Surface & Coatings Technology

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Surface & Coatings Technology 326 (2017) 310–315

Contents lists available at ScienceDirect

Surface & Coatings Technology

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

Effect of alcohol solvents on TiO2 films prepared by sol–gel method


Orawan Wiranwetchayan a,⁎, Surin Promnopas a, Titipun Thongtem b,c,
Arnon Chaipanich a,c, Somchai Thongtem a,c
a
Department of Physics and Materials Science, Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand
b
Department of Chemistry, Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand
c
Materials Science Research Center, Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand

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

Article history: The TiO2 (anatase) films were prepared by sol-gel method using titanium (IV) isopropoxide (TTIP) as a precursor
Received 30 January 2017 and polyvinylpyrrolidone (PVP) as a polymeric precursor dissolved in different alcohol solvents (methanol, eth-
Revised 29 June 2017 anol, isopropanol and 1-butanol) and were heated treatment at 450 °C for 1 h. In this research, the alcohol sol-
Accepted in revised form 29 July 2017
vents have significant influence on the morphology, crystalline degree and energy band gap of TiO2. The
Available online 29 July 2017
number of carbon, boiling point, dielectric constant and viscosity of alcohol solvents were the main factor
Keywords:
influencing the size of TiO2 nanoparticles. The alcohol solvent with higher boiling point and viscosity, and long
TiO2 films length of carbon chain can lead to the smaller crystallite size. The number of carbon and viscosity of solvent
X-ray diffraction and crystallite size can play the important role in the agglomeration of particles. The energy band gaps of the
Electron microscopy films determined by UV-visible spectroscopy were in the 3.00–3.30 eV range.
Spectroscopy © 2017 Elsevier B.V. All rights reserved.

1. Introduction The sol-gel process involves conversion of monomers into a colloidal


solution (sol) acting as the precursor for an integrated network (gel)
The primarily application of titanium dioxide (TiO2) has been used of either discrete particles or network polymer. Typical precursors are
as a pigment in paints, printing ink, cosmetics, soap, toothpaste, poly- metal alkoxide. The final product might have several structures by con-
mer and others in which white coloration is required [1]. TiO2 occurs trolling gelation condition: the precursor hydrolysis, polycondensation,
in three different polymorphs: rutile (tetragonal), anatase (tetragonal) gel condensation and drying/calcination [35,36]. The titanium (IV)
and brookite (orthorhombic) [2]. At present, TiO2 has been widely isopropoxide (TTIP) precursor has been frequently used to produce
used in a number of applications such as photocatalysis [3,4] dye- the TiO2 film because TTIP is a suitable precursor with no toxic of by-
sensitized solar cells [5–8], gas sensors [9–11], batteries [12–14], self- products [37].
cleaning glasses [15], stormwater disinfection [16] and production of In this research, the anatase TiO2 films were prepared on glass sub-
hydrogen fuel [17,18]. strates by sol-gel method in different alcohol solvents (methanol, etha-
Thus, preparation of TiO2 film is the most important field of nol, isopropanol and 1-butanol), with TTIP as a precursor and PVP as a
nanomaterials. Several physical and chemical techniques such as polymeric precursor. The influences of alcohol solvents with different
sputtering [19,20], chemical vapor deposition (CVD) [21,22], metal- carbon numbers, boiling points, viscosity and dielectric constant on
organic chemical vapor deposition (MOCVD) [23], spray pyrolysis [24], the crystalline structure, morphology and energy band gap of TiO2
supersonic aerosol deposition [25], pulsed laser deposition [26], sol- were investigated and discussed.
gel [27–30], hydrolysis [31,32], hydrothermal [31] sparking [33] and
microwave-assisted solvothermal [34] have been used for preparing of
TiO2. Moreover, the properties of TiO2 films, crystalline structure, sur- 2. Experimental procedure
face area, size distribution, porosity and film thickness, are highly con-
trolled by preparation methods. Among them, the sol-gel method is 2.1. Materials
widely used because it requires low temperature processing, large
area coatings, simple and in-expensive. Thus, it is appropriate for fabri- Acetic acid (glacial) 100%, titanium (IV) isopropoxide 97% (TTIP,
cation of porous and homogeneous multicomponent oxide films [30]. Ti(OCH(CH3)2)4), polyvinylpyrrolidone (PVP, Mw = 360,000), 1-
butanol (butan-1-ol, CH3(CH2)3OH), ethanol 99% (C2H6O), methanol
⁎ Corresponding author. (CH3OH) and isopropanol (propan-2-ol, C3H7OH) were used in this
E-mail address: [email protected] (O. Wiranwetchayan). synthesis.

