You are on page 1of 15

Author’s Accepted Manuscript

Characterization and Photocatalytic Properties of


SnO2–TiO2 Nanocomposites Prepared Through
Gaseous Detonation Method

Linsong Wu, Honghao Yan, Xiaojie Li, Xiaohong


Wang
www.elsevier.com/locate/ceri

PII: S0272-8842(16)31889-2
DOI: http://dx.doi.org/10.1016/j.ceramint.2016.10.124
Reference: CERI14003
To appear in: Ceramics International
Received date: 15 September 2016
Revised date: 18 October 2016
Accepted date: 19 October 2016
Cite this article as: Linsong Wu, Honghao Yan, Xiaojie Li and Xiaohong Wang,
Characterization and Photocatalytic Properties of SnO2–TiO2 Nanocomposites
Prepared Through Gaseous Detonation Method, Ceramics International,
http://dx.doi.org/10.1016/j.ceramint.2016.10.124
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Characterization and Photocatalytic Properties of SnO2–

TiO2 Nanocomposites Prepared Through Gaseous

Detonation Method

Linsong Wu, Honghao Yan*, Xiaojie Li, Xiaohong Wang

Department of Engineering Mechanics, Dalian University of Technology, 2 Linguo Street, 116024

Dalian, PR China

*Corresponding author. Tel: +86-0411-84708397. dlutpaper@163.com

Abstract

SnO2-TiO2 nanocomposites were prepared by gaseous detonation method. The effects of SnO2

content on morphology, properties and photocatalytic properties of SnO2-TiO2 nanocomposites

were explored. Structure, components, grain size, morphology and optical properties of samples

were characterized by X-ray diffractometer (XRD), Transmission electron microscope (TEM),

Energy Dispersive Spectrometer (EDS), UV–vis spectroscopy, and Infrared spectroscopy (FIIR).

Photocatalytic properties were measured by methyl orange solution removal experiments under UV

irradiation at an initial concentration of 10 mg/L. Results demonstrated that SnO2 content is

negatively related to the grain size of samples. The TiO2 content in the rutile phase is positively

correlated with the photocatalytic properties of samples. The samples were anatase-rutile mixed

crystal when the molar ratio of Sn: Ti is 1:4. The grain size of mixed crystal in anatase-rutile phase

ranged between 8 nm and 40 nm and the energy gap was 3.03 eV. This phase demonstrated the

highest photocatalytic properties. The methyl orange solution was degraded completely at 25 min

at a constant reaction rate of 0.146 min-1. The rate of photocatalytic reaction was 46% higher than

that of the prepared TiO2 and about 13% higher than that of P25.

Keywords: Nanocomposites; TiO2; Optical properties;

1. Introduction
TiO2 is one of the most widely used photocatalyst. Studies on its applications in

photocatalytic environmental cleaning (organic pollutant degradation in air and water) [1-3] and new

energy sources (preparation of H2, hydrocarbons, and solar cells) [4-6] have achieved rapid

development in the past three decades. This development is attributed to the advantages of TiO2,

including rich reserves, low cost, high chemical stability and photochemical stability, strong

photocatalytic oxidation ability, non-toxic, and good environmental compatibility [7]. However, the

bandwidth (3.0 to 3.2 eV) limits of TiO2 in UV-light region and the over a high recombination rate

of photoelectron-hole destroy the photocatalytic properties of TiO2. Existing studies found that

semiconductor composites could increase the photocatalytic properties of products [8, 9]. Given the

similar crystal structure (rutile phase, tetragonal system) of SnO2 and TiO2, electronic properties

and ionic radius SnO2-TiO2 composite easily form solid solutions. This property enables the band

gap of SnO2 to match TiO2 better thereby decreasing band-gap energy, reducing the recombination

of photo electrons and holes, and increasing photocatalytic properties. Therefore, the research and

preparation of SnO2-TiO2 composites have attracted increasing attention [10-12].

