reliable and robust gas sensing gases with either reducing or oxidizing properties will conse- 90
27 devices, academic research continues to play an important role in quently affect the density of charge carriers in the near-surface 91
28 this field. Meanwhile, with the rapid development of nanoscience, region of each SnO2 grain [40]. Reducing gases will abstract 92
29 the superior control over the shape and the size of the nanocrystals surface-bound oxygen atoms, a process that release immobilized 93
30 enable sensors with an enhancement in sensitivity and selectivity some of the charge carriers (conduction electron). In contrast, 94
31 [7]. Small size, lightweight and high surface-to-volume ratios of oxidizing gases will immobilize further conduction-band electrons 95
32 these nanostructures are the best choice for improving the from the near-surface region by creating additional surface- 96
33 capability to detect gas species. Tin dioxide (SnO2) is one of the acceptor states. As a result, the foreign gases cause a decrease 97
34 most commonly used gas sensing materials and it is n-type (reducing gases) or increase (oxidizing gases) of the depletion layer 98
35 semiconductive material whose electrical conductivity is depen- thickness by tuning the surface-state density of SnO2. Then, the 99
36 dent on the density of pre-adsorbed oxygen ions on the surface. Its charge transfer between the surface adsorbed gaseous species and 100
37 excellent electrical and optical properties and well chemical SnO2 surface atoms leads to change in carrier concentration of 101
38 stability make SnO2 as a suitable material for gas sensors. The basic SnO2 and widening or narrowing of the depletion layer [41]. 102
39 detection principle of its sensor is still the change of the resistance Reducing gases in general increase the conductivity of SnO2 103
40 of the sensing layer with gas adsorption, where target gases have material, while the opposite is observed for oxidizing gases. A 104
41 red-ox reactions with the surface of SnO2. The resistance of SnO2 simple schematic image for this process is depicted in Fig. 1. Some 105
42 changes with the variation of target gas concentration [2]. Thus, more comprehensive discussion of the underlying principles can 106
43 nanostructured SnO2 materials can offer great potential for be found elsewhere [2,40,41]. 107
44 environmental applications including gas sensing [8–10]. To greatly improve the sensitivity of SnO2 materials, it is 108
45 By now, many kinds of SnO2-based materials with good sensing important to develop active materials with large surface areas and 109
46 characteristics have been obtained, including nanofibers [11–14], huge porosity, ensuring the easy reaction between SnO2 and 110
47 nanowires [15–19], nanotubes [20,21], hollow spheres [22,23], analyte gases. Thus, it is generally accepted that a hierarchical SnO2 111
48 hierarchical structures [9,24–27], nanoslab [28,29], etc. However, nanostructure with open porous features has a better performance 112
49 in addition to precisely controlling the size and morphologies of than the dense nanoparticles aggregates. So, three-dimensional 113
50 SnO2 materials [30], there are also other efficient strategies to (3D) SnO2 nanomaterials have received much attention recently. In 114
51 improve their sensing properties such as elemental doping, addition to this, SnO2 composites also can have a large surface area 115
52 composites assembly, noble metal function. There have been and provide effective diffusion of target gases, as well as additional 116
53 many published review papers related on SnO2 and its gas sensors possibilities for further modification such as guest metal loadings. 117
54 [9,10,31–40], but few of them paid special attention to its
55 composite. This is the main motivation of this review paper. SnO2 composite gas sensing materials 118
56 As a simple review of SnO2-based gas sensor materials, the main
57 attention in this work will be focused on various composites SnO2/inorganic metal oxides 119
58 containing SnO2 which have been published in the recent five
59 years. The discussion and contents involve the composites of Heterostructured SnO2 with other semiconducting metal 120
60 inorganic oxides/SnO2, carbon materials/SnO2, polymer/SnO2, oxides will provide another promising strategy to develop novel 121
61 noble metals/SnO2 and their potential uses as gas sensing high-performance gas sensors. As mentioned above, gas sensors of 122
62 materials. By focusing on SnO2 composites (excluding doped individual SnO2 materials are attractive due to their high 123
63 SnO2), we hope to provide readers better understanding on the sensitivity, quick response, and good stability but still have the 124
64 recent development in gas sensing, uncovering their possible main drawbacks of large band gap (3.6 eV, which can only respond Q3 125
65 research in future. to UV illumination) and high electron–hole recombination rate 126

[10]. However, these can be effectively modulated or overcome by 127


66 Gas sensing mechanism of SnO2 compositing SnO2 with other metal oxides such as n-type TiO2, 128

ZnO, WO3, In2O3, CaO, MgO, V2O5, Nb2O5, etc., and p-type NiO, 129
67 Materials that can change their properties depending on the Co3O4, Sb2O3, La2O3, Cu2O, Ag2O, CeO2, etc. (see Table 1 for more 130
68 ambient gas can be used as gas sensing materials. The most details). 131
69 common used sensing materials are semiconductive metal oxides, The heterojunctions formed between SnO2 and these metal 132
70 which can provide sensors with several advantages such as low oxides exhibit interesting electronic properties by coupling the 133

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx 3

Fig. 1. A schematic picture showing the chemical reactions at the surface of n-type gas sensing materials. Left: chemisorption of oxygen from air leading to the immobilization
of conduction electrons in the near surface. Right: reducing gases (e.g., CO) abstract surface-bound oxygen and thus release electrons back into the crystal. (Reprinted from Ref.
[40], copyright with permission from Wiley).

Table 1
List of various SnO2-metal oxide based gas-sensing heterostructures and composites according to their structural dimensions. Q26

Structure Materials Target gas Comments


0D Heterojunction SnO2–SnO NO2 [45]; p–n junction, with high sensitivity and low working
of nanoparticles H2 [46] temperature
SnO2–TiO2 [47,48] H2 Co-precipitated SnO2 and TiO2 [47]
SnO2–In2O3 [49] Cl2 The In/Sn = 12:1 (molar ratio) sample shows best sensing
performance
SnO2–CaO [50]; CuO/SnO2–In2O3 [51]; CO In reducing atmosphere; good selectivity to CO compared to
V–Sn–O [52] VOC and hydrocarbon
SnO2–ZnO [53,54]; SnO2–La2O3 [55]; VOC (ethanol, SnO2 content-dependent response; La2O3@Sb-doped SnO2;
SnO2–TiO2 [56]; SnO2–Sb2O3 [57] acetone, butane, Ag-doped TiO2/SnO2
etc.)
Core–shell/ g-Fe2O3–SnO2 [58]; SnO2–WO3 [59,60]; VOC, humidity Temperature-dependent abnormal p–n transitions [59];
nanosphere ZnO–SnO2 [61]; ZnO-doped SnO2 hollow nanosphere

1D Decorated SnO2–ZnO NO2 [62,63]; Branched SnO2 nanowires loaded with ZnO nanoparticles
nanowires/ [63]
nanorods/ SnO2–CuO H2S [64]; VOC [65] Branched SnO2 nanowires, facilitated ion transfer [64];
nanobelts thermally-oxidized Cu-Sn metallic nanowires [65]
SnO2–In2O3 [66,67] TMA La0.7Sr0.3FeO3 coated on electro-spun In2O3-SnO2
nanofibers
SnO2–Nb2O5 [68] Humidity Low-temperature (60  C) sensing
SnO2–MoO3 [69]; SnO2–TiO2 [70] Ethanol Working temperature lower to 120  C [69];
Composite SnO2–ZnO [71] H2 Numerous heterojunctions resulting dramatically
nanofibers enhanced response
Core–shell SnO2–ZnO [72,73] NO2, CO, C6H6, C7H8 UV-enhanced sensing at RT [72]; dual functional sensing
mechanism [73]
SnO2–Cu2O [74] p–n junction formed
Nanotube SnO2–Ag2O [75]; SnO2–Co3O4 [76]; VOCs Ultra-fast response and ultra-fast recovery [75]; High
composites a-Fe2O3–SnO2 [77]; SnO2–Fe2O3 [78] sensitivity and response time [78]

2D Layered SnO2–CuO [79]; SnO2–Cu2O [80] H2S p–n junction [80]


heterojunctions
SnO2–(WO3,TeO2, Al2O3, NiO, CuO, NO2 [81]; SO2 [82] SnO2–WO3 exhibits the best response to NO2; NiO/
In2O3, ZnO, TiO2, Ag2O, PdO, MgO, V2O5) SnO2sensor shows best response to SO2
Composite film SnO2–CuO H2S [83]; LPG [84] Sol–gel structure
SnO2–V2O5–MgO [85] SO2 5 wt% MgO and 2 wt% V2O5
Sn0.9Ti0.1O [86]; SnO2–La2O3 [87]; Ethanol/ UV-dependent [86]
ZnSnO [88] formaldehyde
SnO2–In2O3 [89]; SnO2–CeO2 [90,91] H2, CO Composition-dependent

3D Hierarchical a-Fe2O3–SnO2 [92]; Ethanol Fe2O3 nanosheets decorated SnO2 hollow nanosphere
heterostructures/ SnO2–NiO [93] Humidity Fast response depends on NiO content
hollow spheres SnO2/a–Fe2O3 [94]; CO, H2S, VOCs Double-shell composite, better for VOCs
SnO2–WO3 [95]; VOCs WO3nanoplate on SnO2 nanoparticle

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

4 J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx

134 advantages of different metal oxides and influencing the efficiency resistance. When exposed in an environment containing NO2, 199
135 of charge transportation [42–44]. On the other hand, by combining NO2 molecules trap surface electrons and result in an increase of 200
136 SnO2 material with various metal oxides, it is possible to develop a resistance. Thus the heterostructured gas sensors present much 201
137 large number of interesting multifunctional heterostructured better gas response. Another simple heterojunctions of SnO2 and 202
138 geometries (e.g., zero-dimensional (0D) core–shell nanoparticles, CaO nanoparticles are demonstrated in Fig. 2(C–D) [50]. During the 203
139 one-dimensional (1D) nanowire heterostructures, two-dimen- growth, calcium primarily was segregated at the SnO2 grain 204
140 sional (2D) layered heterojunctions, and 3D hierarchical hetero- boundaries as CaO and thereby restricted the growth of SnO2 205
141 structures). The resulting multi-component heterostructures have particles. As a result, uniform heterojunctions of SnO2 and CaO 206
142 unique electron transfer or molecule absorption/desorption nanoparticles were formed. The corresponding gas sensor showed 207
143 properties according to their specific geometry and thus are of a detection limit of 1 ppm to CO in air and the heterostructures 208
144 promise to exhibit improved sensitivity, reliable stability, and were stable after operating for 1.5 year. 209
145 controllable selectivity. In addition, it is worth noting that these Core–shell or nanosphere structures are regarded as novel 210
146 metal oxides are normally earth-abundant, easy for scalable interesting 0D heterostructural geometries. Normally 0D struc- 211
147 production, nontoxic and environmental-friendly, making the tures are referred to materials with diameter smaller than 100 nm, 212
148 heterostructured sensors of great promise for large-scale com- however here we will also include some heterostructures in 213
149 mercialization. However, to the best of our knowledge, there has hundreds of nanometer scale to distinguish them with the 214
150 Q4 been no article that systematically summarizes recent critical hierarchical materials below. These core–shell or nanosphere 215
151 progress of such SnO2-based metal oxide sensors. In this section, heterostructures usually have a consistent morphology and 216
152 we conducted a state-of-the-art review on recent progress of novel uniform size, and thus can exhibit constant sensitivity and 217
153 SnO2-based metal oxide heterostructures for gas sensors. As an reproducibility. Most importantly, such spherical heterostructures 218
154 index, various SnO2-based complex metal oxide composites were are very easy to be produced with a hollow structure, which 219
155 first summarized in Table 1. significantly increases the molecular absorption and desorption 220

efficiency and thus leads to increased sensitivity. Zhang et al. [58] 221
156 Zero-dimensional heterojunctions or spherical heterostructures studied the synthesis of core–shell g-Fe2O3–SnO2 hollow nano- 222
157 Zero-dimensional nanostructures normally have aspect ratio particles. As shown in Fig. 2(E–F), a porous SnO2 shell with 223
158 equivalent to 1 in all the directions. In past few years, well- thickness of 22 nm can be observed on the surface g-Fe2O3 spheres. 224
159 developed material synthesis techniques allow us to obtain SnO2 High-resolution TEM study indicated that the SnO2 shell was 225
160 metal oxide with homogeneous nanometric size. Taking advantage actually composed of various interconnected tiny SnO2 nano- 226
161 of their high surface area, these nanomaterials displays high crystals, which resulted in a porous structure with large surface 227
162 electric response when exposed to reducing gases such as CO, NO, area. Gas sensing test indicated that such core–shell heterostruc- 228
163 H2, and various volatile organic compounds (VOCs) [9]. Moreover, tures exhibited very fast response and recovery rate for organic 229
164 when combining SnO2 nanoparticles with other 0D metal oxides, gases. Fig. 2(G–H) shows another representative example of SnO2– 230
165 the obtained nanoclusters or heterostructures normally have a WO3 hybrid hollow spheres [59,60]. These nanospheres are overall 231
166 porous structure, which significantly improves the gas diffusion uniform, with a size of 550 nm and shell thickness of 30 nm. The 232
167 throughout the layers and thus gives the composite materials corresponding gas sensors were found to display a temperature- 233
168 better sensing performance. Most importantly, the combination dependent abnormal p–n transition for various gas molecules at 234
169 of different metal oxides leads to an effective band reconstruc- different temperature. As shown in Fig. 2I, the sensors have no 235
170 tion in the heterojunctions, which changes the charge transfer response to CO, H2, and NO; p-type response to NH3 and ethanol; 236
171 path and significantly modulates the resistance of sensing and n-type response to acetone at 95  C [59]. The same SnO2–WO3 237
172 materials [45]. This further allows us to design plenty of nanospheres were also used for humidity sensing (Fig. 2J–K) [60]. 238
173 interesting gas-sensing heterostructures. Particularly, for in- One can observe that such WO3–SnO2 sensor shows a higher 239
174 stance, heterostructure of SnO2 (n-type) and other p-type metal sensitivity than the other two kinds of humidity sensors (Fig. 2J). 240
175 oxides forms a constant p–n junction at the interface. In these However, it was also worth noting that the response time increased 241
176 p–n junctions, the conduction and valence bands usually bend to 289 s, and the recovery time was 22 s (Fig. 2K). The improved 242
177 and the Fermi levels are equalized associated with the formation sensitivity and decreased response could be attributed to the 243
178 of depletion layers, which directly results in the improvement of increased active sites provided by high surface areas of the hollow 244
179 the conductivity and the acceleration of response and recovery spheres. 245
180 times [45].
181 In Table 1, we summarized various representative examples of One-dimensional heterostructures 246
182 0D SnO2-based metal oxide heterostructures that developed in One-dimensional metal oxide nanowires have been widely 247
183 recent five years. One can observe that the direct contact of metal- used for gas sensors due to their fast charge transfer along single 248
184 oxide nanoparticles (heterojunctions) provides a simple and direction of the crystallized nanostructures. Besides, due to their 249
185 widely-acceptable methodology for composites. These hetero- high aspect ratio, metal oxide nanowires and their heterostruc- 250
186 structured gas sensors normally have increased structural porosity, tures can effectively avoid surface aggregation during materials 251
187 improved sensitivity and good reproducibility, and thus can be preparation and device operation (which is one of the main 252
188 used for the detection of NO2, H2, CO, various VOCs and so on. shortcuts of above 0D composites), and thus are very favorable for 253
189 Fig. 2(A–B) shows a representative SnO-supported SnO2 hybrid the absorption and desorption of various gas molecules, especially, 254
190 with p–n heterojunctions [45]. The gas sensing tests showed that in the long-term recycling use of the gas sensors [96]. A large 255
191 detection of NO2 with a limit of 0.1 ppm could even be realized at a number of SnO2 based 1D heterostructures have been developed in 256
192 low temperature (50  C). The main mechanism for sensing recent years. These include branched nanowires, nanoparticle- 257
193 enhancement is attributed to the p–n heterojunction structure, decorated nanowires/nanobelts/nanotubes, composite nanofibers, 258
194 which forms a depletion region at the interface. The bending of and core–shell nanowire heterostructures (see Table 1 for details). 259
195 forbidden bands occurs at the interface zone and makes the EF These materials show unique sensitivity and selectivity for 260
196 (Fermi level) equilibrium between the band gaps of SnO2 and SnO. different gases based on their heterostructure, and thus allowing 261
197 This actually facilitates the charge transportation at the hetero- us to develop various specific electronic sensors according to the 262
198 junction interface and gives the material a small base line target gases and application environment. 263