http://dx.doi.org/10.1016/j.surfcoat.2017.07.068
0257-8972/© 2017 Elsevier B.V. All rights reserved.
O. Wiranwetchayan et al. / Surface & Coatings Technology 326 (2017) 310–315 311

2.2. Preparation of TiO2 thin films films had strong adhesion with the glass substrates. Those fabricated
in methanol, ethanol and isopropanol contained large spherical parti-
TiO2 films were fabricated by the sol–gel method in different alcohol cles which were aggregated from primary particles to the larger ones.
solvents: methanol, ethanol, isopropanol and 1-butanol. Two solutions But for the film fabricated in 1-butanol, the small particles clustered to-
were prepared: (Sol1) 0.75 ml of TTIP was dissolved in a solution of gether in colonies. The size of aggregated particles (secondary particles)
1.5 ml of alcohol and 1.5 ml of acetic acid with continuous magnetic stir- characterized by SEM is summarized in Table 1. The average diameters
ring for 30 min. (Sol2) 0.3 g of PVP was added to 2.5 ml of alcohol sol- of the aggregate TiO2 particles were about 83 nm, 85 nm and 70 nm for
vent with 30 min stirring. The two solutions were mixed together and those prepared in ethanol, methanol and 1-butanol solvents,
vigorously stirred at room temperature for 120 min. In the end, the mix- respectively.
ture was coated on the glass substrates by doctor blade method and A transmission electron microscopy was used to characterize mor-
followed by 450 °C heat treatment for 1 h. phology and crystalline degree of the films. TEM bright field images
and selected area electron diffraction (SAED) patterns of TiO2 prepared
2.3. Characterization in different solvents are displayed in Fig. 2. Clearly, the particles of films
were mostly spherical shape with random distribution in different di-
Phase and structure of the films were characterized by X-ray diffrac- rections. The primary particle sizes were estimated to be ~13–16 nm.
tion (XRD, Rigaku MiniFlex™ II) operating at 20 kV and 15 mA using Cu- They should be noted that the alcohol solvents can lead to obtain nano-
Kα line in combination with the JCPDS database, Fourier transform particles. These results indicated that PVP can play an important role in
infrared spectroscopy (FTIR, Bruker Tensor 27) operating in the range producing spherical nanoparticles. The corresponding SAED patterns
of 400–4000 cm−1 and Raman spectrophotometry (T64000 HORIBA (inset of Fig. 2) present five diffraction rings. Calculated d-spaces and
Jobin Yvon) using a 50 mW and 514.5 nm wavelength Ar green laser. Miller indices of anatase TiO2 crystals were compared with those of
UV–visible spectra were examined by UV–visible spectroscopy the JCPDS database no. 21-1272 [38] and were summarized in Table 2.
(PerkinElmer Lambda 25). Surface morphology of the TiO2 films was They belonged to the (101), (004), (200), (211) and (204) planes of
studied by field emission scanning electron microscopy (FE-SEM, JEOL the anatase phase (JCPDS database no. 21-1272) [38]. The high-
JSM-6335F) operating at 15 kV and transmission electron microscopy resolution TEM (HR-TEM) images of TiO2 nanoparticles are shown in
(TEM, JEOL JEM-2010) operating at 200 kV. Fig. 3. The d-spacing values of the samples were ~ 0.35 nm and
~0.23 nm, corresponding to the (101) and (004) lattice planes of the an-
3. Results and discussion atase TiO2. The information obtained from d-spacing values was in ac-
cordance with the above results of SAED analysis.
Effect of solvent on the properties of the TiO2 crystals was studied by To further determine crystalline structure, XRD patterns of the sam-
classifying the alcohol with different carbon numbers: methanol, etha- ples produced in ethanol, methanol, 1-butanol and isopropanol solvents
nol, isopropanol and 1-butanol. Generally, the crystallization process after heat treatment at 450 °C for 1 h are shown in Fig. 4. The peaks at 2θ
plays a significant role in defining the crystalline structure, shape and of 25.3°, 37.8°, 48.1°, 53.9°, 55.0°, 62.7°, 68.7°, 70.1° and 75.0° can be
size of the TiO2 particles. The primary particles nucleated from precur- assigned to the (101), (004), (200), (105), (211), (204), (116), (220)
sor solution by molecular addition or aggregation of small subunits. and (215) crystal planes of tetragonal anatase phase (JCPDS database
The particle growth after nucleation was controlled by the alcoholic sol- no. 21-1272), respectively [2–4,38–40]. The XRD results are in accor-
vents. The surface morphology of the as-prepared TiO2 films with differ- dance with those of the SAED and HR-TEM analyses. Comparing the rel-
ent alcohol solvents and annealed at 450 °C for 1 h were examined by ative height of intensity of the (101) peak at 2θ = 25.5° with those of
FE-SEM, as the results presented in Fig. 1. The surfaces of all TiO2 films other peaks, there were the preferential growth direction of these prod-
were smooth and homogenous without crack detection. The TiO2 ucts. Furthermore, XRD analysis indicates that TiO2 is pure anatase

