+ All Categories
Home > Documents > art%3A10.1007%2Fs10854-015-3716-6

art%3A10.1007%2Fs10854-015-3716-6

Date post: 25-Feb-2018
Category:
Upload: beatriz-brachetti
View: 216 times
Download: 0 times
Share this document with a friend

of 10

Transcript
  • 7/25/2019 art%3A10.1007%2Fs10854-015-3716-6

    1/10

    One-step hydrothermal synthesis of tin dioxide nanoparticlesand its photocatalytic degradation of methylene blue

    Morteza Vatanparast1 Mohammad Taghi Taghizadeh1

    Received: 28 July 2015 / Accepted: 28 August 2015/ Published online: 3 September 2015

    Springer Science+Business Media New York 2015

    Abstract A simple hydrothermal method was developed

    for the size-controlled synthesis of tin dioxide (SnO2)nanoparticles, using SnCl45H2O, ethylenediamine and

    hydrazine. The molar ratio of the reactants, temperature,

    reaction time and the surfactant were changed in order to

    investigate the effect of preparation parameters on the mor-

    phology and particle size of SnO2. According to scanning

    electron microscopy (SEM) results, ethylenediamine and

    hydrazine can control the particle growth and play an

    important role in formation of SnO2 nanostructures. The

    prepared SnO2nanoparticles were characterized extensively

    by means of X-ray diffraction, energy-dispersive X-ray

    spectroscopy, SEM, transmission electron microscopy,

    Fourier transform infrared spectroscopy and diffuse reflec-

    tance spectroscopy. Besides, the photocatalytic properties of

    as-synthesized SnO2 nanoparticles have been evaluated for

    the degradation of methylene blue under UV light irradiation.

    1 Introduction

    Semiconductor metal oxides have been widely investigated

    because of their potential applications as battery materials,

    photocatalysts, sensors, and optoelectronic devices [16].As

    an important n-type semiconductor metal oxide, tin dioxide

    (SnO2) with the wide direct energy band gap (Eg = 3.6 eVat

    300 K) has attracted vast interests due to its potential

    applications in many fields, such as photocatalysts [7], solar

    cells [8], gas sensor [9,10], optoelectronic devices [11,12]

    and electrode materials [13]. In general, in most of the

    applications, the performance significantly depends onmorphology of the structures [14]. In recent years, various

    morphologies of SnO2 nanostructures have been synthe-

    sized, such as nanoparticles [15,16], nanorods [17], nano-

    wires [18], nanotubes [19], nanocube [20], nanobelts [21],

    nanoribbons [22] and hollow spheres [23].

    Various approaches have been developed to prepare SnO2nanomaterials, such as microemulsion [24], template-based

    [25], solgel [26], controlled precipitation [27], microwave

    [28], sonochemical [29], solvothermal [30, 31], and

    hydrothermal [32, 33] methods. Among these synthesis

    methods, the hydrothermal method is a simple and inex-

    pensive technique for large scale production [34]. It is not

    only a low temperature preparation method, but also has the

    ability of controlling the size and shape of the final product

    [35, 36]. In the present work, we report the hydrothermal

    synthesis of SnO2 nanoparticles using SnCl45H2O,

    ethylenediamine, and hydrazine as precursors. The factors

    affecting the morphologies and particle sizes of SnO2crys-

    tals, including the molar ratio of the reactants, the

    hydrothermal reaction temperature and time, have been

    systematically investigated. Furthermore, the photocatalytic

    performance of the synthesized SnO2 nanoparticles was

    evaluated by monitoring the photodegradation of methylene

    blue (MB) under UV light irradiation.

    2 Experimental

    2.1 Preparation method

    All the chemicals reagents (purchased from Aldrich) used

    in our experiments were of analytical grade and were used

    as received without further purification.

    & Morteza Vatanparast

    [email protected]

    1 Department of Physical Chemistry, Faculty of Chemistry,

    University of Tabriz, Tabriz, Iran

    1 3

    J Mater Sci: Mater Electron (2016) 27:5463

    DOI 10.1007/s10854-015-3716-6

    http://crossmark.crossref.org/dialog/?doi=10.1007/s10854-015-3716-6&domain=pdfhttp://crossmark.crossref.org/dialog/?doi=10.1007/s10854-015-3716-6&domain=pdf
  • 7/25/2019 art%3A10.1007%2Fs10854-015-3716-6

    2/10

    In a typical experiment, 0.5 g of SnCl45H2O was dis-solved in 40 ml of distilled water. Then, 0.8 ml of

    ethylenediamine and 0.4 ml hydrazine hydrate were added

    drop wise to the solution under stirring and ultimately the

    clear solution was transferred into the Teflon lined auto-

    clave. The autoclave was sealed and maintained at 160 C

    for 5 h (sample 1), then cooled naturally to room temper-

    ature. The obtained white precipitate was centrifuged andwashed with distilled water and ethanol several times and

    was dried at 70 C for 3 h. Finally, it was calcined at

    400 C for 2 h. Some parameters including the molar ratio

    of the reactants, reaction temperature and time were

    changed for reaching the optimized condition. Reaction

    conditions are listed in Table 1.