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx 5

Fig. 2. (A,B) p–n junction SnO2–SnO heterostructures. (Reprinted from Ref. [45], copyright with permission from RSC Publications) (C,D) heterojuctions formed between SnO2
and CaO nanoparticles. (Reprinted from Ref. [50], copyright with permission from Elsevier) (E,F) core–shell g-Fe2O3–SnO2 hollow nanoparticles. (Reprinted from Ref. [58],
copyright with permission from the PCCP Owner Societies) (G,H) SnO2–WO3 hybrid hollow spheres. (I) Response curves of the SnO2–WO3 sensors to different gases at the
same concentration (5000 ppm) at 95  C. (Reprinted with permission from Ref. [59], copyright (2015) American Chemical Society) (J) response and recovery of the three types
of humidity sensors for humidity values ranging from air (35% RH) to 98% RH. (K) response and recovery of a typical humidity sensor based on WO3–SnO2 hybrid hollow
spheres. (Reprinted from Ref. [60], copyright with permission from RSC Publications).

264 Fig. 3(A–B) shows two representative examples of the branched In normal SnO2 nanowires, the H2S sensing is accomplished by the 290
265 nanowire heterostructures [62,64]. As we commented above, the reaction of H2S with chemisorbed oxygen (forms SO2) or with SnO2 291
266 Q5 basic mechanism of gas sensors are based on the adsorption and (forms SnS2), which decreases the overall resistivity. With regard 292
267 desorption of target gas molecules, and thus the surface-to-volume to the CuO-functionalized branched nanowires, the increased 293
268 ratio of materials is of great importance. Compared with the surface area enables more absorption of H2S, which further 294
269 straight nanowire materials, the branched nanowires can provide a decreases the resistivity. Besides, it is possible that the attached 295
270 significantly increased surface-to-volume ratio. Meanwhile, it is CuO will also react with H2S to form CuS, which has a much high 296
271 also possible to combine different characteristics of core nano- conductivity. At the same time, CuO can lead to the evolution of H2 297
272 wires and branches. As a result, the branched nanowires are from the dissociated H2S on SnO2 nanowires, which further 298
273 promising to achieve excellent sensing performance with high increases the gas response rate [64]. 299
274 sensitivity, quick response and recovery time. In Fig. 3A, Kim et al. Another well-developed 1D heterostructure geometry is guest 300
275 [62] developed a heterostructure composed of ZnO-branched SnO2 nanoparticles decorated on SnO2 nanowires. Such heterostructure 301
276 nanowires for the detection of NO2. The observed sensing design combines the features of nanowire sensors and hetero- 302
277 performance improvement was attributed to (1) the generation junction sensors mentioned above. As shown in Fig. 3D, Fiz et al. 303
278 of structural defects including oxygen vacancies during branch [68] studied the gas-sensing performance of Nb2O5 nanoparticles 304
279 growth, (2) the increase of surface area for absorption and decorated SnO2 nanowires. This heterostructure was designed to 305
280 desorption of NO2 molecules, and (3) the resistance modulation overcome the shortcut of undesired cross-sensitivity effect of SnO2 306
281 due to the formation of SnO2/ZnO heterointerfaces. Fig. 3B further sensors at high temperature. As expected, the highest sensitivity 307
282 demonstrates a CuO-functionalized branched SnO2 nanowire was found at a low temperature (60  C), where it was driven by 308
283 heterostructure [64]. During the growth, a copper shell was first ionic conduction mechanisms. The formation of Nb2O5 coatings on 309
284 coated on SnO2 nanowires and followed by a high-temperature the SnO2 nanowires resulted in a higher density of defects, such as 310
285 Q6 annealing. The generated CuO nanoparticles on the nanowire longitudinal twins and stacking faults, formed at the interface due 311
286 backbones and they not only played a catalytic role in growing to the lattice mismatch. These oxygen-related defects allowed the 312
287 branches, but also enhanced the sensing properties with regard to tuning of the resistance and resulted in improved sensing 313
288 H2S gas. Schematic for the charge and mass transportation in SnO2 characteristics. Moreover, Fig. 3E shows the band alignment 314
289 nanowires and the branched heterostructures is shown in Fig. 3C. between the SnO2 and Nb2O5 heterojunctions, which indicates that 315

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

6 J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx

Fig. 3. (A) ZnO-branched SnO2 nanowire heterostructures. (Reprinted from Ref. [62], copyright with permission from Elsevier) (B) CuO-functionalized, branched SnO2
nanowire heterostructures and (C) charge and mass transportation in SnO2 nanowires and the branched heterostructures [64]. (D) Nb2O5 nanoparticles decorated SnO2
nanowires and (E) band alignment between the SnO2 and Nb2O5 heterojunctions. (Reprinted with permission from Ref. [68], copyright (2013) American Chemical Society) (F,
G) SnO2 nanoplates decorated TiO2 nanobelts. (Reprinted with permission from Ref. [70], copyright (2015) American Chemical Society) (H,I) 1D nanofiber heterostructures
composed of nanosized SnO2 and ZnO grains. (Reprinted with permission from Ref. [71], copyright (2015) American Chemical Society) (J) SnO2–ZnO core–shell
heterostructures. (Reprinted with permission from Ref. [73], copyright (2014) American Chemical Society) (K,L) Ag2O-doped SnO2 nanotube heterostructures. (Reprinted
from Ref. [75], copyright with permission from Wiley) (M) a-Fe3O4–SnO2 core–shell nanotubes. (Reprinted from Ref. [77], copyright with permission from Elsevier).

316 the steps in both valence and conduction bands align in the same the large sensing surface of SnO2 nanoplates with high energy 328
317 direction according to the Type II or staggered lineup [68]. The facets, and (3) the rapid electron transport in TiO2 nanobelts. 329
318 efficient electron transfer through the SnO2 core and intrinsic Another type of heterostructural geometry is in regard to the Q7 330
319 presence of oxygen vacancies must be remarked as a second composite nanofibers of different metal oxides. As shown in Fig. 3 331
320 intrinsic advantage of using such SnO2/Nb2O5 heterostructures. (H–I), Katoch et al. [71] developed a 1D nanofiber heterostructures 332
321 With respect to the nanobelt heterostructures, Chen et al. [70] composed of nanosized SnO2 and ZnO grains. This heterostructure 333
322 reported SnO2 nanoplates decorated TiO2 nanobelts (Fig. 3(F–G)) design employed the interesting hydrogen-induced metallization 334
323 as ethanol sensor. The obtained SnO2 nanoplates were demon- phenomenon in ZnO nanofibers and thus showed remarkable 335
324 strated with plenty of high energy facets, making the SnO2/TiO2 sensitivity and selectivity for the detection of H2 gas. These authors 336
325 nanobelt heterostructures very sensitive and selective to ethanol stated that (1) the adsorbed hydrogen generated a metallic thin 337
326 even at low temperatures (43  C). The improved sensing perfor- layer on the boundaries of the ZnO nanograins, drastically 338
327 mance was attributed to (1) the presence of heterojunctions, (2) enhancing the sensitivity and selectivity to H2 gas due to the 339

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx 7

340 SnO2–ZnO heterointerfaces, and (2) the SnO2 nanograins contrib- was for the first time used in the detection of H2S gas. Compared 405
341 uted to the sensing behavior through the formation of SnO2–SnO2 with the existing room temperature H2S sensors, these sensors 406
342 homointerfaces and SnO2–ZnO heterointerfaces. based on the tunneling modulation mechanism exhibited much 407
343 Heterostructure of core–shell nanowires and nanotubes have higher sensitivity, selectivity, and response/recovery rate to low 408
344 also been well-studied. Besides the fast single-direction charge concentration (sub-ppm) H2S at room temperature. The sensing 409
345 transfer property, these heterostructures are very easy to be performance can be further enhanced by laser illumination. 410
346 fabricated in a hollow structure, which is very favorable to gas Therefore, such p–n junction multilayer heterostructure may 411
347 sensing due to their large surface area. Fig. 3J shows a typical core– introduce a new direction of materials exploration for room 412
348 shell nanowire heterostructure of SnO2 and ZnO [74]. In such temperature gas sensors. Fig. 4C shows another gas sensor design 413
349 structures, heterojunctions were created at the interface between by patterning various metal oxide clusters on SnO2 film. Haridas 414
350 the core and shell. The barrier at the heterojunction interface was and Gupta [81] studied the patterned deposition of various metal Q8 415
351 the bottleneck for the current transporting through the system, oxide nano-thin micro-clusters including WO3, TeO2, Al2O3, NiO, 416
352 and the current was mainly dominated by the height and width of CuO, In2O3, ZnO, TiO2, Ag2O and PdO on RF-sputtered SnO2 film and 417
353 the barrier. Choi et al. [73] demonstrated that the sensitivity and tested their gas-sensing performance toward 10 ppm NO2 gas. 418
354 selectivity of SnO2–ZnO core–shell heterostructures had a domi- These authors found that integration of WO3 nano-thin micro- 419
355 nant dependence on the thickness of ZnO shell. A dual functional clusters with SnO2 thin film exhibits enhanced sensitivity, fast 420
356 sensing mechanism based on the contribution of the radial response/recovery time, and good selectivity at a lower operating 421
357 modulation effect of the electron-depleted shell and the electric temperature (100  C). However, they also commented that the 422
358 field smearing effect was proposed to explain the observed sensing formation of p–n junctions at the interface of SnO2 thin film and 423
359 behavior. Park et al. [72] also demonstrated that the sensing nano-thin micro-clusters of CuO, NiO, In2O3, Ag2O, PdO and TeO2 424
360 performance of such SnO2–ZnO core–shell heterostructures could resulted in much higher resistance and thus low sensor response. 425
361 be further enhanced by introducing UV light. Following the study Another vertically layered heterojunction film of SnO2 and CuO 426
362 of core–shell heterostructures, the development of nanotube was demonstrated in Fig. 4E [79]. Fig. 4F further shows an AFM 427
363 heterostructures may hold more potential for gas sensing since it is image of the film surface. The authors found that such SnO2–CuO 428
364 possible to accumulate many structural and electronic merits (e.g., multilayered film with 3 vol% CuO content could exhibit high 429
365 1D charge transport, heterojunctions, high surface-to-volume response and fast response time. The enhanced sensing perfor- 430
366 ratio, etc.) in one material geometry. For instance, Chen et al. [75] mance was attributed to the porous microstructure and the ability 431
367 developed a heterostructure of Ag2O-doped SnO2 nanotubes (Fig. 3 to spill the dissociated H2S molecules of the CuO top layers. 432
368 (K–L)) and found that such combination of SnO2 tubular structure Composite films composed of interconnected particles or sol– 433
369 and catalytic Ag2O dopants possessed interesting sensing behav- gel of SnO2 and other metal oxides have also been widely employed 434
370 iors. The dynamic transients of the sensors demonstrated both for gas sensors. The heterojunctions are uniformly formed 435
371 their ultrafast response and recovery towards ethanol and between adjacent grains or phases of different metal oxides and 436
372 butanone. The ultrafast response was attributed to the rapid thus the composite films exhibit highly improved sensing 437
373 oxidation of gas molecules with O or O2, whereas the remarkable performance due to the modulated charge transfer mechanism 438
374 recovery speed was attributed to the improvement of surface described above. Fig. 4(G–H) shows a gas-sensing device made of 439
375 reactions involving adsorption, dissociation, and the ionization of SnO2–CuO sol–gel that deposited on porous SiO2/Si substrate [83]. 440
376 oxygen by the presence of Ag2O [75]. Fig. 3M demonstrates another The device fabrication was based on a micromachining technology, 441
377 example of a-Fe3O4–SnO2 core–shell nanotubes with hollow which selectively deposited SnO2–CuO sol–gel on the pre-defined 442
378 structure and thus high surface-to-volume ratio [77]. The resulting device active regions. The device was expected compatible with 443
379 gas sensors exhibited a high response, short response–recovery current MEMS technology. The authors found that the device was 444
380 time, and good reproducibility to acetone, attributed to the 1D reliable for detecting H2S gas with a concentration above 2 ppm at 445
381 porous core–shell tubular structure as well as the presence of operating temperature of 200–250  C. It was also worth noting that 446
382 heterojunctions between SnO2 nanoparticles and a-Fe2O3 the device had very good humidity tolerance (up to 80% in air). 447
383 nanotubes. Another type of composite film was regarded to the interstitial 448

compounds formed from SnO2 and other metal oxides. Fig. 4I 449
384 Two-dimensional heterostructured composite shows SEM images of the Sn0.9Ti0.1O2 film [86]. The film is 450
385 2D layers or films of metal oxides are one of the earliest studied consisted of mesoporous nanocrystals of Sn0.9Ti0.1O2 with an 451
386 materials for gas sensors due to their easy production, high average size of 10 nm. Such composite film exhibited abnormal 452
387 reproducibility, and wide feasibility. However, their research photo response. Under UV illumination the resulting gas sensor 453
388 enthusiasm was slightly decreased in recent years. This is probably showed a response of 8.2–1000 ppm methanol. Finally it was 454
389 because of their moderate sensitivity and selectivity as compared noteworthy to mention that such metal oxide film based gas 455
390 with these nanostructured heterojunctions. Nevertheless, 2D sensors had been studied for a long time and thousands of 456
391 layered or composite films might be the most promising material interesting literatures could be found. Here we only presented 457
392 geometry to realize the commercial and industrial application. In several critical examples for each type of film sensors. To sum up, 458
393 this section, we summarized recent research progress of SnO2- due to their low cost, simple production, high response (compared 459
394 based 2D heterojunction or composite films for gas sensors. As with single material), reliable stability, good reproducibility, and 460
395 shown in Table 1, these 2D films can be divided into two categories wide feasibility, these composite films can provide an important 461
396 in general. The first one includes layered or aligned heterojunctions approach for the development of commercially applicable gas 462
397 between SnO2 film and another metal oxide film while the other sensors. 463
398 one indicates the co-deposited composite film or interstitial
399 compounds of SnO2 and other metal oxides. Three-dimensional hierarchical heterostructures 464
400 Fig. 4(A–B) demonstrates the design and fabrication of a quasi- 3D hierarchical heterostructures normally are higher-dimen- 465
401 2D Cu2O/SnO2 p–n horizontal multilayer heterostructure [80]. sional nanomaterials that periodically assembled from low 466
402 These heterostructured films have strictly periodic arrangement in dimensional nano-building blocks. Recently a significant effort 467
403 hundreds of microns and thus can be considered as a resonant has been contributed to the exploration of such novel 3D 468
404 tunneling system. The concept of resonant tunneling modulation hierarchical metal oxide heterostructures for gas sensors. This is 469

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

8 J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx

Fig. 4. (A,B) Quasi-2D Cu2O/SnO2 p–n horizontal multilayer heterostructures. [80] (C,D) Metal oxide nano-thin micro-clusters patterned on SnO2 film. (Reprinted from Ref.
[81], copyright with permission from Elsevier) (E,F) SnO2–CuO multilayered gas sensor. (Reprinted with permission from Ref. [78], copyright (2013) American Chemical
Society) (G,H) CuO/SnO2 sol–gel on porous SiO2 for H2S sensor. (Reprinted from Ref. [83], copyright with permission from Elsevier) (I) mesoporous Sn0.9Ti0.1O2 film. (Reprinted
from Ref. [86], copyright with permission from Elsevier).