Fig. 1. FE-SEM images of TiO2 films prepared by sol-gel method in different alcohol solvents: (a) ethanol, (b) methanol, (c) 1-butanol and (d) isopropanol.
312 O. Wiranwetchayan et al. / Surface & Coatings Technology 326 (2017) 310–315

Table 1
The properties of alcohol solvents and average size of TiO2 nanopaticles.

Alcohol solvents Chemical formulae Dielectric constants Boiling points (°C) Viscosity at 20 °C (mPa·s) Crystallite size (Scherrer) (nm) Particle size (SEM) (nm)

Ethanol C2H6O 24.6 78.24 1.20 11.9 83


Methanol CH3OH 32.6 64.7 0.59 12.2 85
1-Butanol C4H10O 17.5 117.7 2.95 9.3 70
Isopropanol C3H8O 19.9 82.6 2.38 10.1 –

(JCPDS no. 21-1272) [38], regardless of the solvents used for preparing. found to have slightly increase in size in sequence: 1-butanol
The as-prepared films show the main peak almost at the same position, (~ 9.3 nm) b isopropanol (~ 10.1 nm) b ethanol (~ 11.9 nm)
caused by variation in defects of the films and alcohol solvents. The in- b methanol (~12.2 nm).
tensity of main peak is also different, caused by the difference in crystal- The effects of solvent properties, such as viscosity, dielectric con-
line degree and crystallite size. Upon changing of alcohol solvents in stant, boiling point and number of carbons on the crystallite size were
sequence of increasing boiling points (Tmethanol
b = 64.7 °C, Tethanol
b = analyzed as follows. According to the results (Table 1), the crystallite
isopropanol 1-butanol
78.24 °C, Tb = 82.6 °C and Tb = 117.7 °C), the films be- size seems to relate with dielectric constant (ε), viscosity (η) and boiling
came to lessen the crystalline degree, and the film produced in metha- point (Tb) of the alcohol solvents. The solvents with a lower dielectric
nol is the best crystal. constant resulted in a smaller crystallite size than those with a higher
The average crystallite size of TiO2 particles was calculated from the dielectric constant [34]. The particle size decreases as the number of car-
(101) diffraction peak by the Scherrer formula bon in the alcohol increases, leading to smaller crystallite size. More-
over, the different alcohol solvents can play the role in the crystallite
Kλ size of products which were found to relate with their boiling point
Crystallite size ðnmÞ ¼ ð1Þ
FWHM cosθ and viscosity. The crystallite size of TiO2 particles increases as the de-
creasing boiling point and viscosity of alcohol solvents [34].
where the shape factor K is 0.9, λ is the wavelength of Cu-Kα radiation The agglomeration of TiO2 nanoparticles in different alcohol solvents
(0.15406 nm), FWHM is full width at half maximum of the (101) peak was explained. Farrokhi-rad and Ghorbani [42] had measured the zeta
and θ is the Bragg's angle [41]. The crystallite size of TiO2 nanoparticles potential of TiO2 nanoparticles in different alcohols (methanol, ethanol,
prepared in different alcohol solvents are shown in Table 1. They were isopropanol and butanol) and found that the absolute value of zeta