Existing methods for preparing SnO2-TiO2 composites mainly include sol-gel method [13, 14],

hydrothermal method [15], co-electrodeposition method[12], and electrospinning method[16]. The

gaseous detonation method has not been reported in the preparation of SnO2-TiO2 composites.

Gaseous detonation method is a new synthetic technique that prepares nano oxides by detonating

the mixed gas of combustible gas and precursor. It has the characteristics of short production period,

high product purity, small grain size, and ease of operation [17]. Existing studies prepared a number

of nanomaterials, including TiO2, SnO2, and SiO2 [17, 18]; the morphologies of these materials were

characterized and analyzed using gaseous detonation method. However, the properties of these

nanomaterials have not been analyzed deeply. The present study prepares SnO2-TiO2

nanocomposites using gaseous detonation method. These materials were characterized by XRD,

TEM, EDX, UV-vis, and FIIR. The photocatalytic properties of samples with different molar ratios

(Sn: Ti) were analyzed through a degradation experiment of methyl orange solution.

2. Experiment

2.1 Preparation
The experiment used a self-made gaseous temperature-control detonation tube17; this tube is a

closed cylinder container composed of titanium tube with 100 mm inner diameter and 1100 mm

length and a titanium flange plate. The volume of the gaseous temperature-control detonation tube

was 8.6 L. The experiment equipment has a detonation point, a feed port, vacuum meter, and

temperature control system. The precursor was a mixture of certain proportions of TiCl4 (AR,

Sinopharm Chemical Reagent Co., Ltd, China) and SnCl4 (AR, Sinopharm Chemical Reagent Co.,

Ltd, China).

The gaseous detonation tube was heated to 403 K through the temperature control system.

The detonation tube was evacuated by a vacuum pump. Precursor measuring 2 ml was then injected

into the detonation tube and the gasification degree was determined by observing changes of the

pointer of the vacuum meter. Hydrogen was injected. The dosage of hydrogen was calculated

according to numerical changes in the pressure gauge. Oxygen was then injected into the tube until

the pointer of the vacuum meter reaches zero. The detonation tube was put on static for 3 to 5 min.

After completely mixing the gases, they were detonated and white powder was collected 20 min

later. Finally, the products were dried and deacidified. Table 1 lists the experimental parameters.

2.2 Photocatalysis

The photocatalytic activity of products was determined using a 300 W mercury lamp by

degrading methyl orange solution under UV light. The initial concentration of methyl orange dye

was 10 mg/L and the dosage of TiO2 was 100 mg per 400 ml of methyl orange solution. Prior to

photocatalytic reaction, the solution was stirred in the dark for 30 min to achieve adsorption–

desorption equilibrium. Solution samples were collected every 5 min and their absorbency was

tested by ultraviolet spectrophotometer after centrifugation.

2.3 Characterization

The structures and morphologies of the samples were analyzed by XRD (D/MAX 2400 X-ray

powder diffractometer, Cu target (Kα, λ=0.15406 nm), 40 KV tube voltage, 30 A tube current, 10°

to 90° scanning angle and 8°/min scanning rate) and TEM (TF30 field emission transmission

electron microscope). Optical properties of samples were analyzed by UV-vis and FTIR (UV-550
ultraviolet spectrophotometer and IR Affinity-1 infrared spectrometer). The concentration of

methyl orange solution was tested by a 721 spectrophotometer.

3. Results and discussion

3.1 XRD analysis

Fig.1a shows the XRD patterns of samples. Diffraction peaks at 2θ=25.3°, 37.9°, 48.2°, 54.1°

and 55.2° were anatase-phase TiO2 (JCPDS No.73-1764). Diffraction peaks at 2θ=27.4°, 35.7°,

40.9°, 53.9°, 56.2° and 68.4° were rutile-phase TiO2 (JCPDS No.76-0321). Diffraction peaks at

2θ=26.6°, 33.9°, 38.0°, 51.8° and 54.8° were rutile-phase SnO2 (JCPDS No.72-1147). A

comparison between the XRD patterns and standard card shows that Sample 1 was composed of

anatase-phase TiO2 and few rutile-phase TiO2. Sample 2 consisted of anatase-phase TiO2 and few

rutile-phase TiO2. Sample 3 was composed of the mixed crystal of anatase-phase and rutile-phase

TiO2 without the diffraction peak of rutile-phase SnO2. Sample 4 was composed of the mixed

crystal of anatase-phase TiO2 and rutile-phase SnO2 without diffraction peak of rutile-phase TiO2.