470 mainly due to their attractive intrinsic features including high (for the a-Fe2O3 nanosheets). The obtained nanocomposites 485
471 specific surface area, low density, abundant interior voids, good contain a hollow interior sphere and a double-shell structure 486
472 surface permeability, and high possibility to form a large number of (Fig. 5B). The inner shell was identified as a hollow SnO2 sphere, 487
473 internal heterojunctions. Moreover, the morphologies of metal which turned out to be black colored in the image, while the outer 488
474 oxides with hierarchical structures can be tuned to enhance the gas shell was composed of numerous a-Fe2O3 nanosheets. The 489
475 diffusion rate and enlarge the specific surface area, and thus thickness of the a-Fe2O3 shell was 100 nm. Such hierarchical 490
476 further improve the gas sensing performance. In this section we material shows obviously enhanced response in the detection of 491
477 summarized critical recent progress of SnO2-based hierarchical ethanol at 225  C. The authors also demonstrated that the gas 492
478 nanocomposites for gas sensors. sensor shows constant response during a continuous test of 493
479 Representative samples of hierarchical nanostructures of SnO2 20 days. Similar hierarchical structure of SnO2/NiO was prepared 494
480 and other metal oxides are shown in Fig. 5(A–H). Sun et al. [92] by Kim et al. (Fig. 5(C–D)) [93]. The authors demonstrated their 495
481 developed the a-Fe2O3 nanosheets hierarchically assembled SnO2 superior sensing performance of CO with a reduced humidity 496
482 hollow spheres (Fig. 5(A–B)). The material preparation was dependence. Very recently, Nayak et al. [95] demonstrated another 497
483 Q9 achieved by combining a hydrothermal route (for the hollow interesting flower-like hierarchical nanocomposite composed of 498
484 SnO2 nanosphere) and a microwave-assisted hydrolytic reaction SnO2 nanoparticles decorated WO3 nanoplates (Fig. 5E). Fig. 5(F–G) 499

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx 9

Fig. 5. (A,B) Hierarchical Assembly of a-Fe2O3 nanosheets on SnO2 hollow nanospheres. (Reprinted with permission from Ref. [92], copyright (2015) American Chemical
Society) (C,D) NiO-doped hierarchical SnO2 hollow heterostructures. (Reprinted from Ref. [93], copyright with permission from Wiley) (E) flower-like hierarchical
nanocomposite composed of SnO2 nanoparticles decorated WO3 nanoplates. (F–G) WO3 nanoplates loaded with different amount of SnO2 nanoparticles. (I) Band
configuration at the interface of the a-Fe2O3/SnO2 heterostructure in different atmospheres. (Reprinted from Ref. [95], copyright with permission from RSC Publications) (J)
schematic representation of (a) WO3–SnO2 mixed oxides with available barrier potentials (K) energy band structure of a WO3–SnO2 metal oxide heterostructure.

500 further shows several individual WO3 nanoflakes loaded with heterojunctions formation in above SnO2-a-Fe2O3 [92] and SnO2– 524
501 different amount of SnO2 nanoparticles. The authors found that WO3 [95] hierarchical heterostructures. First, combination of 525
502 these nanocomposites showed significantly improved sensing different metal oxide leads to band bending (see Fig. 5(I–K)), which 526
503 performance to VOCs compared with single SnO2 or WO3. establishes a potential barrier at the heterojunctions. The potential 527
504 Meanwhile, their response to different gases (ammonia, ethanol, barriers (Fig. 5J) hinder the electron transport through the 528
505 and acetone) was varied according to the amount of SnO2 loaded nanostructures, and thus provide extra electrons for more 529
506 on the flakes, which was interesting since it might allow us to oxygen species to adsorb on the surface of the sensing layer, 530
507 design sensor devices for a specific target gas by simply varying the which effectively increases the sensor response. Second, the 531
508 material composition. heterojunctions formed at the interface of different metal oxides 532
509 It is generally accepted that when a reducing gas-sensing metal possess high energy and thus can act as the active site to catalyze 533
510 oxide is exposed in air, electrons in conduction band are trapped by the decomposition of target molecules [95]. These decomposi- 534
511 oxygen to form adsorbed oxygen species (O, O2, O2 and O22), tion reactions release free electrons and decrease the sensing 535
512 which leads to the electron depletion on metal oxide surface and layer resistance, leading to further enhancement of sensor 536
513 thus contributes to a large base line resistance of material. On the response. 537
514 other hand, when the material is exposed in an environment
515 containing reducing gas (e.g., ammonia, ethanol, acetone, etc.), the SnO2/carbon materials 538
516 targeting gas molecules react with the absorbed oxygen species.
517 This results in release of electrons to the metal oxides and thereby Carbon nanomaterials have a unique place in nanoscience field 539
518 their resistance decreases. With respect to the hierarchical metal owing to their exceptional electrical, thermal, chemical and 540
519 oxide heterostructures, the observed performance enhancement is mechanical properties and have found many potential applications 541
520 not only because of their hollow structure for gas diffusion and in various areas [97–99]. Two important candidates, carbon 542
521 large surface area for molecule absorption/desorption, but also due nanotubes (CNTs) followed by graphene, have gained much 543
522 to the formation of numerous internal heterojunctions. Schematics attention and impacted many fields, and the number of potential 544
523 in Fig. 5(I–K) shows the band structure realignment as well as the applications continues to grow. 545

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

10 J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx

546 CNTs/SnO2-based gas sensing materials cases, SnO2 functionalized SWCNT network showed improved 612
547 CNTs having a unique structure and amazing 1D structural sensing performance than the un-functionalized SWCNTs. A higher 613
548 feature appear as a potential candidate for gas sensors. Kong et al. Sn2+ content and the greater active surface area might be the key 614
549 [100] found that single-walled carbon nanotubes (SWCNTs) factors that increased the sensing performance of the hybrid 615
550 exhibited high sensitivity to NO2 and NH3 at room temperature structure [116]. 616
551 in 2000 and it inspired researchers to extensively explore CNTs for Even surface modification of CNTs/SnO2 hybrid gas sensors with 617
552 gas sensors. The gas-sensing mechanism of CNTs is based on the noble metals (Pd, Ru, Pt, etc.) was reported to promote sensitivity Q10 618
553 resistance change due to the charge transfer between the adsorbed and improvement of the gas sensors selectivity, because these 619
554 gas molecules and CNTs. Oxidizing gases (e.g., NO2) withdraw metals or their oxides were very efficient catalysts for chemical 620
555 electrons from CNTs, whereas reducing gases (e.g., NH3) donate reactions taking place on the surface [105,115,118]. For H2 621
556 electrons, which can lead to significant changes in the electrical detection, Dhall and Jaggi [115] thought that hydrogen molecules Q11 622
557 conductivity of CNTs [101]. CNTs/SnO2-based gas sensors have could react with Pt nanoparticles and dissociated into H atoms on 623
558 received considerable attention because of their outstanding the surface of noble metal. Then Pt nanoparticles acted as charge 624
559 properties such as faster response, higher sensitivity, lower donor with subsequent transfer of electrons to SnO2 nanoparticles 625
560 operating temperature, and a wider variety of detectable gas that were coated on CNTs. These electrons were shuttled into the 626
561 [41,102–107]. CNT network due to the positioning of CNTs and SnO2 particles. 627
562 CNTs usually exhibit a p-type semiconductor character and Thus the generation of hole–electron recombination in CNTs would 628
563 excellent sensing properties. CNT based p–n heterojunctions also result in an increased resistance of the sensor [115]. 629
564 show some special properties, which to a great extent is resulted
565 from the synergistic reactions between the component species. Graphene/SnO2-based gas sensing materials 630
566 CNT–SnO2 heterojunctions can avoid the aggregation and improve In addition to SnO2/CNTs gas sensing materials, SnO2/graphene 631
567 the sensor stability [10]. Meanwhile, the low Schottky barrier nanohybrid based gas sensors have received quite a lot attention 632
568 between SnO2 and CNTs can facilitate the electron transport from the researchers recently [119–131]. Graphene is a 2D carbon 633
569 between them. The function of interfacial structures and the nanomaterial with a high aspect ratio, specific surface area and 634
570 mechanism of the improved gas-sensing properties of CNTs/SnO2 excellent electronic properties, which facilitate sensing gases at 635
571 composites were discussed by Jia et al. in detail [104]. They found room temperature. Nowadays, the research on 2D materials has 636
572 that the interaction between SnO2 and CNTs could be enhanced by attracted much more attention since the discovery of graphene 637
573 calcination, leading to easy electron transfer between them, [132–138]. Graphene and its derivatives like graphene oxide (GO), 638
574 improved crystallinity and increased pore diameter and pore reduced graphene oxide (rGO) etc. have been reported to show 639
575 volume, thus playing key roles in the improvement of the gas- sensing applications for gas sensors. The first report on graphene 640
576 sensing properties [108]. based gas sensor was in 2007 by Novoselov’s group [139], where a 641
577 Marichy et al. [103] applied atomic layer deposition method to small sensor made from graphene was able to detect individual gas 642
578 coat the inner and outer walls of CNTs with thin films of SnO2 with molecules by changing the local carrier concentration in graphene. 643
579 various thicknesses. The p–n heterojunction formed between the Yuan and Shi [4] summarized the recent advancements on the 644
580 n-type SnO2 and the p-type carbon support played an important synthesis of graphene materials and the techniques applied for 645
581 role in the sensing mechanism. The SnO2 on CNTs was considered fabricating gas sensors. More recently, Chatterjee et al. [119] 646
582 to be involved in the receptor function while the carbon support reviewed the state of the art of gas sensors based on graphene and 647
583 provided mainly the electronic conduction path. As depicted in metal oxide hybrid nanostructures for detection of various 648
584 Fig. 6(a–b), in the case of MOx@CNT-based sensor, a depletion common toxic gases, including SnO2, ZnO, WO3, Cu2O and 649
585 region exists at the heterojunction of the MOx shell and the tube, Co3O4. In this section, recently-developed SnO2/graphene compo- 650
586 and a junction is created between two coated CNTs at the attaching sites are briefly summarized. 651
587 point, generating an additional potential barrier. When meeting an Similarly to CNTs, the incorporation of n-type SnO2 nano- 652
588 oxidizing target gas, such as NO2, adsorbs on the surface of the particles on p-type graphene sheets can create a hybrid 653
589 metal oxide, a new depletion region is thus created at the surface of nanostructure and form p–n junction, which will improve the 654
590 the shell. Indeed, the adsorbed NO2 extracts electrons from the gas-sensing performances [140,141]. Lin et al. [142] reported the 655
591 metal oxide shell, modifying the width of the depletion layer at the hydrothermal fabrication of SnO2/graphene composite with a 3D 656
592 surface of the metal oxide. This will alter the depletion layer at the network structure and the gas sensors using the composite 657
593 n-MOx/p-carbon junction and therefore modifies the whole exhibited good sensing properties to NH3 from 10 to 50 ppm at 658
594 resistance of the heterostructured sensor, as schematically shown room temperature. A 3D mesoporous SnO2/rGO composite was 659
595 in Fig. 6(c–d). More detailed discussions can be found in Ref. [103]. prepared by hydrothermal and lyophilization method, as reported 660
596 The detectable gases using CNTs/SnO2 include alcohol vapor [41], by Li et al. [143]. The supported SnO2 nanocrystals with an average 661
597 sulfur hexafluoride [109], acetone [104], hydrogen [110], NO2 [111], size of 2–7 nm and the specific surface area of the 3D mesoporous 662
598 methane, CO and NO [107]. composite was higher than 440 m2/g. The composite exhibited a 663
599 At the same time, some papers on SWCNTs/SnO2 as gas sensing good linearity for NO2 detection, and the limit of detection was 664
600 materials were also reported [102,112–117]. Mao et al. [102] calculated to be as low as about 2 ppm at 55  C [143]. Xiao et al. 665
601 reported a schematic of the gas-sensing platform of SWCNTs/SnO2, [144] fabricated discoid SnO2 modified by rGO. The response of the 666
602 as shown in Fig. 6(e). In the sensor, the SWCNTs worked as the sensor made of the composite to 1 ppm NO2 at 75  C was nearly one 667
603 conducting channels. When direct adsorption H2 molecules onto order of magnitude higher than that of SnO2, and the detection 668
604 SnO2, they would induce electron transfer from H2 to the SWCNTs, limit was improved to 50 ppb. 669
605 thereby changing the sensor conductivity. Su et al. [116] Chen et al. had carried out a series of work on SnO2/graphene Q12 670
606 summarized the possible underlying mechanism of SnO2/SWCNTs gas sensing materials [145,146]. A gas-sensing platform with SnO2 671
607 hybrid nanostructures for CO and NH3 detection. The morphology, nanocrystals–rGO was studied and it showed excellent response to 672
608 shape and valence state of the SnO2 nanoparticles could be target gases with a low 1 ppm detection limit for NO2 at room 673
609 controlled by adjusting the deposition potential and charge temperature [145]. Later, they developed In-doped SnO2/rGO 674
610 density, and post annealing of the nanostructures to determine composite by one-pot aqueous method and it showed a much 675
611 its effect on NH3 gas sensor sensitivity as chemiresistors. In all their higher sensitivity than SnO2/rGO. The dopants induced a large 676

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx 11

Fig. 6. (a–b) shows the TEM images of SnO2@CNTs, CNTs were treated at (a) 700  C and (b) 3000  C. The junction existing between the tubes of the networking at their crossing
points, as shown by the dotted square in part (c), as well as the different depletion regions existing as a function of the presence of the target gas (d). (Reprinted with
permission from Ref. [103], copyright (2013) American Chemical Society) (e) schematic of the gas-sensing platform of SWCNTs decorated with SnO2 nanocrystals. (Reprinted
from Ref. [102], copyright with permission from RSC Publications).