Fig. 2. TEM images and SAED patterns of TiO2 prepared in different alcohol solvents: (a) ethanol, (b) methanol, (c) 1-butanol and (d) isopropanol.
O. Wiranwetchayan et al. / Surface & Coatings Technology 326 (2017) 310–315 313

Table 2
The d-spacing values and Miller indices of anatase TiO2 prepared in different alcohol sol-
vents as compared with those of the JCPDS database no. 21-1272 [38].

d-Spacing values (nm) Miller indices (hkl)

JCPDS database Ethanol Methanol Butanol Isopropanol

0.352 0.347 0.351 0.352 0.355 (101)


0.238 0.231 0.231 0.239 0.231 (004)
0.189 0.190 0.184 0.187 0.188 (200)
0.167 0.166 0.163 0.166 0.166 (211)
0.148 0.147 0.144 0.146 0.145 (204)

potential decreases with increasing the molecular size (carbon chain


length) of alcohol. The zeta potential demonstrated the variation of
the dispersion stability of TiO2 particles in alcohol solvents. The higher
zeta potentials presented an increasing of electrostatic repulsion be-
tween the two neighboring primary particles, leading to their separa-
tion and reduction the frequency of successful collision of particles
participating in Brownian motion [42,43]. Thus, the particles were re-
duced in different sizes. The particles in the suspension undergo ran-
domized Brownian motion which is significant for particle size of b 1 Fig. 4. XRD patterns of TiO2 films prepared in different alcohol solvents comparing with
the JCPDS database.
μm. The diffusion coefficient of colloidal particles (D) in a suspension
was determined by fundamental Einstein's equation [42]

kT Thus, TiO2 nanoparticles can diffuse and collide with each other more
D¼ ð2Þ
3πηd frequently in the suspension in sequence: methanol N ethanol N
isopropanol N 1-butanol, leading to the agglomerate growth rate in-
where k is Boltzmann's constant, T is absolute temperature, d is particle creases. Dogon and Golombok [44] had concluded that smaller particles
diameter and η is the viscosity of solvent. For the smaller particles and agglomerated faster than larger particles. Due to the SEM images shown
lower viscosities, colloidal particles have larger diffusion coefficient. in Fig. 1, TiO2 nanoparticles in methanol produced larger particle size

Fig. 3. HR-TEM images of TiO2 prepared in different alcohol solvents: (a) ethanol, (b) methanol, (c) 1-butanol and (d) isopropanol.
314 O. Wiranwetchayan et al. / Surface & Coatings Technology 326 (2017) 310–315