Sample 5 was composed of rutile-phase SnO2 accompanied by anatase-phase TiO2. Diffraction

peak was observed in rutile-phase TiO2. Fig.1a shows that increased SnO2 content gradually

decreased the diffraction peaks of the anatase phase, whereas the diffraction peaks of the rutile

phase increased gradually. This finding indicates that SnO2 content could significantly influence the

crystalline phase of TiO2 and excessive SnO2 could translate TiO2 from anatase into rutile phase.

Diffraction peaks were sharp and strong, which indicate high crystallinity of the samples.

Fig.1b shows the magnification of diffraction peak around 27.4°. The diffraction peaks of

rutile phase of Samples 3, 4 and 5 shifted to some extent, which is caused by the lattice parameter

changes of the samples. Table 2 shows the lattice parameters of the samples. Mixing Sn4+ with TiO2

crystal replaced Ti4+ in the rutile-phase TiO2 and formed solid solution[14]; thus, rutile-phase TiO2

and rutile-phase SnO2 did not coexist in the XRD pattern. Table 2 shows the calculated results of

anatase-phase and rutile-phase contents in TiO2. Changes in SnO2 content were the major cause of

the rapid transformation of samples from anatase to rutile phase[20, 21]. Sn4+ can easily replace Ti4+ in

lattice and form solid solution that disperses evenly in the structure because of the similar

structures of rutile-phase SnO2 and rutile-phase TiO2 and the similar radius of Sn4+ and Ti4+ (0.072
nm and 0.075 nm). This transformation facilitates the generation of rutile phase.

The mean grain size of samples could be calculated by the Scherrer formula: D=Kλ/βcosθ,

where D is the average thickness of grains in the normal of crystal face that corresponds to the

diffraction peak [the average grain size (nm)], λ is X-ray wavelength (0.154 nm), β is the half-peak

breadth of the corresponding diffraction peak (radian), θ is the Bragg angle of diffraction peak, and

K is a constant (0.89). Table 2 shows the calculated results. Increased SnO2 content can effectively

reduce the average grain size of samples. According to El-Maghrapy et al., [13] the existence of

SnO2 inhibits the growth of TiO2 particles.

3.2 TEM analysis

Fig. 2 shows the TEM images of samples. The grain size of samples varied between 8 nm and

40 nm. Product particles were basically spherical and clustered. The mean grain size of samples

was calculated by particle size in TEM images. Table 2 shows the results. The TEM images show

that the mean grain size of all samples was about 20 nm. The mean grain size decreased gradually

from 24.5 nm to 17.7 nm with the increase of SnO2 content. This increase was in accordance with

XRD analysis with a slight difference in granularity. This finding could be attributed to the mean

grain size calculated by TEM images, which was only a part of samples. Another reason was the

small particles that did not count in the statistics because of obvious clustering phenomenon.

Compared with P25, the composites prepared by gaseous detonation method have smaller and

meaner grain size.

3.3 UV-vis spectral analysis

The energy band structure of semiconductor is an important factor that determines its

photocatalytic properties. Fig. 3 shows the UV-vis absorption spectra of the samples. All samples

have evident absorption peaks in the UV region. Compared with Sample 1, the absorption spectra

of Samples 3, 4 and 5 exhibited slight blue shift, which may be caused by decreasing grain size[22].