677 number of oxygen vacancies in the SnO2 nanocrystals, leading to an typical p-type semiconductor for the detection of NO2. Their exact 689
678 Q13 increase in the number of oxygen ion species to react with NO2 sensing mechanism of them was not well explained. Ge et al. [149] 690
679 molecules [146]. Zhang et al. [147,148] also performed some good fabricated Ag/SnO2/graphene ternary nanocomposites by a wet- 691
680 work to further improve the sensing properties of SnO2/rGO chemical method and found that its lowest detection concentra- 692
681 composite by the incorporation of CNTs [147] and Ag nanoparticles tion to acetone was only 0.005 ppm. 693
682 [148] to form ternary hybrids. The sensor based on rGO/CNT/SnO2 Russo et al. [150] reported the hydrogen-sensing performance 694
683 hybrids exhibited a high response, fast response and recovery rate, of SnO2/rGO and Pt-SnO2/rGO nanostructures, and the composites 695
684 good selectivity as well as good stability for detection of NO2 at showed improved sensing performance relative to that of the 696
685 room temperature [147]. The introduction of Ag nanoparticles into corresponding pure and binary systems as a consequence of 697
686 the SnO2/rGO hybrids could significantly enhance the NO2 sensing synergistic effects between the different components in the 698
687 performances at room temperature, compared to SnO2/rGO composites. Fig. 7(a–c) shows the TEM image of SnO2, SnO2/rGO 699
688 hybrids [148]. However, both ternary hybrids exhibited as a and Pt-SnO2/rGO. Fig. 7(d) shows the comparison of the response 700

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

12 J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx

Fig. 7. TEM images of (a) SnO2, (b) SnO2/rGO, and (c) Pt-SnO2/rGO, (d) comparison of the response to 1% H2 in air for all sensors investigated. (Reprinted from Ref. [150],
copyright with permission from Wiley) (e) a model of a potential barrier to electronic conduction at heterojunction boundary and grain boundary. (Reprinted from Ref. [144],
copyright with permission from Elsevier).

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx 13

701 to 1% hydrogen in air. The Pt/SnO2/rGO sensor exhibited high concentration of CO, down to below ppm level (1–25 ppm) with 764
702 response to hydrogen at concentration between 0.5 to 3% in air at ultrafast response–recovery time (in s), high response, good 765
703 near room temperature (50  C), with response and recovery times stability, and reproducibility [160]. In order to reduce the power 766
704 of 3–7 and 2–6 s, respectively. The properties of the composites consumption of Pd/SnO2 based CO gas sensor, several parameters 767
705 were related to the heterojunction barrier formatted at SnO2–rGO such as Pd loading, organic binder content and sensor design 768
706 interface and the presence of Pt which acted as a catalytic promoter should be optimized [161–164]. For example, a low concentration 769
707 to further improve the sensitivity [150]. As shown in Fig. 7(e), two of Pd nanodots could enhance the response of sensor, but extensive 770
708 different potential barriers and depletion layers coexist in SnO2/ addition of Pd into the sensing materials would result in the 771
709 rGO heterostructures. One is on the interface between SnO2 degradation of sensing characteristics [163,165]. Meanwhile, the 772
710 nanograins and the other is at the interface between rGO and SnO2 surface area and porosity of SnO2 also affect the sensing properties, 773
711 [144]. In the air, oxygen molecules are adsorbed on the surface of thus a high gas sensing performance can be derived from the large 774
712 SnO2/rGO by capturing free electrons from the composite. When surface area, high activity and well dispersion of Pd additive, as 775
713 exposed to NO2, causing the extraction of electron to be extracted well as high porosity [166]. 776
714 from SnO2 and widening both depletion layers. Electrospun SnO2 nanofibers are an ideal substrate to support 777
715 All above previous reports demonstrated that SnO2/graphene Pd and the resultant composites usually exhibit excellent gas 778
716 composite is a very promising candidate material for gas sensors. sensing properties, because the resultant SnO2 fibers are porous 779
717 SnO2/graphene composite has the following advantages to and with a high surface area, as well as long continuous fibers up to 780
718 improve its sensing properties. (I) Such hybrid composite can several centimeters [167–169]. Yang et al. [170] reported that 781
719 reduce the operation temperature significantly, even as low as unloaded and Pd-loaded SnO2 nanofiber mats could be synthesized 782
720 room temperature range. (II) The interaction with graphene by electrospinning followed by hot pressing at 80  C and 783
721 support can effectively prevent the aggregation of SnO2 nano- calcination at 450 or 600  C, and that Pd-loading could effectively 784
722 particles and reduce its particle size. (III) The resistance of SnO2/ suppress the SnO2 grain growth during the calcination process. The 785
723 graphene can be reduced by several orders of magnitude, Pd-loaded SnO2 sensors had 4 orders of magnitude higher 786
724 compared with pure SnO2 material. resistivity and exhibited significantly enhanced sensitivity to H2 787

and lower sensitivity to NO2 compared to their unloaded 788


725 SnO2/noble metals sensing materials counterparts due to the enhanced electron depletion at the 789

surface of the PdO-decorated SnO2 crystallites and the catalytic 790


726 In above sections, additives including inorganic oxide materials effect of PdO [170]. Wang et al. [171] also used electrospun SnO2 791
727 and carbon species are used for increasing the gas response of SnO2 nanofibers to load different contents of Pd and to investigate their 792
728 to gases. In order to increase the sensor signals, to diminish the gas sensing properties. 793
729 cross sensitivity to water and to reduce the operation temperature In several reports, two kinds of additive are simultaneously 794
730 as well as the response–recovery times, small quantities of noble added into SnO2 in addition to Pd [57,172–175]. Jeong et al. [172] 795
731 metals (Pd, Pt, Au, Ru and Ag etc.) can be added to the most metal applied Pd and Sn co-deposited on SnO2 nanowires surface and the 796
732 oxides in a popular manner [151,152]. The gas response of SnO2 composite were shown to be a highly sensitive hydrogen sensor 797
733 sensors can be further enhanced by loading noble metals or metal with fast response time at room temperature. Compared with the 798
734 oxide catalysts to chemically sensitize the oxide semiconductor SnO2 nanowires deposited with Pd or Sn nanoparticles alone, the 799
735 due to their catalytic properties. Noble metal is usually high- Pd/Sn-deposited SnO2 nanowires exhibited a significant improve- 800
736 effective oxidation catalyst and its ability can be used to enhance ment in the sensitivity and reversibility of sensing hydrogen gas in 801
737 the reactions on gas sensor surfaces. A wide diversity of methods, the air at room temperature. Two factors should be responsible for 802
738 including impregnation, sol–gel, sputtering and thermal evapora- the enhanced performances. One was that Sn was oxidized only to 803
739 tion, has been used for depositing noble metal additives into oxide suboxides by Pd and the other was the near-perfect hydrophobicity 804
740 semiconductors [6]. In this section, we will put our focus on noble of its surface, as shown in Fig. 8. Grobmann et al. [173] found that 805
741 metals/SnO2 sensing materials with different metals. the addition of Sb to Pd/SnO2 resulted in the significant decrease in 806

baseline resistance; up to two orders of magnitude in dry air at 807


742 Pd/SnO2 sensing materials 300  C and three orders of magnitude in humid air at 300  C [173]. 808
743 Pd has been the most widely applied to incorporate with SnO2 SnO2 also showed a reduced resistance owing to the substitution of 809
744 grains in all the noble metals for gas sensing applications in the Sn4+ by Sb5+, because their ionic radii were matching [174]. 810
745 past five years [153–159]. Palladium is a suitable oxidation catalyst There have several reports in the literature on the gas sensing 811
746 for SnO2 based sensors for reducing gases typically. The detected enhancement by UV-light irradiation at room temperature for 812
747 target gases in the reported papers are mostly reducing gases by detecting both reducing and oxidizing gases [176,177], and UV- 813
748 using Pd/SnO2, such as CO, H2, and methane. Moreover, it can light irradiation is a simple route to improve performance of 814
749 balance the decreased response to the target gases if water is sensors. More recently, Saboor et al. [178] investigated the NO2 gas 815
750 presented in the surrounding atmosphere. Pd is considered an sensing performance of Pd-loaded SnO2 film sensors under UV- 816
751 example of the so-called Fermi-level control type and it or PdO light irradiation, using a UV-LED of 365 nm wavelength and light 817
752 clusters usually present at the surface to control the position of the intensities ranging from 0–137 mW/cm2 at room temperature. For 818
753 Femi level for SnO2. Palladium was reported to exhibit the most the sensors with poor gas response under no UV-light irradiation 819
754 active among various catalysts including Pt, Ag, Ni, Pd, Au, NiO showed a considerable improvement in the response under UV- 820
755 Au2O3 on SnO2 to detect methane at a relatively low operating light irradiation. A drastic decrease in the resistance in air of the 821
756 temperature at 220  C due to the dominant roles played by both sensor under UV-light irradiation could be a possible reason for the 822
757 Fermi level energy control mechanism and spillover mechanism enhanced response value. In addition to UV-light irradiation, the 823
758 [81]. Thus, the recent progress in Pd/SnO2 sensing materials will be incorporation of rGO with Pd/SnO2 to form a ternary composite 824
759 briefly discussed in the following section. could be also effective. The sensor based on Pd/SnO2/rGO ternary 825
760 There are many factors that influence the performance of gas nanocomposite film responded more strongly to low concen- 826
761 sensors. Gas-sensing characterization demonstrated an enhanced trations of NH3 gas at room temperature than SnO2 and SnO2/rGO, 827
762 CO-sensing performance for SnO2 nanowires decorated with Pd attributing to the higher conductivity and catalytic activity of the 828
763 nanoparticles and the SnO2 based sensors could detect a very low Pd/SnO2/rGO composite [179]. 829

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

14 J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx

Fig. 8. Images of a water droplet before and after being dropped on the surface of the Sn/Pd-deposited SnO2 nanowire (top), STEM elemental mapping image for Sn/Pd-
deposited SnO2 nanowire (bottom–left), and conductance change, DG/G0, measured at room temperature versus time plot for Pd/Sn-deposited SnO2 nanowire (bottom–
right). (Reprinted with permission from Ref. [172], copyright (2013) American Chemical Society).

830 In most cases, SnO2 is selected as a support to deposit catalytic parameters greatly influenced the growth behavior of Pt nano- 855
831 Pd nanoparticles on its surface, where Pd grains are thus generally particles in terms of size and formation density and that Pt 856
832 dispersed on the surface of SnO2 grains. While, Gyger et al. [180] functionalized SnO2 could detect reducing gases such as CO, 857
833 prepared core@shell nanocomposites with Pd0 encapsulated by toluene and benzene [190]. 858
834 SnO2 shells via water-in-oil microemulsion. Extraordinarily high The Pt loading will introduce a well-known spillover effect, 859
835 sensor signals could be observed when exposing the Pd@SnO2 which remarkably promotes the dissociation of oxygen molecules 860
836 nanocomposite to CO in humid air. whose atomic products then diffuse to the metal oxide support. As 861

shown in Fig. 9(b), more electrons are transferred to the adsorbed 862
837 Pt/SnO2 sensing materials oxygen from SnO2, resulting in the obvious decrease of carrier 863
838 One of the most often-used noble metals especially in concentration within SnO2 nanoparticles. When being exposed to 864
839 commercial sensors is platinum. Similarly to Pd, Pt is also a CO, Pt clusters catalyze the oxidation of CO, causing the electrons 865
840 superior catalyst and its role in the promotion of sensitivity of SnO2 trapped by the surface oxygen species to be released. Thus it will 866
841 is being studied now [181–187]. decrease the depletion layer and expand the conduction channel 867
842 Controlling the configurations of Pt particles has been deemed [185]. 868
843 very important for gas sensing. Lin et al. [188] applied atomic layer The operation conditions for gas sensors also influence its 869
844 deposition (ALD) method to control the morphology, size, and performance for Pt/SnO2 materials. Pt-loaded SnO2 nanofiber mats 870
845 concentration of Pt particles on SnO2 nanowires by varying the were prepared by electrospinning followed by calcination and 871
846 cycling numbers and found that the sensing properties could be subsequent mixing with organic binder and colloidal Pt nano- 872
847 altered by Pt catalyst and the modification of Schottky barrier particles, and their gas sensing properties were influenced by Pt 873
848 junctions on the nanowire surface in the vicinity of Pt nano- loading, operation temperature and gas concentration [190]. The 874
849 particles. Their assumption was likely justified by observing the Pt-loaded SnO2 sensors were of much high sensitivity to H2 gas 875
850 electrical properties of Pt-decorated nanowires obtained after than their pristine SnO2 counterpart, reaching a maximum 876
851 repeating deposition cycles a number of times in air at room response of 16.6 upon exposure to 2.5 ppm H2 in dry air at the 877
852 Q14 temperature, as shown in Fig. 9(a) [188]. Choi and Kim [189] also operating temperature of 300  C [190]. 878
853 found a method to control Pt nanoparticles on SnO2 nanowires by There have developed several methods to improve the sensing 879
854 applying g-ray radiolysis. They reported that the processing properties of Pt/SnO2 material. Similarly to Pd/SnO2, Haridas et al. 880

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx 15

Fig. 9. (a) Dependence of electrical conductance of Pt-decorated SnO2 nanowires on the ALD reaction cycle number measured in air. (Reprinted from Ref. [188], copyright with
permission from RSC Publications) (b) schematic diagram of the mechanism depicting the Pt-loading enhances the sensing performance of SnO2 nanoparticles sensors by
spillover effect. (Reprinted from Ref. [185], copyright with permission from Elsevier).