    2.2 Characterization

    The X-ray diffraction (XRD) study was done by Italstracture

    X-ray diffractometer model MPD 3000 with a radiation

    source from Cu Ka at scan range of 10\ 2h\80. Energydispersive X-ray (EDX) spectrum and scanning electron

    micrographs (SEM) were obtained using a field-emission

    scanning electron microscopy (MIRA3 FEG-SEM, Tescan

    Brno, Czech Republic). Transmission electron microscope

    (TEM) images were obtained on a Philips EM208 trans-

    mission electron microscope with an accelerating voltage of

    100 kV. Fourier transform infrared (FT-IR) spectra were

    recorded on Tensor 27 (Bruker) spectrophotometer in KBr

    pellets. TheUVVis diffused reflectance spectra(DRS)were

    obtained from UVVis Scinco 4100 spectrometer. UVVis

    absorption spectra were recorded on a Perkin-Elmer Lambda

    2S UVVis spectrophotometer.

    2.3 Photocatalytic experiment

    The photocatalytic activities of the samples were deter-

    mined by the degradation of aqueous MB under UV light.

    In a typical photocatalytic test performed at room tem-

    perature, 20 mg of photocatalyst nanoparticles was added

    into 150 ml MB aqueous solution (5 mg/l), and then dis-

    persed by stirring in a dark condition for 30 min to

    establish adsorption/desorption equilibrium. Later, a series

    of UV lamps (4 9 15 W, Philips) was switched on a 20 cm

    distance over the suspension surface. Then, the changes of

    MB concentration were monitored on the basis of its UVvisible absorption peak at 664 nm.

    3 Results and discussion

    Figure1 shows the XRD patterns of the sample 1 before

    and after the calcination process. The XRD pattern of

    sample 1 before calcination appears to be amorphous-like

    with sharp peak at 31.74, 34.36, 36.22, 47.50, 56.58. It

    should be mentioned that we were not able to find any

    standard XRD data that match this XRD pattern. However,

    the diffraction pattern of sample 1 after calcination can beindexed to a pure tetragonal phase of SnO2 with space

    group P42/mnm (JCDPS No. 41-1445) including the lattice

    constants of a, b = 4.7382 and c = 3.1871. The crystallite

    size of the as-synthesized SnO2, D, was calculated from the

    major diffraction peaks using Scherrer equation [37]:

    D Kk

    b cos h 1

    where K is the so-called shape factor, which usually takes a

    value of about 0.9,k is the X-ray wavelength used in XRD,

    h is the Bragg angle; b is the breadth of the observed

    diffraction line at its half-intensity maximum. The esti-mated particle size of SnO2 by the Scherrer equation is

    about 10 nm. The purity of the nanostructures was inves-

    tigated by means of EDX analysis. EDX analysis of SnO2nanoparticles (sample 1) is shown in Fig. 2. Peaks asso-

    ciated with Sn and O are obviously observed. Au peaks are

    Table 1 Reaction conditions

    for SnO2 nanoparticles Sample no. Hydrazine (ml) Ethylenediamine (ml) Temperature (C) Time (h) Surfactant

    1 0.4 0.8 160 5

    2 0.4 0.05 160 5

    3 0.4 0.2 160 5

    4 0.4 1.6 160 5

    5 0 0 160 5

    6 NH3 160 5

    7 NaOH 160 5

    8 0.4 0.8 120 5

    9 0.4 0.8 200 5

    10 0.4 0.8 160 10

    11 0.4 0.8 160 15

    12 0.4 0.8 160 5 PEG

    J Mater Sci: Mater Electron (2016) 27:5463 55

    1 3

  • 7/25/2019 art%3A10.1007%2Fs10854-015-3716-6

    3/10

    detected because the samples were gold coated for con-

    duction. Moreover, no impurity peaks are observed, which

    indicates a high level of purity in the product.