The energy gap of samples was calculated using Eg=1240/λg, where λg was the absorption

wavelength threshold determined by the intercept method. The following results were obtained:

3.07 eV, 2.9 eV, 3.0 eV, 3.03 eV and 3.0 eV. The energy gap of all samples was about 3.0, which

indicates that the existence of SnO2 could significantly affect the energy gap of samples.
3.4 Infrared spectroscopic analysis

Fig. 4 shows the infrared spectroscopy (IR) of the samples. Stretching vibration at 3600 cm-1 and

bending vibration at 1620 cm-1 represent the radical and physical hydroxyl absorbed by water of the

TiO2 surface[23, 24]. The absorbed water molecules and free hydroxyl radicals are common

characteristics of semiconductor oxides and the basic conditions of photocatalysis. Stretching

vibrations within 400 cm-1 to 700 cm-1 represent Ti-O-Ti, Sn-O-Sn and Ti-O-Sn bonds in TiO2 and

SnO2[21,25].

3.5 Photocatalysis analysis

Fig.5a shows the relation curves between the degradation of methyl orange solution and time.

Samples have strong photocatalytic properties and degraded completely after 25 min. compared

with pure TiO2, SnO2–TiO2 composites with different proportions can increase photocatalytic

activity. The photocatalytic properties of samples increased gradually with the increase of SnO2

content thereby reaching the peak at Sn:Ti=1:4 (Sample 4). The photocatalytic properties of

Sample 4 were significantly higher than those of commercial TiO 2 (P25). Photocatalytic

properties gradually declined with the continuous increase of SnO2 content. Fig.5b shows that the

photocatalytic degradation of methyl orange solution conforms to the following first-order

reaction:

ln  C / C0   kt

where C0 is the initial concentration of methyl orange solution (mg/L), C is the current

concentration of methyl orange solution (mg/L), t is illumination time (min), and k is apparent

reaction rate constant (min-1). After fitting, k values were 0.100 min-1, 0.114 min-1, 0.144 min-1,

0.146 min-1 and 0.119 min-1. SnO2-TiO2 composites with different proportions had relatively

higher photocatalytic rates than pure TiO2. The photocatalytic rates of Sample 2, 3 (Sn: Ti=1:7)

and 4 (Sn:Ti=1:4) were about 14%, 44% and 46% higher than that of pure TiO 2. The

photocatalytic rates of Sample 4 was 13% higher than that of P25 (k=0.129 min-1). The

photocatalytic activity of Sample 5 decreased significantly compared with that of Sample 4. This

finding is caused by the sharp reduction of anatase-phase content in Sample 5. However, the

photocatalytic activity of Sample 4 remains 19% higher than that of pure TiO2.
When TiO2 receives photon radiation with energy larger than or equal to Eg, electrons on

the valence band were motivated and transited onto the conduction band and holes were left on

the valence band thereby forming e--h+ pairs. Photo electrons and holes can transfer to TiO 2

surface through diffusion to facilitate redox reaction with substances adsorbed on the TiO 2

surface or initiate composite reaction on the TiO2 particle surface or in the TiO2 particles.

Electrons and holes react with H 2O, O2 and OH- on the TiO2 surface and generate hydroxyl

radicals (•OH) with strong oxidizing property, superoxide radicals (O 2•-) and H2O2. This process

facilitates the decomposition of organic matters on the TiO2 surface. This process explains the

strong photocatalysis of TiO2. The energy band structure of SnO2 is the main cause of the higher

photocatalytic activity of SnO2-TiO2 composites than pure TiO2 [27, 28]. The energy gap of SnO2

(3.6 eV) is larger than that of TiO2 (3.2eV), and its conduction band potential and valence band

potential are lower than those of TiO 2. Some scholars suggest that the band structure of energy

can form heterojunction after the recombination of SnO 2 and TiO2 [8, 28, 29]. Electrons transit from

TiO2 surface to the conduction band of SnO2, whereas holes will transit from the valence band of

SnO2 to the valence band of TiO2. The migration of photon-induced carrier reduces the

recombination of TiO2 and SnO2, and prolongs the service life of electrons and holes. This

mechanism explains the higher photocatalytic activity of SnO 2-TiO2 nanocomposites than that of

TiO2.