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

16 J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx

881 [191] reported that under UV illumination the room temperature nanosheets are due to less agglomerated and hollow architectures, 946
882 response of SnO2 film sensor loaded with Pt catalyst was about catalytic activity of Au, as well as enhanced electron depletion on 947
883 4400 for 200 ppm liquefied petroleum gas and the response SnO2 surface [203]. 948
884 increased with increase in thickness of Pt clusters till 10 nm, SnO2 nanomaterials being co-doped with inorganic metal oxide 949
885 attributing to the efficient catalytic dissociation of gases molecules and Au nanoparticles were received a lot of interest such as Cu, V 950
886 Q15 and the spillover process. Haridas et al. [191] found that the role of and Zn [204–206]. Functionalizing host SnO2 with guest metal 951
887 catalysis in gas sensing mechanism toward CO gas was important cations is effective to enhance sensor response because the guest 952
888 for SnO2 sensor with Pt catalyst and that if the rate of CO oxidation metals possess variable electronic structures or induce defect sites 953
889 was too low or too high, the sensitivity was low in spite of high Pt to promote chemical reactions via catalytic or electronic effects 954
890 dispersion. Thus, they could obtain samples with different catalytic [205]. Li et al. [206] employed electrospinning method to prepare 955
891 activity with an identical Pt loading by sintering at different porous SnO2 nanofibers containing both Cu2+ and Au. Au is an 956
892 temperatures from 400 to 800  C. The highest sensitivity was effective route to reduce the activation energy of the reaction 957
893 shown for the sensor obtained with Pt/SnO2 sample sintered at between gas molecules and ionosorbed oxygen species to reduce 958
894 800  C [192]. the work temperature of the sensors. When SnO2 was also doped 959
895 The role of Pt in the gas sensing process was further studied by with Cu2+, the grain size of SnO2 decreased because of the 960
896 Hubner et al. [193]. They applied X-ray absorption near edge mismatched atomic radius between them. The co-doping of Cu2+ 961
897 structure spectrum to determine the state of the Pt under the and Au caused a synergistic effect which considerably enhanced 962
898 working conditions of the 0.2 wt% Pt/SnO2 sensors before and after the sensitivity [206]. 963
899 reduction in 2 vol% H2 in He at 600  C [193]. Even under rather
900 strong reducing conditions, no metallic Pt could be observed. Pt Other metals and their mixture with SnO2 964
901 seemed to either influence the electronic structure as a whole and/ In addition to above three noble metals, Ru or its oxide were 965
902 or created new adsorption sites in the SnO2 lattice that lead to also applied to modify SnO2 nanocrystals in order to improve its 966
903 enhanced selectivity and sensitivity. The role of Pt in gas sensing is sensing properties [205–207]. In comparative sensor tests to NH3 967
904 thus very complex from above results. using nanocrystalline SnO2 doped or modified by various additives, 968
905 In addition to above experimental researches, the oxidation it was RuOx modifier that exhibited the highest sensibilization 969
906 process of CO on the Pt doped SnO2 was also studied by Li using effect because RuOx-promoted materials could enhance affinity 970
907 first-principles method based on the density functional theory and reactivity to NH3 gas [207]. Ru/SnO2 sensing materials also 971
908 [194]. His results proved that the presence of Pt could induce new showed high responses to trimethylamine, ethanol, acetone etc. 972
909 electronic states near the Fermi energy and facilitated the [208,209]. However, only a few papers were reported to concern 973
910 formation of oxygen vacancy. They subsequently concluded the Ag/SnO2 composites for gas sensing application [210] and the 974
911 sensing mechanism of Pt/SnO2 toward CO gas. morphological configuration of Ag grains on SnO2 strongly 975

governed the sensing characteristics, i.e., Ag nano-islands could 976


912 Au/SnO2 sensing materials enhance gas response, but continuous Ag layers deteriorated it 977
913 After their study on Pt and Pd roles in gas sensing, Hubner et al. [211]. 978
914 [195] applied high energy resolved fluorescence detected X-ray There are also several papers reported on using Pd–Au binary 979
915 absorption spectroscopy and simultaneous DC resistance, and metals to decorate SnO2 [212,213]. Wen et al. [212] applied 980
916 work function change measurements to prove that Au was present electroless deposition to prepare SnO2/Pd/Au mixed thin films onto 981
917 in form of small metallic particles at the surface of SnO2 and it the surface of an alumina substrate. The sensor based on the 982
918 could be attributed to the “spill-over effect” to react with reducing composite film had fast response in the range of 134–1469 ppm 983
919 gases such as CO and H2, which were greatly different from Pd and toward hydrogen gas at room temperature [212]. They thought 984
920 Pt that were distributed at an atomic level on the surface and in the that the sensing mechanism was most probably due to conductive 985
921 bulk of SnO2 to have a tremendous effects on its bulk and surface changes by hydrogen atoms dissociated on the surface of Pd to 986
922 electronic properties [195]. Zhang et al. [196] used hollow SnO2 increase carrier density of n-type SnO2. The presence of two noble 987
923 hollow spheres for the first time to deposit Au nanoparticle and it metals could not only provide chemical sensitization but also 988
924 exhibited excellent properties to sensing ethanol due to its large electronic sensitization [213]. 989
925 surface area, abundant channels for gas diffusion and the catalytic
926 effects from Au [197]. The incorporation of Au nanoparticles could SnO2/polymer sensor materials 990
927 result in a composition that was capable of selectively sensing low
928 concentrations (upto 10 ppm CO in air) at the temperatures below Conducting polymers, such as polyaniline (PANI), polypyrrole 991
929 50  C [198]. Meanwhile, the content of Au nanoparticles incorpo- (PPy), polythiophene (PTH) and their derivatives, have been used 992
930 rated in SnO2 substrate is likely to be an important parameter to as the active layers of gas sensors since early 1980s [214]. Although 993
931 upgrade the sensing ability [199–201]. polymer-based gas sensing materials are applied to detect various 994
932 More recently, the designed preparation of pure and Au-loaded gases, they are usually used to detect a variety of VOCs or solvent 995
933 SnO2 hollow multilayered nanosheets for CO detection was vapors under a rather low work temperature [215]. It is well known 996
934 reported by Bing et al. [202]. The formation process involved that the electrical conductivity of some conducting polymers is 997
935 several unambiguous stages to obtain the final pure and Au-loaded affected by exposure to diverse organic and inorganic gases. 998
936 SnO2 hollow configurations. Sensors based on SnO2 hollow Semiconductive inorganic metal oxide/conducting polymer hybrid 999
937 nanospheres, double-shelled hollow nanoboxes, and hollow materials are currently of great interest for exploring high 1000
938 multilayered nanosheets are fabricated and their performances performance sensor materials, due to their synergetic or comple- 1001
939 were compared. As shown in Fig. 10, Au/SnO2 hollow nanosheets at mentary behavior that are not available from their single 1002
940 220  C exhibit about 7.6, 5.7, 5.0 times enhancement in response to counterpart. In this section, SnO2/polymer gas sensing composites 1003
941 50 ppm CO compared to hollow nanospheres, double-shelled are concerned and discussed briefly. Compared with inorganic 1004
942 hollow nanoboxes, and hollow multilayered nanosheets, respec- metal oxide, carbon nanomaterials and noble metals, the study of 1005
943 tively. At the same time, its response and recovery times are much SnO2/polymer composites is less investigated. 1006
944 shorter than others, showing a rapid response performance [201]. In 2009, Deshpande et al. [216] first fabricated SnO2 intercalat- 1007
945 The significantly improved performance for Au/SnO2 hollow ed PANI composite at room temperature by solution technique. 1008

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx 17

Fig. 10. (a) Dynamic response transients of sensors based on different hollow SnO2 nanostructures and (b) response transients of Au-loaded SnO2 sample to CO with the
increased concentration at 220  C (bar: 1 mm). (Reprinted from Ref. [202], copyright with permission from Elsevier).

1009 Pure PANI films became more resistive on exposure to NH3 gas, and 10 wt% anatase TiO2 and SnO2 nanoparticles, respectively, had 1025
1010 while the composite became less resistive on a similar exposure. the best response time (75 s) with a recovery time of 117 s at 1026
1011 The SnO2/PANI composite films showed good sensitivity, repro- environmental conditions and its response was strongly composi- 1027
1012 ducibility with relatively faster response for ammonia gas, at room tion-related. They also found that the presence of SnO2 nano- 1028
1013 temperature. They also provided a suitable explanation for such particles in the composites caused the response and recovery time Q17 1029
1014 behavior of these composite films [216]. Subsequently, several of the sensors to have a significant decrease [48]. 1030
1015 interesting papers ware reported on SnO2/PANI composite gas Joulazadeh and Navarchian [223] fabricated PPy/SnO2 nano- 1031
1016 sensing materials. Patil et al. [217,218] synthesized SnO2/PANI and composite for ammonia, methanol and ethanol detection. The 1032
1017 investigated its performances as sensing materials for detecting sensors based on PPy had high selectivity for ammonia with 1033
1018 liquefied petroleum gas [217] and humidity [218] under low respect to methanol and ethanol, because ammonia could affect 1034
1019 temperatures. The SnO2/PANI sensing materials were usually the interchain conductivity in PPy macromolecules, whereas 1035
1020 investigated to detect NH3 [218–222] and NO2 [220,221] gases. The alcohols could only influence the interchains conduction process. 1036
1021 gas sensing mechanisms to detect NO2 and NH3 by SnO2/PANI Meanwhile, PPy behaved as p-type semiconductor whereas the 1037
1022 Q16 composite were further discussed in Ref. [220] while Nasirian and SnO2 was n-type semiconductor. Therefore the p–n heterojunc- 1038
1023 Moghaddam [48] applied the ternary composite of TiO2/SnO2/PANI tions would be formed in the interface of PPy/SnO2 hybrids, which 1039
1024 for hydrogen sensing application. The sensor materials with 20% would generate a unique electron donor–acceptor system [223]. 1040

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

18 J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx

1041 NH3 molecules adsorbed in the nanocomposite would withdrew nanostructures or composites [231]. Recently, it had been 1066
1042 the protons in the N+–H site in PPy, leading to a decrease in the extensively applied to fabricate polymer/SnO2 sensor materials 1067
1043 holes concentration of PPy chain. Therefore, PPy was deprotonated which had a large aspect ratio, high surface area to volume ratio 1068
1044 and exhibited increased resistance. The composite thus displayed a and large porosity, including PPy/SnO2 [224,225], PANI/SnO2 1069
1045 larger change in resistance and higher response magnitude [232] composite. Jiang et al. [224] investigated the synergic Q19 1070
1046 compared to PPy alone [224,225]. effect within n-type inorganic-p-type organic hybrid in gas 1071
1047 Q18 Wu et al. investigated a lot of SnO2-polymer sensor materials sensors and used 1D SnO2/PPy nanofibers prepared by electro- 1072
1048 previously [226–229]. They firstly reported the fabrication of spinning as a model. Interfacial Ag nanoparticles were further 1073
1049 polythiophene/SnO2 composites by the in situ chemical oxidative introduced into the hybrid. They divided the cross-section of the 1074
1050 polymerization method and the composites exhibited higher whole SnO2–PPy core–shell into four regions and found that the 1075
1051 sensitivity to a low NOx concentration at low temperature of 90  C thicknesses of the PPy layer and the region within the PPy shell 1076
1052 [228]. They also synthesized SnO2 hollow spheres/PTH hybrid attached to the SnO2 core were critical keys for the sensing 1077
1053 material by an in situ polymerization method and the composite performances. It showed readers a better understand the 1078
1054 showed much higher response and fast response–recovery time in relationship between the microstructures and their gas sensing 1079
1055 comparison to pure PTH. However, when compared with pure n- performances [224]. Fig. 11 is the TEM images of SnO2–PPy core– 1080
1056 type SnO2, the response of the composite was lower due to their shell hybrid fibers with different polymerization times and PPy 1081
1057 different sensing mechanisms [229]. With the assistance of thicknesses [224]. 1082
1058 ultrasound, SnO2 nanoparticles could be coated by PTH during Komolov et al. investigated the electrical conductivity of thin 1083
1059 polymerization process in the presence of SnO2 and the composite films based on SnO2 nanoparticles and copper phthalocyanine- 1084
1060 containing 20 wt% SnO2 showed the maximum sensitivity at room 3,40 ,400 ,4000 -tetrasulfonic acid tetrasodium salt(CuPc-4SO3Na) mol- 1085
1061 temperature to liquefied petroleum gas. It was mainly attributed to ecules to detect ammonia [233]. They found that, in case of a 1086
1062 the effects of p–n heterojunctions between the two components single-component film of SnO2 nanoparticles, an increase in the 1087
1063 [230]. electrical conductivity reached 100% and was not completely 1088
1064 Electrospinning method has been deemed to be the simplest reversible at room temperature after evacuation of the gas. For the 1089
1065 and most versatile technique capable of generating 1D polymer single-component CuPc-4SO3Na film and the composite CuPc- 1090

Fig. 11. TEM images of SnO2-PPy core–shell hybrid nanofibers with different polymerization times, (a) 0.5 h, (b) 1 h, (c) 2 h, and (d) 3 h. The insets show the optical
photographs. (Reprinted from Ref. [224], copyright with permission from RSC Publications).

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx 19

Fig. 12. (a) Schematic illustration of the preparation process for MoS2/SnO2 nanohybrids. The inset photographs show the MoS2 suspension in water before and after adding
the SnCl4 solution. (b) An SEM image of the 2H-MoS2 nanosheets. (c–d) SEM images of the hybrid, (c) overview of a MoS2/SnO2 film, (d) magnified image in the center.
(Reprinted from Ref. [237], copyright with permission from Wiley).

1091 4SO3Na/SnO2 film, the electrical conductivities increased by a and selectivity to NO2 with a low signal noise. The lower detection 1130
1092 factor of 400 and 150, respectively. Upon evacuation of ammonia to limit could reach a concentration of 0.5 ppm. 1131
1093 the base pressure, the electrical conductivity decreased to the The in situ vertical growth of ultrathin SnO2 nanosheets on 1132
1094 initial values for 1 s. The gas adsorption to the organic material quasi-1D SiC nanofibers forming a hierarchical architecture via a 1133
1095 surface provides a better reversibility than to the SnO2 surface simple hydrothermal method was reported by Wang et al. [238]. 1134
1096 upon gas exposure/evacuation [234]. At the same time the gas Compared with commercial and pure SnO2 nanosheets, the SnO2/ 1135
1097 molecules adsorbed on the organic surface (within the composite SiC hierarchical composite exhibited a superior gas sensing 1136
1098 material) provide the formation of the polarization layer which performance including high sensitivity, excellent reproducibility, 1137
1099 affects the conduction of the SnO2 component of the composite outstanding selectivity and stability toward ethanol. Its perfor- 1138
1100 [235]. mance was attributed to the synergic effect of SnO2–SiC hetero- 1139

junction and the hierarchical structure as well as the vertical 1140


1101 Composite of SnO2/other materials growth of SnO2 nanosheets morphology. Karakuscu et al. [239] also 1141

applied thermal evaporation to decorate the cell walls of 1142


1102 In addition to above SnO2 composites for gas sensors, there are commercial SiC foams with 1D SnO2 nanobelts to form a porous 1143
1103 also other reports on the gas sensing materials, such as SnO2/MoS2, structure. Its gas sensing properties to NH3 and NO2 were 1144
1104 Q20 SnO2/SiC, Pd/SnO2/SiO2/Si heterojunction films, etc. systematically investigated in both dry and humid air atmosphere 1145
1105 MoS2 has been selected as a good choice to make a composite with/without UV activation. The SiC/SnO2 composite showed a 1146
1106 with SnO2. 2D materials can screen charge fluctuations better than higher performance to detect low concentration of gases than bare 1147
1107 1D materials, meanwhile, the planar structure like graphene eases SnO2 and SiC [239]. 1148
1108 pattern fabrication and measurement. Thus, 2D MoS2 can be also Lian et al. [240] fabricated a series of Pd/SnO2/SiO2/Si Q21 1149
1109 used as an ideal substrate for SnO2 nanoparticles. Yan et al. [236] heterojunction sensors that were produced using magnetron 1150
1110 reported that a new composite that could be successfully sputtering method. The sensor showed excellent H2 gas sensitivity. 1151
1111 fabricated by a two-step low temperature hydrothermal method. Pd/SnO2/SiO2/n or p-Si heterojunction films exhibited a response of 1152
1112 The fabricated SnO2 nanoparticles were dispersed on the surfaces 17363–1.0% H2 or 611–1317% to 0.05% H2 at room temperature. 1153
1113 of MoS2 nanosheets that were a promising supporting substrate for The sensor had some advantages including easy-fabrication, 1154
1114 the preventing the interparticle aggregation of SnO2. Experimental cheapness, operating at room temperature, good selectivity and 1155
1115 results showed that SnO2/MoS2 composites exhibit superior gas- stability, etc. [240]. 1156
1116 sensing performance to ethanol in comparison with pure SnO2
1117 nanoparticles [236]. Cui et al. [237] also fabricated the nanohybrid Conclusions and outlooks 1157
1118 of SnO2 nanocrystal-decorated crumpled MoS2 nanosheets and it
1119 exhibited high sensitivity, excellent selectivity, and repeatability to Pure SnO2 and its composite are unique and attractive gas 1158
1120 NO2 under a practical dry air environment. These SnO2 nano- sensing materials for sensors fabrication to detect toxic, flammable 1159
1121 particles were strong p-dopants and greatly enhanced the stability or explosive gases. Up to now, extensive studies have established 1160
1122 of MnS2 nanosheets to prevent the interaction between the oxygen that three basic factors are important in controlling the gas sensing 1161
1123 in air and the MnS2, as shown in Fig. 12 [237]. Fig. 12(a) illustrates properties of SnO2 gas sensors. The factors are (1) the grain size of 1162
1124 the synthesis procedure for crumpled 2D MoS2/SnO2 nanohybrids. SnO2 particles, (2) microstructure of the sensing materials and 1163
1125 The MoS2/SnO2 nanohybrids exhibit a distinct crumpled structure body, and (3) surface modification of SnO2 particles. The sensors 1164
1126 (Fig. 12c). The magnified images (Fig. 12d) reveal that the MoS2/ made of SnO2 composites usually exhibited improved perform- 1165
1127 SnO2 nanohybrids still maintain the 2D structure of MoS2, but with ances compared with those of pure SnO2 in term of sensitivity, 1166
1128 a large number of wrinkles on their surfaces. Its sensing reversibility, response time, detection limit, etc. Thus, more and 1167
1129 performance showed a high sensitivity, excellent repeatability, more investigations have proved that design and fabrication of 1168