    Figure3a, b show SEM images of obtained products

    before and after calcination of sample 1, respectively, in

    the presence of 0.8 ml hydrazine and 0.4 ml ethylenedi-

    amine. As can be seen, in both of them, spherical

    nanoparticles were obtained and after calcination, the

    particle size of them was increased (Fig. 3b). For investi-

    gation of ethylenediamine effect, several experiments with

    different concentration of ethylenediamine including 0.05,

    0.2 and 1.6 ml (sample 24, respectively) were performed.

    As shown in Fig. 4a, b, by decreasing the ethylenediamine

    concentration, products with finer and more homogenous

    particle size were obtained. By increasing the concentra-

    tion to 1.6 ml (sample 4), increasing and nonuniformity of

    particle size is obvious (Fig. 4c). Therefore, for obtaining

    finer and more uniform structure, an optimum amount ofethylenediamine is required. In the presence of ethylene-

    diamine and hydrazine, releasing of ions is controllable and

    so nucleation and growing mechanism in different con-

    centration can be controlled and by adjusting conditions,

    appropriate nanostructures can be achievable. Ethylenedi-

    amine is able to form a complex with metal ions. Thus,

    releasing of metal ions is strongly influenced by

    ethylenediamine concentration. On the other hand, the high

    amount of ethylenediamine concentration cant be useful

    because presence of high concentration of ethylenediamine

    leads to aggregation of nanoparticles due to carbon burning

    while annealing. Also, hydrazine hydrate compared toNaOH and NH3, can release hydroxide ion slower and

    more controllable. According to the above description,

    mechanism of SnO2 nanoparticles formation can be as

    follow:

    Coordination: Sn4 nen $ Sn en n 4

    2

    Hydrolyzation: NH2CH2CH2NH2 2H2O! NH3 CH2CH2NH

    3 2OH

    3

    N2H4 H2O ! N2H5 OH

    4

    Precipitation: Sn en n 4

    4OH annealing

    400 C 2 h SnO2

    2H2O nenen ethylenediamine5

    For direct studying the ethylenediamine and hydrazine

    effect and comparison with NaOH and NH3, three experi-

    ments were done (Fig.5) in absence of hydrazine and

    ethylenediamine (sample 5), in presence of NH3and NaOH

    (sample 6 and 7, respectively). As can be seen, in absence

    of hydrazine and ethylenediamine, larger particles were

    obtained in comparison with sample 1 prepared in the

    presence of 0.4 ml ethylenediamine and 0.8 ml hydrazine

    and in many parts, aggregation and agglomeration can be

    seen (Fig. 5a). In other words, with simple hydrolyze thatthe concentration of hydroxide ion is very low, fine and

    uniform nanoparticles of SnO2 cant be obtained. In low

    concentration, nucleation rate is low and as a result, large

    particles will be achieved. In the presence of NH3 and

    NaOH which the concentration of hydroxide ion is high,

    larger particle with less uniformity can be synthesized. At

    high concentration, the nucleation rate is high and too fine

    particles were obtained. Being the particle so fine leads to

    instability of particles and as a result, the particles can be

    Fig. 1 XRD patterns of sample 1 before (a) and after (b) calcination

    Fig. 2 EDX pattern of SnO2 nanoparticles (sample 1)

    56 J Mater Sci: Mater Electron (2016) 27:5463

    1 3

  • 7/25/2019 art%3A10.1007%2Fs10854-015-3716-6

    4/10

    aggregated together (Fig. 5b, c, respectively). Figure6a, b

    shows SEM images of samples 8 and 9 formed at 120 and

    200 C, respectively. As shown, by decreasing the tem-

    perature from 160 to 120 C and by increasing temperature

    from 160 to 200 C, the particle size got larger and smaller,

    respectively. Indeed, by increasing temperature, the parti-

    cles became finer and the finest and the most uniform

    particles with size of 1015 nm were obtained at 200 C.