4. Conclusions

SnO2-TiO2 nanocomposites with different Sn/Ti are prepared through gaseous detonation

method. Results show that SnO2-TiO2 nanocomposites can enable SnO2 to facilitate the production

of rutile phase and decrease the mean grain size of products. The photocatalytic properties of

products significantly increased with the increasing of SnO2 content. The photocatalytic activity of

samples peaked when Sn: Ti=1: 4, increasing Sn: Ti up 1:1.5 is accompanied by reduction in

photocatalytic reactivity of the samples due to the loss of titania anatase phase. High photocatalytic

properties are attributed to the synergistic effect of SnO2–TiO2 composites. Heterojunction is

formed after combination of SnO2 and TiO2, which reduces the recombination of electrons and

holes thereby increasing the photocatalytic properties of samples. This study proves the feasibility
of gaseous detonation method in preparing nanocomposite photocatalyst and lays the foundation

for future preparation of other composite photocatalyst using the gaseous detonation method.

ACKNOWLEDGMENT

This project was financially supported by the National Natural Science Foundation of China

(Nos. 10872044, 10972051, 10902023 and 11672068).

References

[1] T. Ochiai, A. Fujishima, Photoelectrochemical properties of TiO2 photocatalyst and its applications for
environmental purification, Journal of Photochemistry and Photobiology C: Photochemistry Reviews, 13
(2012) 247-262.
[2] A. Fujishima, X. Zhang, Titanium dioxide photocatalysis: present situation and future approaches,
Comptes Rendus Chimie, 9 (2006) 750-760.
[3] Y. Paz, Application of TiO2 photocatalysis for air treatment: Patents’ overview, Applied Catalysis B:
Environmental, 99 (2010) 448-460.
[4] S. Horikoshi, H. Hidaka, N. Serpone, Environmental remediation by an integrated
microwave/UV-illumination technique, Journal of Photochemistry and Photobiology A: Chemistry, 159 (2003)
289-300.
[5] S. Zhou, Y. Liu, J. Li, Y. Wang, G. Jiang, Z. Zhao, D. Wang, A. Duan, J. Liu, Y. Wei, Facile in situ synthesis of
graphitic carbon nitride (g-C3N4)-N-TiO2 heterojunction as an efficient photocatalyst for the selective
photoreduction of CO2 to CO, Applied Catalysis B: Environmental, 158-159 (2014) 20-29.
[6] Y. Ma, X. Wang, Y. Jia, X. Chen, H. Han, C. Li, Titanium dioxide-based nanomaterials for photocatalytic fuel
generations, Chemical reviews, 114 (2014) 9987-10043.
[7] A. Fujishima, X.T. Zhang, D.A. Tryk, TiO2 photocatalysis and related surface phenomena, Surf. Sci. Rep., 63
(2008) 515-582.
[8] U.I. Gaya, A.H. Abdullah, Heterogeneous photocatalytic degradation of organic contaminants over
titanium dioxide: A review of fundamentals, progress and problems, Journal of Photochemistry and
Photobiology C: Photochemistry Reviews, 9 (2008) 1-12.
[9] H. Su, C. Peng, J. Wu, Effects of composition on the phase transformation behavior of anatase TiO2 in
photocatalytic ceramics, Materials Research Bulletin, 48 (2013) 4963-4966.
[10] A. Enesca, L. Isac, L. Andronic, D. Perniu, A. Duta, Tuning SnO2–TiO2 tandem systems for dyes
mineralization, Applied Catalysis B: Environmental, 147 (2014) 175-184.
[11] V.R. de Mendonça, O.F. Lopes, R.P. Fregonesi, T.R. Giraldi, C. Ribeiro, TiO2-SnO2 heterostructures
applied to dye photodegradation: The relationship between variables of synthesis and photocatalytic
performance, Applied Surface Science, 298 (2014) 182-191.
[12] M. Huang, S. Yu, B. Li, D. Lihui, F. Zhang, M. Fan, L. Wang, J. Yu, C. Deng, Influence of preparation
methods on the structure and catalytic performance of SnO2-doped TiO2 photocatalysts, Ceramics
International, 40 (2014) 13305-13312.