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

20 J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx

1169 SnO2 composites is a preferential route to obtain high-perfor- [17] D. Han, B.Q. Han, R. Han, S.J. Deng, Y. Wang, Q. Li, Y.D. Wang, New J. Chem. 38
1170 (2014) 2443. 1220
mance gas sensors.
1171 [18] Y.B. Shen, X.M. Cao, B.Q. Zhang, D.Z. Wei, J.W. Ma, W.G. Liu, C. Han, Y.T. Shen, J.
In this work, an overview is provided for the recent progress of Alloys Compd. 593 (2014) 271. 1221
1172 gas sensing studies on SnO2 composites, typically in the last five [19] S.H. Mohamed, J. Alloys Compd. 510 (2012) 119.
1173 years, including almost categories of composite systems, such as [20] J. Wu, Q. Huang, D. Zeng, S. Zhang, L. Yang, D. Xia, Z. Xiong, C. Xie, Sens.
Actuators B: Chem. 198 (2014) 62. 1222
1174 SnO2/inorganic metal oxide, SnO2/CNTs, SnO2/graphene, SnO2/ [21] L. Shi, H.L. Lin, Langmuir 27 (2011) 3977.
1175 noble metals, SnO2/polymer, and SnO/other materials. Among [22] X.B. Wang, Y.Y. Wang, F. Tian, H.J. Liang, K. Wang, X.H. Zhao, Z.S. Lu, K. Jiang, L.
1176 Yang, X.D. Lou, J. Phys. Chem. C 119 (2015) 15963. 1223
these composites, SnO2/inorganic metal oxide, SnO2/graphene,
1177 [23] R.V. Hagen, M. Sneha, S. Mathur, J. Am. Ceram. Soc. 97 (2014) 1035.
and SnO2/noble metals exhibit very promising for practical [24] H.B. Ren, W. Zhao, L.Y. Wang, S.O. Ryu, C.P. Gu, J. Alloys Compd. 653 (2015) 611.
1178 application, due to their low working temperature, high sensitivity [25] W.X. Jin, S.Y. Ma, Z.Z. Tie, W.Q. Li, J. Luo, L. Cheng, X.L. Xu, T.T. Wang, X.H. Jiang,
1179 1224
and stability. However, there are also still many challenges in this Y.Z. Mao, Appl. Surf. Sci. 353 (2015) 71.
1180 [26] F. Wei, H. Zhang, M. Nguyen, M. Ying, R. Gao, Z. Jiao, Sens. Actuators B: Chem.
field. For example, how to precisely control the dispersion of guest 1225
1181 215 (2015) 15.
materials into the SnO2 block? etc. [27] I. Paulowicz, V. Hrkar, S. Kaps, V. Cretu, O. Lupan, T. Braniste, V. Duppel, I.
1182 1226
The introduction of components with different chemical Tiginyannu, L. Kienle, R. Adelung, Y.K. Mishra, Adv. Electron. Mater. 1 (2015)
1183 1500081. 1227
compositions can be conveniently realized with the rapid
1184 [28] X. Zhao, W. Shi, H. Mu, H. Xie, F. Liu, J. Alloys Compd. 659 (2016) 60.
development of preparation strategies. Although significant [29] S. Maeng, S.W. Kim, D.H. Lee, S.E. Moon, K.C. Kim, A. Maiti, ACS Appl. Mater.
1185 progress has been made in the design and synthesis of SnO2 Interfaces 6 (2014) 357. 1228
1186 composites for gas sensing application, further work are still [30] Z.W. Chen, Z. Jiao, M.H. Wu, C.H. Shek, C.M.L. Wu, J.K.L. Lai, Prog. Mater. Sci. 56
(2011) 901. 1229
1187 required to better understand the synergistic interactions between [31] M. Batzill, U. Diebold, Prog. Surf. Sci. 79 (2005) 47.
1188 each component in the composite and the underlying sensing [32] I.D. Kim, A. Rothschild, H.L. Tuller, Acta Mater. 61 (2013) 974.
1189 Q22 mechanism for each of them. On basis of these questions being [33] V.V. krivetskiy, M.N. Rumyantseva, A.M. Gaskov, Russ. Chem. Rev. 82 (2013)
917. 1230
1190 well resolved, better scientific composition and structure design [34] H. Korotcenkov, S.D. Han, B.K. Cho, V. Brinzari, Crit. Rev. Solid State 34 (2009)
1191 for SnO2 composites can be achieved. Then high-performance gas 1231
1.
1192 sensors can be fabricated. [35] H.J. Kim, J.H. Lee, Sens. Actuators B: Chem. 192 (2014) 607.
1193 [36] N. Barsan, D. Koziej, U. Weimar, Sens. Actuators B: Chem. 121 (2007) 18.
Most previously reported results confirmed that the SnO2
[37] G. Korotcenkov, Mater. Sci. Eng. B 139 (2007) 1.
1194 composites demonstrated an enhanced gas-sensing performance [38] A.V. Marikutsa, M.N. Rumyantseva, A.M. Gaskov, A.M. Samoylov, Inorg. Mater.
1195 1232
compared with single SnO2 alone. Most of composite are binary 51 (2015) 1329.
1196 [39] A. Gurlo, Nanoscale 3 (2011) 154.
with two different components. However, there are also some
1197 [40] M. Tiemann, Chem. Eur. J. 13 (2007) 8376.
valuable attempts using ternary composites for gas sensing [41] V. Kamble, A. Umarji, AIP Adv. 5 (2015) 037138.
1198 application and the composite also usually exhibits better [42] Y. Li, C. Chopra, J. Catal. 329 (2015) 514.
1199 Q23 performance than single or binary systems. It is also an easy [43] J. Wang, Z. Chen, Y. Liu, C.H. Shek, C.M.L. Wu, J.K.L. Lai, Sol. Energy Mater. Sol.
Cells 128 (2014) 254. 1233
1200 and efficient route to obtain excellent SnO2-based gas sensing [44] Z. Chen, C.H. Shek, C.M.L. Wu, Nanoscale 7 (2015) 15532.
1201 materials at present. We hope that the present review paper can [45] L. Li, C.M. Zhang, W. Chen, Nanoscale 7 (2015) 12133.
1202 stimulate the further development of SnO2-based composites for [46] A. Shanmugasundaram, B. Pratyay, L. Satyanarayana, S.V. Manorama, Sens.
Actuators B: Chem. 185 (2013) 265. 1234
1203 gas sensor application with high performances in the future. [47] D. Shaposhnik, R. Pavelko, E. Llobet, F. Gispert-Guirado, X. Vilanova, Procedia
Eng. 25 (2011) 1133. 1235
[48] S. Nasirian, H.M. Moghaddam, Appl. Surf. Sci. 328 (2015) 395.
1204 Acknowledgements [49] P. Li, H.Q. Fan, Y. Cai, Sens. Actuators B: Chem. 185 (2013) 110.
[50] S. Ghosh, M. Narjinary, A. Sen, R. Bandyopadhyay, S. Roy, Sens. Actuators B:
Chem. 203 (2014) 490. 1236
1205 Q24 This project was supported by the Experimental technology
[51] H. Yamaura, Y. Iwasaki, S. Hirao, H. Yahiro, Sens. Actuators B: Chem. 153
1206 research project of Zhejiang University No. SYB201602 and 1237
(2011) 465.
1207 “Electronic Science and technology” Zhejiang Open Foundation [52] C.T. Wang, M.T. Chen, D.L. Lai, J. Am. Ceram. Soc. 94 (2011) 4471.
1208 [53] O. Singh, R.C. Singh, Mater. Res. Bull. 47 (2012) 557.
of the Most Important Subjects.
[54] S. Hemmati, A.A. Firooz, A.A. Khodadadi, Y. Mortazavi, Sens. Actuators B:
Chem. 160 (2011) 1298. 1238

1209 [55] M.Y. Wang, L.F. Zhu, C.Y. Zhang, G.S. Gai, X.W. Ji, B.H. Li, Y.W. Yao, Sens.
References Actuators B: Chem. 224 (2016) 478. 1239
[56] V.K. Tomer, S. Duhan, J. Mater. Chem. A 4 (2016) 1033.
[1] T. Wagner, S. Haffer, C. Weinberger, D. Klaus, M. Tiemann, Chem. Soc. Rev. 42 [57] S. Majumdar, Ceram. Int. 41 (2015) 14350.
1210 (2013) 4036. [58] S.F. Zhang, F. Ren, W. Wu, J. Zhou, X.H. Xiao, L.L. Sun, Y. Liu, C.Z. Jiang, Phys.
[2] X. Liu, S.T. Cheng, H. Liu, S. Hu, D.Q. Zhang, H.S. Ning, Sensors 12 (2012) 9635. Chem. Chem. Phys. 15 (2013) 8228. 1240
Q25 [3] A. Afzal, N. Cioffi, L. Sabbatini, L. Toris, Sens. Actuators B: Chem. 171–172 [59] H. Li, W.Y. Xie, T.J. Ye, B. Liu, S.H. Xiao, C.X. Wang, Y.R. Wang, Q.H. Li, T.H. Wang,
1211 (2012) 25. ACS Appl. Mater. Interfaces 7 (2015) 24887. 1241
[4] W.J. Yuan, G.Q. Shi, J. Mater. Chem. A 1 (2013) 10078. [60] H. Li, B. Liu, D.P. Cai, Y.R. Wang, Y. Liu, L. Mei, L.L. Wang, D.D. Wang, Q.H. Li, T.H.
[5] J. Zhang, X.H. Liu, G. Neri, N. Pinna, Adv. Mater. 28 (2016) 795. Wang, J. Mater. Chem. A 2 (2014) 6854. 1242
[6] C.X. Wang, L.W. Yin, L.Y. Zang, D. Xiang, R. Gao, Sensor 10 (2010) 2088. [61] X.C. Ma, H.Y. Song, C.S. Guan, Sens. Actuators B: Chem. 188 (2013) 193.
[7] M.D. Arienzo, D. Cristofori, R. Scotti, F. Morazzoni, Chem. Mater. 25 (2013) [62] H.W. Kim, H.G. Na, Y.J. Kwon, H.Y. Cho, C.M. Lee, Sens. Actuators B: Chem. 219
1212 3675. (2015) 22. 1243
[8] Z. Chen, D. Pan, B. Zhao, G. Ding, Z. Jiao, M. Wu, C.H. Shek, C.M.L. Wu, J.K.L. Lai, [63] G.Y. Lu, J. Xu, J.B. Sun, Y.S. Yu, Y.Q. Zhang, F.M. Liu, Sens. Actuators B: Chem. 162
1213 ACS Nano 4 (2010) 1202. (2012) 82. 1244
[9] H.K. Wang, A.L. Rogach, Chem. Mater. 26 (2014) 123. [64] S.S. Kim, H.G. Na, S.W. Choi, D.S. Kwak, H.W. Kim, J. Phys. D: Appl. Phys. 45
[10] S. Das, V. Jayaraman, Prog. Mater. Sci. 66 (2014) 112. (2012) 205301. 1245
[11] R.A. Kadir, Z.Y. Li, A.Z. Sadek, R.A. Rani, A.S. Zoolfakar, M.R. Field, J.Z. Ou, A.F. [65] X.P. Li, Z.Y. Gu, J.H. Cho, H.W. Sun, P. Kurup, Sens. Actuators B: Chem. 158
1214 Chirmes, K. Kalantar-zadeh, J. Phys. Chem. C 118 (2014) 3129. (2011) 199. 1246
[12] J.J. Wu, D.W. Zeng, S.Q. Tian, K. Xu, D.G. Li, C.S. Xie, J. Mater. Sci. 50 (2015) [66] Q. Qi, Y.C. Zou, M.H. Fan, Y.P. Liu, S. Gao, P.P. Wang, Y. He, D.J. Wang, G.D. Li,
1215 7725. Sens. Actuators B: Chem. 203 (2014) 111. 1247
[13] Z. Chen, D. Pan, Z. Li, Z. Jiao, M. Wu, C.H. Shek, C.M.L. Wu, J.K.L. Lai, Chem. Rev. [67] S. Xu, J. Gao, L.L. Wang, K. Kan, Y. Xie, P.K. Shen, L. Li, K.Y. Shi, Nanoscale 7
1216 114 (2014) 7442. (2015) 14643. 1248
[14] J.P. Cheng, B.B. Wang, M.G. Zhao, F. Liu, X.B. Zhang, Sens. Actuators B: Chem. [68] R. Fiz, F. Hernandez-Ramirez, T. Fischer, L. Lopez-Conesa, S. Estrade, F. Peiro, S.
1217 190 (2014) 78. Mathur, J. Phys. Chem. C 117 (2013) 10086. 1249
[15] L. Luo, Q. Jiang, G. Qin, K. Zhao, G. Du, H. Wang, H. Zhao, Sens. Actuators B: [69] L.L. Xing, S. Yuan, Z.H. Chen, Y.J. Chen, X.Y. Xue, Nanotechnology 22 (2011)
1218 Chem. 218 (2015) 205. 225502. 1250
[16] E. Brunet, T. Maier, G.C. Mutinati, S. Steinhauer, A. Kock, C. Gspan, W. Grogger, [70] G.H. Chen, S.Z. Ji, H.D. Li, X.L. Kang, S.J. Chang, Y.N. Wang, G.W. Yu, J.R. Lu, J.
1219 Sens. Actuators B: Chem. 165 (2012) 110. Claverie, Y.H. Sang, H. Liu, ACS Appl. Mater. Interfaces 7 (2015) 24950. 1251