    An increase of the reaction temperature always results in an

    increase of the rate of reaction; therefore at relatively high

    temperatures more SnO2nuclei will form before the growthprocess, which causes the formation of more particles with

    smaller diameters. For more investigation, TEM analysis

    was used to observe the morphology and size of the SnO2nanoparticles. Figure7 shows the TEM images of sample 9

    in different scale scopes. The sizes of nanoparticles

    obtained from the TEM images are in close agreement with

    the XRD diffraction patterns, which show sizes of

    1015 nm. For evaluating of time effect, in addition to 5 h

    (sample 1), two experiments were performed for 10 and

    15 h (samples 10 and 11, respectively). By increasing the

    time till 10 h, finer and more uniform structures can be

    synthesized (Fig.8a) compared to products obtained for

    5 h (sample 1). But, by increasing the time to 15 h (sample

    11), the particles got larger and the uniformity was

    decreased (Fig.8b). In other words, for achieving fine and

    uniform nanoparticles in hydrothermal method, the opti-

    mum amount of time is required and this optimum time in

    our study for formation of SnO2 is 10 h. For investigating

    of surfactant effect, an experiment was done in the pres-

    ence of PEG (sample 12) and related SEM image ofobtained product is shown in Fig. 9. As can be seen, in the

    presence of surfactant, finer and more uniform particles

    were prepared in comparison with products prepared in the

    absence of surfactant (sample 1). However, as noted ear-

    lier, by adjusting conditions, so fine and uniform

    nanoparticles can be obtained (see Fig. 6b). In other words,

    by adjusting ethylenediamine concentration, time and

    temperature, the surfactant can be eliminated from the

    reaction.

    Fig. 3 SEM images of sample 1 before (a) and after (b) calcination

    J Mater Sci: Mater Electron (2016) 27:5463 57

    1 3

  • 7/25/2019 art%3A10.1007%2Fs10854-015-3716-6

    5/10

    Figure10 represents FT-IR spectra of the sample 1

    before and after the calcination. The broad bands around

    3440 and 1650 cm-1 in Fig.10a correspond to stretching

    and bending vibrations of absorbed water molecules

    respectively [38], which can identify the presence of

    physisorbed water molecules linked to the SnO2 nanopar-

    ticles. The peaks at 2930 cm-1 are attributed to the CH2

    asymmetric vibrations [39]. The peak at 1550 cm-1 is

    associated with the bending mode of NH [40]. After

    calcination, these bands disappear because of the pyrolysis

    Fig. 4 SEM images of SnO2 nanoparticles a sample 2, b sample 3, c sample 4

    58 J Mater Sci: Mater Electron (2016) 27:5463

    1 3

  • 7/25/2019 art%3A10.1007%2Fs10854-015-3716-6

    6/10

    of the organic precursor. The peaks in the region

    8501350 cm-1 are assigned to hydroxyl bending mode of

    different types of surface hydroxyl groups [29]. The peaks

    at 665 and 550 cm-1 are assigned to the SnOSn

    antisymmetric vibrations [29], where after calcination

    broad band appear in this region. From the FT-IR results, it

    is concluded that further heat treatment is necessary to

    remove the residual water and organic precursor.

    The UVVis diffuse reflectance spectroscopy (DRS) of

    SnO2 nanoparticles (sample 1) are shown in Fig.11. The

    Fig. 5 SEM images of SnO2 nanoparticles a sample 5, b sample 6, c sample 7

    J Mater Sci: Mater Electron (2016) 27:5463 59

    1 3

  • 7/25/2019 art%3A10.1007%2Fs10854-015-3716-6

    7/10

    fundamental absorption edge in most semiconductors fol-

    lows the exponential law. Using the absorption data, the

    direct band gap (Eg) for SnO2 nanoparticles can be esti-

    mated by Taucs relationship (6):

    ahm Bhm Eg1=2 6

    where hm, B and a are incident photon energy, a constant

    related to the material and the absorption coefficient,

    Fig. 6 SEM images of SnO2 nanoparticles a sample 8, b sample 9

    Fig. 7 TEM images of sample 9 in different scale

    60 J Mater Sci: Mater Electron (2016) 27:5463

    1 3

  • 7/25/2019 art%3A10.1007%2Fs10854-015-3716-6

    8/10

    Fig. 8 SEM images of SnO2 nanoparticles a sample 10, b sample 11

    Fig. 9 SEM images of SnO2 nanoparticles (sample 12)

    J Mater Sci: Mater Electron (2016) 27:5463 61

    1 3

  • 7/25/2019 art%3A10.1007%2Fs10854-015-3716-6

    9/10

    respectively [41]. The Eg value is calculated as 3.8 eV

    (326 nm) for the SnO2nanoparticles (the sample 1), which

    shows a blue shift compared with SnO2 bulk (3.6 eV).

    To demonstrate the potential applicability of as-syn-

    thesized SnO2 nanoparticles, we investigated the photo-

    catalytic decomposition of MB as the test reaction.