than those in ethanol and 1-butanol, as the results summarized in


Table 1.
Fig. 5a shows the Raman spectra of TiO2 films prepared in ethanol,
methanol, 1-butanol and isopropanol solvents. The tetragonal anatase
has I41/amd space group and contains six Raman active modes
(A1g + 2B 1g + 3E g ) [2]. All of the samples have similar peaks
appearing at 143, 394, 513 and 635 cm− 1, corresponding to the Eg,
B1g, A1g and Eg modes of anatase phase, respectively. They have been
known that Eg and B1g peaks are mainly caused by symmetric stretching
vibration and bending vibration of O\\Ti\\O, while A1g peak is caused
by anti-symmetric bending vibration of O\\Ti\\O [39,45]. Meanwhile,
Raman peaks of the samples shows different intensities controlled by
their crystallite size and crystalline degree [46]. The improvement of
crystallinity of anatase phase prepared in methanol solvent can be ex-
plained by the decrease in boiling point and viscosity of alcohol with de-
creasing in the length of carbon chains.
The FTIR spectra disclose chemical bonding information of the TiO2
films prepared in different solvents as the results shown in Fig. 5b. The
absorption broad band at 3421 cm−1 assigned to the hydroxyl groups
(O\\H) [45,47,48] and that at 1637 cm− 1 ascribed to the H\\O\\H
bending vibration, indicating that TiO2 contains physically absorbed
water on the surface [42,49]. The band at 1384 cm−1 can be assigned
to C\\OH or C\\H stretching vibration. Meanwhile, the FTIR spectra of
TiO2 films show intense broad band in the vicinity of 400–800 cm−1,
corresponding to Ti\\O\\Ti asymmetric stretching as expected for
TiO2 [47,48]. They should be noted that the broad band at 3421 cm−1
and narrow band at 1637 cm−1 gradually changed with different alco-
hol solvents and crystalline degree of TiO2 particles. In conclusion, the
crystallization process was influenced by the surface hydroxyl groups
interacted with the alcohol solvents [49]. Raman and FTIR spectra indi-
cated the typical anatase TiO2 phase and these results were also in ac-
cordance with the above XRD results.
UV–visible spectroscopy and the plots of (αhυ)1/2 versus photon en-
ergy (hυ) of TiO2 films prepared in different alcohol solvents are shown
in Fig. 6. The transmittance of the TiO2 films changed with the alcohol
solvents. When photon energy is greater than energy gap, the absorp-
tion is linearly increased with increasing of photon energy. Inclination
of the linear portion was controlled by charged diffusion from conduc-
tion band to valence band. The band gap (Eg) of a semiconductor can
be calculated by the relationship between the absorption coefficient
and the incident photon energy by the following equation

n
Fig. 5. (a) Raman and (b) FTIR spectra of TiO2 films prepared in different alcohol solvents. αhν ¼ Aðhν−Eg Þ ð3Þ

Fig. 6. Transmittance and the plots of (αhυ)1/2 versus photon energy (hυ) of TiO2 films prepared in different alcohol solvents.
O. Wiranwetchayan et al. / Surface & Coatings Technology 326 (2017) 310–315 315