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx 21

[71] A. Katoch, J.H. Kim, Y.J. Kwon, H.W. Kim, S.S. Kim, ACS Appl. Mater. Interfaces [122] H.C. Kim, C.S. Park, K.M. Kang, M.H. Hong, Y.J. Choi, H.H. Park, New J. Chem. 39
1252 7 (2015) 11351. (2015) 2256. 1287
[72] S.H. Park, S.Y. An, Y.H. Mun, C.M. Lee, ACS Appl. Mater. Interfaces 5 (2013) [123] Y.L. Cao, Y.Z. Li, D.Z. Jia, J. Xie, RSC Adv. 4 (2014) 46179.
1253 4285. [124] G. Neri, S.G. Leonardi, M. Latino, N. Donato, S. Baek, D.E. Conte, P.A. Russo, N.
[73] S.W. Choi, A. Katoch, G.J. Sun, J.H. Kim, S.H. Kim, S.S. Kim, ACS Appl. Mater. Pinna, Sens. Actuators B: Chem. 179 (2013) 61. 1288
1254 Interfaces 6 (2014) 8281. [125] D.Z. Zhang, A.M. Liu, H.Y. Chang, B.K. Xia, RSC Adv. 5 (2015) 3016.
[74] J.H. Kim, A. Katoch, S.H. Kim, S.S. Kim, ACS Appl. Mater. Interfaces 7 (2015) [126] I.S. Kang, H.M. So, G.S. Bang, J.H. Kwak, J.O. Lee, C.W. Ahn, Appl. Phys. Lett. 101
1255 15351. (2012) 123504. 1289
[75] X. Chen, Z. Guo, W.H. Xu, H.B. Yao, M.Q. Li, J.H. Liu, X.J. Huang, S.H. Yu, Adv. [127] K. Inyawilert, A. Wisitsoraat, C. Scrpachaubwong, A. Tuantranont, S.
1256 Funct. Mater. 21 (2011) 2049. Phanichphant, C. Liewhiran, Sens. Actuators B: Chem. 209 (2015) 40. 1290
[76] S.L. Bai, H.Y. Liu, R.X. Luo, A.F. Chen, D.Q. Li, RSC Adv. 4 (2014) 62862. [128] Z. Zhang, X. Zou, L. Xu, L. Liao, W. Liu, J. Ho, X. Xiao, C. Jiang, J. Li, Nanoscale 7
[77] Q.X. Yu, J.H. Zhu, Z.Y. Xu, X.T. Huang, Sens. Actuators B: Chem. 213 (2015) 27. (2015) 10078. 1291
[78] H. Shan, C.B. Liu, L. Liu, J.B. Zhang, H.Y. Li, Z. Liu, X.B. Zhang, X.Q. Bo, X. Chi, ACS [129] L.Z. Yu, L.C. Zhang, H.J. Song, X.M. Jiang, Y. Lv, CrystEngComm 16 (2014) 3331.
1257 Appl. Mater. Interfaces 5 (2013) 6376. [130] S.K. Lee, D. Chang, S.W. Kim, J. Hazard. Mater. 268 (2014) 110.
[79] M.K. Verma, V. Gupta, Sens. Actuators B: Chem. 166 (2012) 378. [131] D.Z. Zhang, H.Y. Chang, P. Li, R.H. Liu, Q.Z. Xue, Sens. Actuators B: Chem. 225
[80] G.L. Cui, M.Z. Zhang, G.T. Zou, Sci. Rep. 3 (2013) 1250. (2016) 233. 1292
[81] D. Haridas, V. Gupta, Sens. Actuators B: Chem. 166–167 (2012) 156. [132] B. Li, Y. He, S. Lei, S. Najmaei, Y. Gong, X. Wang, J. Zhang, L. Ma, Y. Yang, S. Hong,
[82] P. Tyagi, A. Sharma, M. Tomar, V. Gupta, Sens. Actuators B: Chem. 224 (2016) J. Hao, G. Shi, A. George, K. Keyshar, X. Zhang, P. Dong, L. Ge, R. Vajtai, J. Lou, Y. 1293
1258 282. Jung, P. Ajayan, Nano Lett. 15 (2015) 5089. 1294
[83] A. Ebrahimi, A. Pirouz, Y. Abdi, S. Azimi, S. Mohajerzadeh, Sens. Actuators B: [133] K.Y. Ma, W.J. Zhao, J.P. Cheng, F. Liu, X.B. Zhang, J. Colloid Interface Sci. 468
1259 Chem. 173 (2012) 802. (2016) 238. 1295
[84] S. Singh, N. Verma, A. Singh, B.C. Yadav, Mat. Sci. Semicond. Process. 18 (2014) [134] C. Chen, L. Wang, Y. Liu, Z. Chen, D. Pan, Z. Li, Z. Jiao, P. Hu, C.H. Shek, C.M.L.
1260 88. Wu, J.K.L. Lai, M.H. Wu, Langmuir 29 (2013) 4111. 1296
[85] S.C. Lee, B.W. Hwang, S.J. Lee, H.Y. Choi, S.Y. Kim, S.Y. Jung, D. Ragupathy, D.D. [135] J.P. Cheng, L. Liu, J. Zhang, F. Liu, X.B. Zhang, J. Electroanal. Chem. 722 (2014)
1261 Lee, J.C. Kim, Sens. Actuators B: Chem. 160 (2011) 1328. 23. 1297
[86] Z.D. Lin, C.L. Guo, Q.Q. Fu, W.L. Song, Mater. Lett. 102 (2013) 47. [136] R. Huang, L. Wang, Q. Zhang, Z. Chen, Z. Li, D. Pan, B. Zhao, M. Wu, C.M.L. Wu,
[87] G.Z. Zhang, S.P. Zhang, L. Yang, Z.J. Zou, D.W. Zeng, C.S. Xie, Sens. Actuators B: C.H. Shek, ACS Nano 11 (2015) 11351. 1298
1262 Chem. 188 (2013) 137. [137] M. Malek Alaie, M. Jahangiri, A.M. Rashidi, A. Haghighi Asl, N. Izadi, J. Ind. Eng.
[88] Q.J. Jiang, C.J. Wu, L.S. Feng, L. Gong, Z.Z. Ye, J.G. Lu, Appl. Surf. Sci. 357 (2015) Chem. 29 (2015) 97. 1299
1263 1536. [138] J. Xu, J. Wu, L. Luo, X. Chen, H. Qin, V. Dravid, S. Mi, C. Jia, J. Power Souces 274
[89] L.I. Trakhtenberg, G.N. Gerasimov, V.F. Gromov, T.V. Belysheva, O.J. Ilegbusi, (2015) 816. 1300
1264 Sens. Actuators B: Chem. 169 (2012) 32. [139] F. Schedin, A.K. Geim, S.V. Morozov, E.W. Hill, P. Blake, M.I. Katsnelson, K.S.
[90] S. Somacescu, R. Scurtu, G. Epurescu, R. Pascu, B. Mitu, P. Osiceanu, M. Movoselov, Nat. Mater. 6 (2007) 652. 1301
1265 Dinescu, Appl. Surf. Sci. 278 (2013) 146. [140] D. Zhang, J. Liu, H. Chang, A. Liu, B. Xia, RSC Adv. 5 (2015) 18666.
[91] S. Durrani, M.F. Al-Kuhaili, I.A. Bakhtiari, M.B. Haider, Sensors 12 (2012) 2598. [141] D.M. Guo, P.J. Cai, J. Sun, W.N. He, X.H. Wu, T. Zhang, X. Wang, X.T. Zhang,
[92] P. Sun, C. Wang, J. Liu, X. Zhou, X. Li, X. Hu, G.Y. Lu, ACS Appl. Mater. Interfaces Carbon 99 (2016) 571. 1302
1266 7 (2015) 19119. [142] Q.Q. Lin, Y. Li, M.J. Yang, Sens. Actuators B: Chem. 173 (2012) 139.
[93] H.R. Kim, A. Haensch, I.D. Kim, N. Barsan, U. Weimar, J.H. Lee, Adv. Funct. [143] L. Li, S.J. He, M.M. Liu, C.M. Zhang, W. Chen, Anal. Chem. 87 (2015) 1638.
1267 Mater. 21 (2011) 4456. [144] Y. Xiao, Q.Y. Yang, Z.Y. Wang, R. Zhang, Y. Gao, P. Sun, Y.F. Sun, G.Y. Lu, Sens.
[94] P. Sun, X. Zhou, C. Wang, K. Shimanoe, G.Y. Lu, N. Yamazoe, J. Mater. Chem. A 2 Actuators B: Chem. 227 (2016) 419. 1303
1268 (2014) 1302. [145] S. Mao, S.M. Cui, G.H. Lu, K.H. Yu, Z.H. Wen, J.H. Chen, J. Mater. Chem. 22
[95] A.K. Nayak, R. Ghosh, S. Santra, P.K. Guha, D. Pradhan, Nanoscale 7 (2015) (2012) 11009. 1304
1269 12460. [146] S.M. Cui, Z.H. Wen, E.C. Mattson, S. Mao, J.B. Chang, M. Weinert, C.J.
[96] J.H. Park, M.S. Cho, D. Lim, J.G. Park, J. Nanosci. Nanotech. 14 (2014) 8038. Hirachmugl, M. Gajdardziska-Josifovska, J.H. Chen, J. Mater. Chem. A 1 (2013) 1305
[97] B. Li, X. Wang, H.Y. Jung, Y.L. Kim, J.Y. Robinson, M. Zalalutdinov, S.H. Hong, J. 4462. 1306
1270 Hao, P.M. Ajayan, K.T. Wan, Y.J. Jung, Sci. Rep. 5 (2015) 15908. [147] S. Liu, Z.Y. Wang, Y. Zhang, C.B. Zhang, T. Zhang, Sens. Actuators B: Chem. 211
[98] M. Li, F. Liu, J.P. Cheng, J. Ying, X.B. Zhang, J. Alloys Compd. 635 (2015) 225. (2015) 318. 1307
[99] X. Wang, J.F. Najem, S.C. Wong, K.T. Wan, J. Appl. Phys. 111 (2012) 024315. [148] Z.Y. Wang, Y. Zhang, S. Liu, T. Zhang, Sens. Actuators B: Chem. 22 (2016) 893.
[100] J. Kong, N.R. Franklin, C.W. Zhou, M.G. Chapline, S. Peng, K. Cho, H.J. Dai, [149] S. Ge, H.Z. Zheng, Y.F. Sun, Z. Jin, J.H. Shan, C. Wang, H. Wu, M.Q. Li, F.L. Meng, J.
1271 Science 287 (2000) 622. Alloys Compd. 659 (2016) 127. 1308
[101] S.M. Cui, H.H. Pu, G.H. Lu, Z.H. Wen, E.C. Mattson, C. Hirschmug, M. [150] P.A. Russo, N. Donato, S.G. Leonardi, S. Baek, D.E. Conte, G. Neri, N. Pinna,
1272 Gadjardziska-Josifovska, M. Weinert, J.H. Chen, ACS Appl. Mater. Interfaces 4 Angew. Chem. Int. Ed. 51 (2012) 11053. 1309
1273 (2012) 4898. [151] C.S. Lee, J.H. Choi, Y.H. Park, J. Ind. Eng. Chem. 29 (2015) 321.
[102] S. Mao, S. Cui, K. Yu, Z. Wen, G. Lu, J. Chen, Nanoscale 2 (2012) 1275. [152] J.Y. Jung, C.S. Lee, J. Ind. Eng. Chem. 17 (2011) 237.
[103] C. Marichy, P.A. Russo, M. Latino, J.P. Tessonnier, M.G. Willinger, N. Donato, G. [153] M. Yuasa, T. Kida, K. Shimanoe, ACS Appl. Mater. Interfaces 4 (2012) 4231.
1274 Neri, N. Pinna, J. Phys. Chem. C 117 (2013) 19729. [154] M. Choudhary, V.N. Mishra, R. Dwivedi, J. Electron. Mater. 42 (2013) 2793.
[104] Y. Jia, P.Y. Wu, Y.P. Jiang, Q.Y. Zhang, S.S. Zhou, F. Fang, D.Y. Peng, New J. Chem. [155] L.L. Xing, B. He, Z.H. Chen, X.Y. Xue, Solid State Sci. 15 (2013) 42.
1275 38 (2014) 1100. [156] M. Choudhary, V.N. Mishra, R. Dwivedi, J. Mater. Sci.: Mater. Electron. 24
[105] Q.H. Hu, S.T. Liu, Y.F. Lian, Phys. Status Solidi A 211 (2014) 2729. (2013) 2824. 1310
[106] S. Salehi, E. Nikan, A.A. Khodadadi, Y. Mortazavi, Sens. Actuators B: Chem. 205 [157] C. Liewhiran, N. Tamaekong, A. Wisitsoraat, A. Tuantranont, S. Phanichphant,
1276 (2014) 261. Sens. Actuators B: Chem. 176 (2013) 893. 1311
[107] S.B. Naghadeh, S. Vahdatifar, Y. Mortazavi, A.A. Khodadadi, A. Abbasi, Sens. [158] K.C. Lee, Y.J. Chiang, y.C. Lin, F.M. Pan, Sens. Actuators B: Chem. 226 (2016)
1277 Actuators B: Chem. 223 (2016) 252. 457. 1312
[108] J.M. Xu, J. Zhang, B.B. Wang, F. Liu, J. Alloys Compd. 619 (2015) 361. [159] A.A. Zhukova, M.N. Rumyantseva, V.B. Zaytsev, A.V. Arbakumov, A.M. Gaskov,
[109] H.F. Dai, P. Xiao, Q. Lou, Phys. Status Solidi A 208 (2011) 1714. J. Alloys Compd. 565 (2013) 6. 1313
[110] S. Majumdar, P. Nag, P.S. Devi, Mater. Chem. Phys. 147 (2014) 79. [160] D.D. Trung, N.D. Hoa, P.V. Tong, N.V. Duy, T.D. Dao, H.V. Chung, T. Nagao, N.V.
[111] R. Leghrib, A. Felten, J.J. Pireaux, E. Llobet, Thin Solid Films 520 (2011) 966. Hieu, J. Hazard. Mater. 265 (2014) 124. 1314
[112] Y. Zhang, S. Cui, J. Chang, L.E. Ocola, J. Chen, Nanotechnology 24 (2013) [161] B. Kim, Y. Lu, A. Hannon, M. Meyyappan, J. Li, Sens. Actuators B: Chem. 177
1278 025503. (2013) 770. 1315
[113] S. Mubeen, M. Lai, T. Zhang, J.H. Lim, A. Muchandani, M.A. Dehusses, N.V. [162] M. Choudhary, C.N. Mishra, R. Dwivedi, J. Mater. Sci.: Mater. Electron. 25
1279 Myung, Electrochim. Acta 92 (2013) 484. (2015) 1331. 1316
[114] Z.K. Horastani, S.M. Sayedi, M.H. Sheikhi, Sens. Actuators B: Chem. 202 (2014) [163] N.M. Shaalan, T. Yamazaki, T. Kikuta, Sens. Actuators B: Chem. 166–167 (2012)
1280 461. 671. 1317
[115] S. Dhall, N. Jaggi, Sens. Actuators B: Chem. 210 (2015) 742. [164] L.P. Oleksenko, N.P. Maksymovych, E.V. Sokovykh, I.P. Matushko, A.I. Buvailo,
[116] H.C. Su, M. Zhang, W. Bosze, N.V. Myung, J. Electrochem. Soc. 161 (2014) B283. N. Dollahon, Sens. Actuators B: Chem. 196 (2014) 298. 1318
[117] F. Rigoni, G. Drera, S. Pagliara, A. Goldoni, L. Sangaletti, Carbon 80 (2014) 356. [165] Z.M. Cui, A. Mechai, L. Guo, W.G. Song, RSC Adv. 3 (2013) 14979.
[118] V.M. Aroutiounian, A.Z. Adamyan, E.A. Khachaturyan, Z.N. Adamyan, K. [166] J. Zhao, W.N. Wang, Y.P. Liu, J.M. Ma, X.W. Li, Y. Du, G.Y. Lu, Sens. Actuators B:
1281 Hernadi, Z. Pallai, Z. Nemeth, L. Forro, A. Magrez, E. Horvath, Sens. Actuators Chem. 160 (2011) 604. 1319
1282 B: Chem. 177 (2013) 308. [167] Y. Lin, W. Wei, Y.J. Li, F. Li, J.R. Zhou, D.M. Sun, Y. Chen, S.P. Ruan, J. Alloys
[119] S.G. Chatterjee, S. Chatterjee, A.K. Ray, A.K. Chakraborty, Sens. Actuators B: Compd. 651 (2015) 690. 1320
1283 Chem. 221 (2015) 1170. [168] S. Xu, K. Kan, Y. Yang, C. Jiang, J. Gao, L.Q. Jiang, P.K. Shen, L. Li, K.Y. Li, J. Alloys
[120] R. Ghosh, A.K. Nayak, S. Santra, D. Pradhan, P.K. Guha, RSC Adv. 5 (2015) Compd. 618 (2015) 240. 1321
1284 50165. [169] W.Q. Li, S.Y. Ma, Y.F. Li, X.B. Li, C.Y. Wang, X.H. Yang, L. Cheng, Y.Z. Mao, J. Luo,
[121] N. Tammanoon, A. Wisitsoraat, C. Sriprachuabwong, D. Phokharatul, A. D.J. Gengzang, G.X. Wang, X.L. Xu, J. Alloys Compd. 605 (2014) 80. 1322
1285 Tuantranont, S. Phanichphant, C. Liewhiran, ACS Appl. Mater. Interfaces 7 [170] D.J. Yang, I. Kamienchick, D.Y. Youn, A. Rothschild, I.D. Kim, Adv. Funct. Mater.
1286 (2015) 24338. 20 (2010) 4258. 1323