    Moreover, the photocatalytic decomposition was per-

    formed under UV-light illumination, and the degradation

    rate for the decomposition of MB was estimated by

    observing the changes in absorbance (absorption intensity

    vs. irradiation time) obtained by UVVis spectra. Fig-

    ure12a shows the absorption spectrum of an aqueous

    solution of MB in the presence of the sample 1. As can be

    seen, the intensity of the characteristic absorption peak

    decreased monotonously with the increase in the exposure

    time. Figure12b indicated that in the absence of SnO2

    nanoparticles, nearly no change in the absorption spectrais observable, while in the presence of SnO2 nanoparti-

    cles, more than 90 % of MB was decolorized after

    120 min.

    4 Conclusions

    In summary, SnO2 nanoparticles were successfully syn-

    thesized by hydrothermal method using SnCl45H2O,

    ethylenediamine and hydrazine. Besides, the effects of

    molar ratio of the reactants, temperature, reaction time and

    surfactant on the morphology and particle size of thenanoparticles were investigated. The results show that the

    ethylenediamine concentration plays a key role in con-

    trolling the particle size of the final products. Hydrazine

    hydrate compared to NaOH and NH3, can release

    hydroxide ion slower and more controllable, which leads to

    formation of finer and more homogenous nanoparticles.

    SnO2 nanoparticles were characterized by XRD, EDX,

    SEM, TEM, FT-IR and DRS techniques. Moreover, the as-

    synthesized SnO2 nanoparticles showed the excellent

    Fig. 10 FT-IR spectrums of sample 1 before (a) and after

    (b) calcination

    Fig. 11 The UVVis diffusereflectance spectroscopy (DRS)

    of SnO2 nanoparticles (sample

    1). The insetshows the plots of

    (ahm)2 versus the energy of light

    (hm)

    62 J Mater Sci: Mater Electron (2016) 27:5463

    1 3

  • 7/25/2019 art%3A10.1007%2Fs10854-015-3716-6

    10/10

    photocatalytic ability for the degradation of MB as a model

    of pollutant.

    References

    1. J.A. Rodriguez, M. Fernandez-Garca, Synthesis, Properties, and

    Applications of Oxide Nanomaterials (Wiley, New York, 2007)

    2. M. Epifani, R. Daz, J. Arbiol, E. Comini, N. Sergent, T. Pagnier,

    P. Siciliano, G. Faglia, J.R. Morante, Adv. Funct. Mater. 16, 1488

    (2006)

    3. M. Vatanparast, M. Ranjbar, M. Ramezani, S.M. Hosseinpour-

    Mashkani, M. Mousavi-Kamazani, Superlattices Microstruct. 65,

    365 (2014)

    4. S.M. Hosseinpour-Mashkani, M. Ramezani, M. Vatanparast,

    Mater. Sci. Semicond. Process. 26, 112 (2014)

    5. T. Hammad, J. Salem, S. Kuhn, M. Draaz, R. Hempelmann, F.

    Kodeh, J. Mater. Sci. Mater. Electron. 26, 5495 (2015)

    6. L. Nejati-Moghadam, D. Ghanbari, M. Salavati-Niasari, A.

    Esmaeili-Bafghi-Karimabad, S. Gholamrezaei, J. Mater. Sci.

    Mater. Electron. 26, 6075 (2015)

    7. W.W. Wang, Y.J. Zhu, L.X. Yang, Adv. Funct. Mater. 17, 59

    (2007)

    8. Y.S. Kim, B.-K. Yu, D.-Y. Kim, W.B. Kim, Sol. Energy Mater.

    Sol. Cells 95, 2874 (2011)

    9. H. Zhang, W. Zeng, Y. Zhang, Y. Li, B. Miao, W. Chen, X. Peng,

    J. Mater. Sci. Mater. Electron. 25, 5006 (2014)

    10. D.D. Trung, N. Van Toan, P. Van Tong, N. Van Duy, N.D. Hoa,

    N.V. Hieu, Ceram. Int. 38, 6557 (2012)

    11. T. Tao, Q.-Y. Chen, H.-P. Hu, Y. Chen, Mater. Chem. Phys.126,

    128 (2011)

    12. G. Xi, J. Ye, Inorg. Chem. 49, 2302 (2010)

    13. Z. Wen, F. Zheng, K. Liu, Mater. Lett.68, 469 (2012)

    14. M. Wang, Y. Gao, L. Dai, C. Cao, X. Guo, J. Solid State Chem.

    189, 49 (2012)

    15. J. Zhang, S. Wang, Y. Wang, M. Xu, H. Xia, S. Zhang, W.

    Huang, X. Guo, S. Wu, Sens. Actuators, B 139, 369 (2009)