where A is a constant, ν is the frequency of the incident photon and h is [14] Y. Zhao, W. Zhu, G.Z. Chen, E.J. Cairns, J. Power, Sources 327 (2016) 447–456.
[15] Z. Liu, X. Zhang, T. Murakami, A. Fujishima, Sol. Energy Mater. Sol. Cells 92 (2008)
the Planck's constant, n = 2 for indirect transition [2,41]. The absorption 1434–1438.
coefficient (α) was calculated by α ≈ d1 ln ½1T, where T and d are the [16] G. Wang, W. Feng, X. Zeng, Z. Wang, C. Feng, D.T. McCarthy, A. Deletic, X. Zhang,
Water Res. 94 (2016) 363–370.
transmittance and the film thickness, respectively [2]. For the plot of
[17] S. Sun, P. Gao, Y. Yang, P. Yang, Y. Chen, Y. Wang, Appl. Mater. Interfaces 8 (2016)
(αhυ)1/2 vs (hυ), the indirect energy gap of TiO2 films was obtained 18126–18131.
by extrapolating the linear portion of the plot to zero absorbance. [18] B. Wu, D. Liu, S. Mubeen, T.T. Chuong, M. Moskovits, G.D. Stucky, J. Am. Chem. Soc.
138 (2016) 1114–1117.
They were found to be 3.30, 3.25, 3.20 and 3.00 eV for ethanol, 1-buta-
[19] C. Guillѐn, J. Montero, J. Herrero, J. Mater. Sci. 49 (2014) 5035–5042.
nol, methanol and isopropanol, respectively. [20] B. Houng, Y.H. Shih, S.H. Lu, W.C. Chien, J. Electroceram. 36 (2016) 87–93.
[21] D. Guo, A. Ito, T. Goto, R. Tu, C. Wang, Q. Shen, L. Zhang, J. Adv. Ceram. 2 (2) (2013)
4. Conclusions 162–166.
[22] C. Edusi, G. Sankar, I.P. Parkin, Chem. Vap. Depos. 18 (2012) 126–132.
[23] R.A. Antunes, M.C.L. de Oliveira, M.F. Pillis, Int. J. Electrochem. Sci. 8 (2013)
The influence of alcohol solvents (methanol, ethanol, isopropanol 1487–1500.
and 1-butanol) on TiO2 films prepared by sol–gel method is reported. [24] S. Dhanapandian, A. Arunachalam, C. Manoharan, J. Sol-Gel Sci. Technol. 77 (2016)
119–135.
The study showed that all of the fabricated samples had the crystalline [25] J.-J. Park, D.-Y. Kim, S.S. Latthe, J.-G. Lee, M.T. Swihart, S.S. Yoon, Appl. Mater. Inter-
structure of anatase phase. The different alcohol solvents have the influ- faces 5 (2013) 6155–6160.
ence on the morphology, crystalline degree and energy band gap. In ad- [26] A. Ishii, Y. Nakamura, I. Oikawa, A. Kamegawa, H. Takamura, Appl. Surf. Sci. 347
(2015) 528–534.
ditional, the crystallite size and agglomeration of TiO2 particles were [27] O. Wiranwetchayan, Q. Zhang, X. Zhou, Z. Liang, P. Sngjai, G.Z. Cao, Chalcogenide
found to be controlled by different alcohol solvents related to the num- Lett. 9 (2012) 157–163.
ber of carbon in alcohol, boiling point, dielectric constant and viscosity. [28] M.A. Behnajady, H. Eskandarloo, Res. Chem. Intermed. 41 (2015) 2001–2017.
[29] R.S. Sabry, Y.K. Al-Haidarie, M.A. Kudhier, J. Sol-Gel Sci. Technol. 78 (2016) 299–306.
In this research, the energy band gaps of the as-prepared TiO2 were de-
[30] S. Šegota, L. Ćurkovic, D. Ljubas, V. Svetličić, I.F. Houra, N. Tomašić, Ceram. Int. 37
termined to be 3.00–3.30 eV. (2011) 1153–1160.
[31] S. El-Sherbiny, F. Morsy, M. Samir, O.A. Fouad, Appl. Nanosci. 4 (2014) 305–313.
[32] J. Yue, Z. Chen, Y.E.L. Chen, J. Zhang, Y. Song, Y. Zhai, Nanoscale Res. Lett. 9 (2014)
Acknowledgment
1–6.
[33] W. Thongsuwan, T. Kumpika, P. Singjai, Curr. Appl. Phys. 8 (2008) 563–568.