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008
G Model
JIEC 3043 1–22

22 J.P. Cheng et al. / Journal of Industrial and Engineering Chemistry xxx (2016) xxx–xxx

[171] Z.J. Wang, Z.Y. Li, T.T. Jiang, X.R. Xu, C. Wang, ACS Appl. Mater. Interfaces 5 [204] L.Z. Yu, H.J. Song, Y.R. Tang, L.C. Zhang, Y. Lv, Sens. Actuators B: Chem. 203
1324 (2013) 2013. (2014) 726. 1352
[172] S.H. Jeong, S. Kim, J. Cha, M.S. Son, S.H. Park, H.Y. Kim, M.H. Cho, M.H. [205] C.T. Wang, H.Y. Chen, Y.C. Chen, Sens. Actuators B: Chem. 176 (2013) 945.
1325 Whangbo, K.H. Yoo, S.J. Kim, Nano Lett. 13 (2013) 5938. [206] Z.Y. Li, X.G. Wang, T. Lin, J. Mater. Chem. A 2 (2014) 13655.
[173] K. Grobmann, K.E. Kovacs, D.K. Pham, L. Madler, N. Barsan, U. Weimar, Sens. [207] A. Marikutsa, V. Krivetskiy, L. Yashina, M. Rumyantseva, E. Konstantinova, A.
1326 Actuators B: Chem. 158 (2011) 388. Ponzoni, E. Comini, A. Abakumov, A. Gaskov, Sens. Actuators B: Chem. 175 1353
[174] S. Kundu, P.K. Warran, S.M. Mursalin, M. Narjinary, J. Mater. Sci: Mater. (2012) 186. 1354
1327 Electron. 26 (2015) 9865. [208] K.M. Kim, K.I. Choi, H.M. Jeong, H.J. Kim, H.R. Kim, J.H. Lee, Sens. Actuators B:
[175] L.P. Chikhale, J.Y. Patil, A.V. Rajgure, R.C. Pawar, I.S. Mulla, S.S. Suryavanshi, J. Chem. 166–167 (2012) 733. 1355
1328 Alloys Compd. 608 (2014) 133. [209] C. Sankar, V. Ponnuswamy, M. Manickam, R. Mariappan, R. Suresh, Appl. Surf.
[176] E. Comini, G. Faglia, G. Sberveglier, Sens. Actuators B: Chem. 78 (2001) 73. Sci. 349 (2015) 931. 1356
[177] C.L. Guo, Z.D. Lin, W.L. Song, X.H. Wang, Y.Y. Huang, K. Wang, J. Nanopart. Res. [210] Z.K. Horastani, S.M. Sayedi, M.H. Sheikhi, E. Rahimi, Mat. Sci. Semicond.
1329 15 (2013) 1998. Process. 35 (2015) 38. 1357
[178] F.H. Saboor, T. Ueda, K. Kamada, T. Hyodo, Y. Mortazavi, A.A. Khodadadi, Y. [211] I.S. Hwang, J.K. Choi, H.S. Woo, S.J. Kim, S.Y. Jung, T.Y. Seong, I.D. Kim, J.H. Lee,
1330 Shimizu, Sens. Actuators B: Chem. 223 (2016) 429. ACS Appl. Mater. Interfaces 3 (2011) 3140. 1358
[179] P.G. Su, L.Y. Yang, Sens. Actuators B: Chem. 223 (2016) 202. [212] X.G. Wen, M. Wang, C. Wang, J.Z. Jiang, Electrochim. Acta 56 (2011) 6524.
[180] F. Gyger, A. Sackmann, M. Hubner, P. Bockstaller, D. Gerthsen, H. Lichtenberg, [213] Y.V. Kaneti, J. Yue, J. Morciau, C. Chen, M. Liu, Y. Yuan, X. Jiang, A. Yu, Sens.
1331 J.D. Grunwaldt, N. Barsan, U. Weimar, C. Feldmann, Part. Part. Syst. Charact. 31 Actuators B: Chem. 219 (2015) 83. 1359
1332 (2014) 591. [214] H. Bai, G.Q. Shi, Sensors 7 (2007) 267.
[181] M.H.S. Abadi, M.N. Hamidon, A.H. Shaari, N. Abdullah, R. Wagiran, Sensors 11 [215] S. Sattari, A. Reyhani, M.R. Khanlari, M. Khabazian, H. Heydari, J. Ind. Eng.
1333 (2011) 7724. Chem. 20 (2014) 1761. 1360
[182] S. Vahdatifar, A.A. Khodadadi, Y. Mortazavi, Sens. Actuators B: Chem. 191 [216] N.G. Deshpande, Y.G. Gudage, R. Sharma, J.C. Vyas, J.B. Kim, Y.P. Lee, Sens.
1334 (2014) 421. Actuators B: Chem. 138 (2009) 76. 1361
[183] M.H. Saveri, Y. Mortazavi, A.A. Khodadadi, Sens. Actuators B: Chem. 206 [217] D. Patil, K. Kolhe, H.S. Potdar, P. Patil, J. Appl. Phys. 110 (2011) 124501.
1335 (2015) 617. [218] D. Patil, P. Patil, Y.K. Seo, Y.K. Hwang, Sens. Actuators B: Chem. 148 (2010) 41.
[184] S. Rane, S. Arbuj, A. Rane, S. Gosavi, J. Mater. Sci.: Mater. Electron. 26 (2015) [219] H.L. Tai, Y.D. Jiang, G.Z. Xie, J.S. Yu, J. Mater. Sci. Technol. 26 (2010) 605–613.
1336 3707. [220] C.A. Betty, S. Choudhury, S. Arora, Sens. Actuators B: Chem. 220 (2015) 288.
[185] K. Wang, T.Y. Zhao, G. Lian, Q.Q. Yu, C.H. Luan, Q.L. Wang, D.L. Cui, Sens. [221] G.D. Khuspe, S.T. Navale, D.K. Bandgar, R.D. Sakhare, M.A. Chougule, V.B. Patil,
1337 Actuators B: Chem. 184 (2013) 33. Electron. Mater. Lett. 10 (2014) 191. 1362
[186] T.V.K. Karthik, M. de la L. Olvera, A. Maldonado, V. Velumurugan, Mat. Sci. [222] S.L. Bai, Y.L. Tian, M. Cui, J.H. Sun, Y. Tian, R.X. Luo, A.F. Chen, D.Q. Li, Sens.
1338 Semicond. Process. 37 (2015) 143. Actuators B: Chem. 226 (2016) 540. 1363
[187] Y.B. Shen, T. Yamazaki, Z.F. Liu, D. Meng, T. Kikuta, J. Alloys Compd. 488 (2009) [223] M. Joulazadeh, A.H. Navarchian, Synth. Met. 210 (2015) 404.
1339 L21. [224] T.T. Jiang, Z.J. Wang, Z.Y. Li, W. Wang, X.R. Xu, X.C. Liu, J.F. Wang, C. Wang, J.
[188] Y.H. Lin, Y.C. Hsueh, P.S. Lee, C.C. Wang, J.M. Wu, T.P. Perng, H.C. Shih, J. Mater. Mater. Chem. C 1 (2013) 3017. 1364
1340 Chem. 21 (2011) 10552. [225] Y. Li, H.T. Ban, M.J. Yang, Sens. Actuators B: Chem. 224 (2016) 449.
[189] S.W. Choi, S.S. Kim, J. Mater. Res. 27 (2012) 1688. [226] L.N. Geng, S.R. Wang, P. Li, Y.Q. Zhao, S.M. Zhang, S.H. Wu, Chin. J. Inorg. Chem.
[190] B.H. Jang, O. Landau, S.J. Choi, J.W. Shin, A. Rothschild, I.D. Kim, Sens. 21 (2005) 977. 1365
1341 Actuators B: Chem. 188 (2013) 156. [227] L.N. Geng, Y.Q. Zhao, X.L. Huang, S.R. Wang, S.M. Zhang, S.H. Wu, Sens.
[191] D. Haridas, A. Chowdhuri, K. Sreenivas, V. Gupta, Sens. Actuators B: Chem. 153 Actuators B: Chem. 120 (2007) 568. 1366
1342 (2011) 152. [228] F.H. Kong, Y. Wang, J. Zhang, H.J. Xia, B.L. Zhu, Y.M. Wang, S.R. Wang, S.H. Wu,
[192] I. Kocemba, J. Rynkowski, Sens. Actuators B: Chem. 155 (2011) 659. Mater. Sci. Eng. B 150 (2008) 6. 1367
[193] M. Hubner, D. Koziej, M. Bauer, N. Barsan, K. Kvashnian, M.D. Rossell, U. [229] M. Xu, J. Zhang, S. Wang, X. Guo, H. Xia, Y. Wang, S. Zhang, W. Huang, S. Wu,
1343 Weimar, J.D. Grunwaldt, Angew. Chem. Int. Ed. 50 (2011) 2841. Sens. Actuators B: Chem. 146 (2010) 8. 1368
[194] S.S. Li, Z.S. Lu, Z.X. Yang, X.L. Chu, Sens. Actuators B: Chem. 202 (2014) 83. [230] S.S. Barkade, D.V. Pinjari, U.T. Nakate, A.K. Singh, P.R. Gogate, J.B. Naik, S.H.
[195] M. Hubner, D. Koziej, J.D. Grunwaldt, U. Weimar, N. Barsan, Phys. Chem. Sonawane, A.B. Pandit, Chem. Eng. Process. 74 (2013) 115. 1369
1344 Chem. Phys. 14 (2012) 13249. [231] B.B. Wang, X.X. Fu, F. Liu, S.L. Shi, J.P. Cheng, X.B. Zhang, J. Alloys Compd. 587
[196] J. Zhang, X.H. Liu, S.H. Wu, M.J. Xu, X.Z. Guo, S.R. Wang, J. Mater. Chem. 20 (2014) 82. 1370
1345 (2010) 6453. [232] H.J. Sharma, N.D. Sonwane, S.B. Kondawar, Fiber Polym. 16 (2015) 1527.
[197] S.H. Park, S.Y. An, S.Y. Park, C.H. Jin, W.I. Lee, C.M. Lee, Appl. Phys. A 110 (2013) [233] A.S. Komolov, E.F. Lazneva, N.B. Gerasimova, M.V. Zimina, P. Si, Yu. A. Panina,
1346 471. Phys. Solid State 57 (2015) 2550. 1371
[198] P. Manjula, S. Arunkumar, S.V. Manorama, Sens. Actuators B: Chem. 152 [234] K.D. Schierbaum, U. Weimar, W. Gopel, Sens. Actuators B: Chem. 7 (1992) 709.
1347 (2011) 168. [235] F.I. Bohrer, C.N. Colesniuc, J. Park, M.E. Ruidiaz, I.K. Schuller, A.C. Kummel, W.
[199] A. Katoch, J.H. Byun, S.W. Choi, S.S. Kim, Sens. Actuators B: Chem. 202 (2014) C. Trogler, J. Am. Chem. Soc. 131 (2009) 478. 1372
1348 38. [236] H.H. Yan, P. Song, S. Zhang, Z.X. Yang, Q. Wang, RSC Adv. 5 (2015) 79593.
[200] J. Guo, J. Zhang, H.B. Gong, D.X. Ju, B.Q. Cao, Sens. Actuators B: Chem. 226 [237] S.M. Cui, Z.H. Wen, X.K. Huang, J.B. Chang, J.H. Chen, Small 11 (2015) 2305.
1349 (2016) 266. [238] B. Wang, Y.D. Wang, Y.P. Lei, S. Xie, N. Wu, Y.Z. Gou, C. Han, Q. Shi, D. Fang, J.
[201] Y.G. Li, L. Qiao, D. Yan, L.L. Wang, Y. Zeng, H.B. Yang, J. Alloys Compd. 586 Mater. Chem. C 4 (2016) 295. 1373
1350 (2014) 399. [239] A. Karakuscu, A. Ponzoni, E. Comini, G. Sberveglieri, C. Vakifahmetoglu, Int. J.
[202] Y.F. Bing, Y. Zeng, S.R. Feng, L. Qiao, Y.Z. Wang, W.T. Zheng, Sens. Actuators B: Appl. Ceram. Technol. 11 (2014) 851. 1374
1351 Chem. 227 (2016) 362. [240] C.C. Lian, Q.Z. Xue, Z.D. Han, H.P. Lu, F.J. Xia, Z.F. Yan, L.J. Deng, Sens. Actuators
[203] Y. Li, N. Chopra, Phys. Chem. Chem. Phys. 17 (2015) 12881. B: Chem. 227 (2016) 438. 1375

Please cite this article in press as: J.P. Cheng, et al., A review of recent developments in tin dioxide composites for gas sensing application, J. Ind.
Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.08.008

You might also like