    16. A.E. Shalan, M. Rasly, I. Osama, M.M. Rashad, I.A. Ibrahim,

    Ceram. Int. 40, 11619 (2014)

    17. B. Cheng, J.M. Russell, W. Shi, L. Zhang, E.T. Samulski, J. Am.

    Chem. Soc. 126, 5972 (2004)

    18. Y. Wang, X. Jiang, Y. Xia, J. Am. Chem. Soc.125, 16176 (2003)

    19. L. Zhao, M. Yosef, M. Steinhart, P. Goring, H. Hofmeister, U.

    Gosele, S. Schlecht, Angew. Chem. Int. Ed. 45, 311 (2006)

    20. S. Cao, W. Zeng, H. Zhang, Y. Li, J. Mater. Sci. Mater. Electron.

    26, 2871 (2015)

    21. S.H. Sun, G.W. Meng, G.X. Zhang, T. Gao, B.Y. Geng, L.D.

    Zhang, J. Zuo, Chem. Phys. Lett. 376, 103 (2003)

    22. R. Zou, J. Hu, Z. Zhang, Z. Chen, M. Liao, CrystEngComm 13,

    2289 (2011)

    23. J. Zhang, S. Wang, Y. Wang, Y. Wang, B. Zhu, H. Xia, X. Guo,

    S. Zhang, W. Huang, S. Wu, Sens. Actuators, B 135, 610 (2009)

    24. J.X. Zhou, M.S. Zhang, J.M. Hong, J.L. Fang, Z. Yin, Appl. Phys.

    A 81, 177 (2005)

    25. J. Ye, H. Zhang, R. Yang, X. Li, L. Qi, Small6, 296 (2010)

    26. M. Ionita, G. Cappelletti, A. Minguzzi, S. Ardizzone, C. Bianchi,

    S. Rondinini, A. Vertova, J. Nanopart. Res. 8, 653 (2006)

    27. C.A. Ibarguen, A. Mosquera, R. Parra, M.S. Castro, J.E. Rodr-

    guez-Paez, Mater. Chem. Phys. 101, 433 (2007)

    28. A. Kar, S. Kundu, A. Patra, J. Phys. Chem. C 115, 118 (2011)

    29. D.N. Srivastava, S. Chappel, O. Palchik, A. Zaban, A. Gedanken,

    Langmuir 18, 4160 (2002)

    30. M.M. Rashad, I.A. Ibrahim, I. Osama, A.E. Shalan, Bull. Mater.

    Sci. 37, 903 (2014)

    31. A.E. Shalan, I. Osama, M.M. Rashad, I.A. Ibrahim, J. Mater. Sci.

    Mater. Electron. 25, 303 (2014)

    32. C. Wang, Y. Zhou, M. Ge, X. Xu, Z. Zhang, J.Z. Jiang, J. Am.

    Chem. Soc. 132, 46 (2010)

    33. A.A. Firooz, A.R. Mahjoub, A.A. Khodadadi, Mater. Lett. 62,

    1789 (2008)

    34. A. Sonia, Y. Djaoued, B. Subramanian, R. Jacques, M. Eric, B.

    Ralf, B. Achour, Mater. Chem. Phys. 136, 80 (2012)

    35. D. Kim, Y.-D. Huh, Mater. Lett.65, 2100 (2011)36. Y.C. Zhang, Z.N. Du, K.W. Li, M. Zhang, Sep. Purif. Technol.

    81, 101 (2011)

    37. H.P. Klug, L.E. Alexander, X-Ray Diffraction Procedures: For

    Polycrystalline and Amorphous Materials (Wiley, New York,

    1974)

    38. S. Yang, L. Gao, J. Am. Ceram. Soc.89, 1742 (2006)

    39. S. Bourbigot, M. Le Bras, R. Delobel, P. Breant, J.-M. Tremillon,

    Carbon 33, 283 (1995)

    40. S. Nie, Y. Hu, L. Song, Q. He, D. Yang, H. Chen, Polym. Adv.

    Technol. 19, 1077 (2008)

    41. A. Singhal, B. Sanyal, A.K. Tyagi, RSC Adv. 1, 903 (2011)

    Fig. 12 The photocatalytic performance of SnO2 nanoparticles (sample 1). a Absorption spectra of MB solution photocatalyzed by SnO2nanoparticles.b MB concentration dependence on irradiation time

    J Mater Sci: Mater Electron (2016) 27:5463 63

    1 3


Recommended