[13] E.M. El-Maghraby, Effect of Sn ratio on the photocatalytic degradation of methylene blue and soot of
ink by TiO2–SnO2 nanostructured thin films, Physica B: Condensed Matter, 405 (2010) 2385-2389.
[14] S. Guo, D. Li, Y. Zhang, Y. Zhang, X. Zhou, Fabrication of a Novel SnO2 Photonic Crystal Sensitized by CdS
Quantum Dots and Its Enhanced Photocatalysis under Visible Light Irradiation, Electrochimica Acta, 121
(2014) 352-360.
[15] E. Arpac, F. Sayilkan, M. Asilturk, P. Tatar, N. Kiraz, H. Sayilkan, Photocatalytic performance of Sn-doped
and undoped TiO2 nanostructured thin films under UV and vis-lights, Journal of hazardous materials, 140
(2007) 69-74.
[16] H. Shi, M. Zhou, D. Song, X. Pan, J. Fu, J. Zhou, S. Ma, T. Wang, Highly porous SnO2/TiO2 electrospun
nanofibers with high photocatalytic activities, Ceramics International, 40 (2014) 10383-10393.
[17] H.H. Yan, L.S. Wu, X.J. Li, X.H. Wang, Detonation Synthesis of SnO2 Nanoparticles in Gas Phase, Rare
Metal Mat. Eng., 42 (2013) 1325-1327.
[18] H. Yan, L. Wu, X. Li, X. Wang, S. Wang, Influences of relative amount of substance of precursor on
nanoSiO_2 particles prepared by oxyhydrogen gaseous deflagration, Explosion and Shock Waves, 32 (2012)
581-584.
[19] A. Kusior, J. Klich-Kafel, A. Trenczek-Zajac, K. Swierczek, M. Radecka, K. Zakrzewska, TiO2–SnO2
nanomaterials for gas sensing and photocatalysis, Journal of the European Ceramic Society, 33 (2013)
2285-2290.
[20] R. Sasikala, A. Shirole, V. Sudarsan, T. Sakuntala, C. Sudakar, R. Naik, S.R. Bharadwaj, Highly dispersed
phase of SnO2 on TiO2 nanoparticles synthesized by polyol-mediated route: Photocatalytic activity for
hydrogen generation, International Journal of Hydrogen Energy, 34 (2009) 3621-3630.
[21] M.F. Abdel-Messih, M.A. Ahmed, A.S. El-Sayed, Photocatalytic decolorization of Rhodamine B dye using
novel mesoporous SnO2–TiO2 nano mixed oxides prepared by sol–gel method, Journal of Photochemistry
and Photobiology A: Chemistry, 260 (2013) 1-8.
[22] H. Lin, C. Huang, W. Li, C. Ni, S. Shah, Y. Tseng, Size dependency of nanocrystalline TiO2 on its optical
property and photocatalytic reactivity exemplified by 2-chlorophenol, Applied Catalysis B: Environmental, 68
(2006) 1-11.
[23] Y.H. Tseng, C.S. Kuo, C.H. Huang, Y.Y. Li, P.W. Chou, C.L. Cheng, M.S. Wong, Visible-light-responsive
nano-TiO(2) with mixed crystal lattice and its photocatalytic activity, Nanotechnology, 17 (2006) 2490-2497.
[24] X. Wang, L. Sø, R. Su, S. Wendt, P. Hald, A. Mamakhel, C. Yang, Y. Huang, B.B. Iversen, F. Besenbacher,
The influence of crystallite size and crystallinity of anatase nanoparticles on the photo-degradation of
phenol, Journal of Catalysis, 310 (2014) 100-108.
[25] L.-C. Chen, C.-M. Huang, M.-C. Hsiao, F.-R. Tsai, Mixture design optimization of the composition of S, C,
SnO2-codoped TiO2 for degradation of phenol under visible light, Chemical Engineering Journal, 165 (2010)
482-489.
[26] X. Li, R. Xiong, G. Wei, Preparation and photocatalytic activity of nanoglued Sn-doped TiO2, Journal of
hazardous materials, 164 (2009) 587-591.
[27] L.R. Hou, C.Z. Yuan, Y. Peng, Synthesis and photocatalytic property of SnO2/TiO2 nanotubes composites,
Journal of hazardous materials, 139 (2007) 310-315.
[28] S. Valencia, F. Cataño, L. Rios, G. Restrepo, J. Marín, A new kinetic model for heterogeneous
photocatalysis with titanium dioxide: Case of non-specific adsorption considering back reaction, Applied
Catalysis B: Environmental, 104 (2011) 300-304.
[29] Y. Qu, X. Duan, Progress, challenge and perspective of heterogeneous photocatalysts, Chemical Society
reviews, 42 (2013) 2568-2580.
Rutile SnO2
a JCPDS No.72-1147