We wish to thank the Thailand Research Fund and the Thailand's Of- [34] Y.-C. Wu, Y.-C. Tai, J. Nanopart. Res. 15 (2013) 1–11.
fice of the Higher Education Commission for providing financial support [35] B. Duymaz, Z.V. Yigit, M.G. Seker, F. Dündar, Acta Phys. Pol. A 129 (2016) 872–874.
[36] S. Segota, L. Curkovic, D. Ljubas, V. Svetlicic, I.F. Houra, N. Tomasic, Ceram. Int. 37
through the contact MRG6080011. (2011) 1153–1160.
[37] S. Collette, J. Hubert, A. Batan, K. Baert, M. Raes, I. Vandendael, A. Daniel, C.
References Archambeau, H. Terryn, F. Reniers, Surf. Coat. Technol. 289 (2016) 172–178.
[38] Powder Diffract. File, JCPDS-ICDD, 12 Campus Boulevard, Newtown Square, PA
[1] D.A.H. Hanao, C.C. Sorrell, J. Mater. Sci. 46 (2011) 855–874. 19073-3273, U.S.A.2001.
[2] W. Promnopas, S. Promnopas, T. Phonkhokkong, T. Thongtem, D. Boonyawan, L. Yu, [39] Y. Cao, Q. Li, C. Li, J. Li, J. Yang, Appl. Catal. B 198 (2016) 378–388.
O. Wiranwetchayan, A. Phuruangrat, S. Thongtem, Surf. Coat. Technol. 306 (2016) [40] M. Stefan, O. Pana, C. Leostean, C. Bele, D. Silipas, M. Senila, E. Gautron, J. Appl. Phys.
69–74. 116 (2014) 1–13.
[3] C.B.D. Marien, C. Marchal, A. Koch, D. Robert, P. Drogui, Environ. Sci. Pollut. Res. 24 [41] O. Wiranwetchayan, W. Promnopas, K. Hongsith, S. Choopun, P. Singjai, S.
(2017) 12582–12588. Thongtem, Res. Chem. Intermed. 42 (2016) 3655–3672.
[4] X.H. Lin, Y. Wu, J. Xiang, D. He, S. Fong, Y. Li, Appl. Catal. B 199 (2016) 64–74. [42] M. Farrokhi-rad, M. Ghorbani, Ceram. Int. 38 (2012) 3893–3900.
[5] S. Kathirvel, C. Su, Y.J. Shiao, Y.-F. Lin, B.R. Chen, Sol. Energy 132 (2016) 310–320. [43] G. Borsoi, B. Lubelli, R. van Hees, R. Veiga, A. Santos Silvad, L. Collae, L. Fedele, P.
[6] J. Liu, J. Luo, W.-G. Yang, Y.-L. Wang, L.-Y. Zhu, Y.-Y. Xu, Y. Tang, Y.-J. Hu, C. Wang, Y.- Tomasin, Colloids Surf. A Physicochem. Eng. Asp. 497 (2016) 171–181.
G. Chen, W. Shi, J. Mater. Sci. Technol. 31 (2015) 106–109. [44] D. Dogon, M. Golombok, J. Pet. Explor. Prod. Technol. 5 (2015) 91–98.
[7] S. Nagai, G. Hirano, T. Bessho, K. Satoria, Vib. Spectrosc. 72 (2014) 66–71. [45] F. Tian, Y. Zhang, J. Zhang, C. Pan, J. Phys. Chem. C 116 (2012) 7515–7519.
[8] S.J. Kwon, H.B. Im, J.E. Nam, J.K. Kang, T.S. Hwang, K.B. Yid, Appl. Surf. Sci. 320 (2014) [46] M.K. Ahmad, S.M. Mokhtar, C.F. Soon, N. Nafarizal, A.B. Suriani, A. Mohamed, M.H.
487–493. Mamat, M.F. Malek, M. Shimomura, K. Murakami, J. Mater. Sci. Mater. Electron. 27
[9] D. Mardare, N. Cornei, C. Mita, D. Florea, A. Stancu, V. Tiron, A. Manole, C. Adomnitei, (2016) 7920–7926.
Ceram. Int. 42 (2016) 7353–7359. [47] S.H. Othman, S. Abdul Rashid, T.I. Mohd Ghazi, N. Abdullah, J. Nanomater. 2012
[10] J. Wu, C. Zhang, Q. Li, L. Wu, D. Jiang, J. Xia, Solid State Ionics 292 (2016) 32–37. (2012) 1–10.
[11] O. Alev, E. Sennik, N. Kılınç, Z.Z. Öztürka, Procedia Eng. 120 (2015) 1162–1165. [48] A. Qu, H. Xie, X.Xu.Y. Zhang, S. Wen, Y. Cui, Appl. Surf. Sci. 375 (2016) 230–241.
[12] R. Gao, Z. Jiao, Y. Wang, L. Xu, S. Xia, H. Zhang, Chem. Eng. J. 304 (2016) 156–164. [49] B.M. Rao, S.C. Roy, J. Phys. Chem. C 118 (2014) 1198–1205.
[13] L. Yang, H.Z. Li, J. Liu, Y. Lu, S. Li, J. Min, N. Yan, Z. Men, M. Lei, J. Alloys Compd. 689
(2016) 812–819.

You might also like