Sample 5

Sample 4
Intensity(a.u.)

Sample 3

Sample 2

Sample 1

Rutile TiO2
JCPDS No.76-0321

Anatase TiO2
JCPDS No.73-1764

20 30 40 50 60 70 80
2/degree
TiO2(Rutile )
SnO2(Rutile)

Sample 5
Intensity(a.u.)

Sample 4

Sample 3

26 27 28 29 30
2/degree

Fig. 1 (a) XRD pattern of samples; (b) Magnification of diffraction peak around 27.4°
Fig. 2 TEM images of the samples, (a) Sample 1; (b) Sample 2; (c) Sample 3; (d) Sample 4; (e)

Sample 5; (f) P25

1.0

0.8 Sample 3
Absorption(a.u.)

Sample 5 Sample 2
0.6
Sample 4
Sample 1

0.4

0.2

0.0
200 250 300 350 400 450 500
Wavelength(nm)

Fig. 3 UV–vis absorption spectrum of the samples


Sample 5

Sample 4

Transmittance(a.u.) Sample 3

Sample 2

Sample 1

500 1000 1500 2000 2500 3000 3500 4000 4500


-1
Wavemumber(cm )

Fig. 4 IR spectra of the samples

1.0
Sample 1
Sample 2
Sample 3
0.8
Sample 4
Sample 5
P25
0.6
Ct/C0

0.4

0.2

a
0.0
0 5 10 15 20 25
Time(min)
4
Sample 1 k=0.100
Sample 2 k=0.114
Sample 3 k=0.144
Sample 4 k=0.146
3 Sample 5 k=0.119
P25 k=0.129

-ln(Ct/C0)

1
b

0 5 10 15 20 25
Time(min)

Fig.5 (a) Variation of photocatalytic degradation of methyl orange dye with time of irradiation;

(b) First-order kinetics of the degradation of methyl orange dye with irradiation time

Table 1 Experiment parameters

Sample Dosage of SnCl4 Theoretical Volume Volume Initial

and TiCl4 (ml) molar ratio fraction of H2 fraction of O2 pressure

(Sn: Ti) (MPa)

1 0 and 2.0 -- 0.1 0.8 0.1

2 0.1 and 1.9 1 : 20 0.1 0.8 0.1

3 0.3 and 1.7 1:6 0.1 0.8 0.1

4 0.5 and 1.5 1 : 3.2 0.1 0.8 0.1

5 0.7 and 1.3 1: 1.7 0.1 0.8 0.1

Table 2 Particle size, texture, and component parameters of SnO2-TiO2 nanocomposites


Sample 1 2 3 4 5

Crystallite Anatase 18.4 16.9 20.7 18.7 16.6

sizea (nm) Rutile 29.2 33.4 14.9 12.1 12.4

Composition(A: R) 85:15 94:6 52:48 51:49 30:70

Lattice parameter(a) -- -- 4.624 4.734 4.724

Lattice parameter(c) -- -- 2.598 3.158 3.188

Crystallite sizeb (nm) 24.5 20.3 20.7 19.4 17.7

Actual molar ratioc -- 1:40 1:7 1:4 1:1.5

(Sn:Ti)

a. Calculated results of the Scherrer formula; b. Results of TEM; c. Results of EDX analysis

You might also like