Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

J Nanopart Res (2018) 20:23

https://doi.org/10.1007/s11051-018-4140-7

RESEARCH PAPER

Microwave-assisted synthesis of TiO2 nanoparticles:


photocatalytic activity of powders and thin films
G. S. Falk & M. Borlaf & M. J. López-Muñoz &
J. C. Fariñas & J. B. Rodrigues Neto & R. Moreno

Received: 1 June 2017 / Accepted: 19 January 2018


# Springer Science+Business Media B.V., part of Springer Nature 2018

Abstract A simple, rapid, and effective synthesis meth- that of a thin film of commercial TiO2 powder prepared
odology for the preparation of high-performance TiO2 under similar conditions.
nanoparticles and thin films by combining colloidal sol-
gel and microwave-assisted hydrothermal synthesis was Keywords TiO2 . Nanoparticles . Microwave synthesis .
developed. The obtained results indicate that the heating Thin films . Photocatalysis
with microwaves at 180 °C for 20 min was enough to
synthesize crystalline TiO2 nanoparticles, presenting an-
atase as a major phase with a crystal size of ~ 7 nm and a
Introduction
specific surface area of 220 m2 g−1. A secondary thermal
treatment improved the crystallinity and induced the
Titanium dioxide (TiO2) is a material of great interest
anatase-to-rutile transformation. The highest photocata-
due to its excellent set of chemical, physical, optical, and
lytic activity was found for the as-synthesized powder
electronic properties that make this material to be a
without any additional thermal treatment. Thin films
reference in several technological applications. A huge
were also prepared by dip-coating and its high photo-
number of applications such as air and water treatment
catalytic activity showed a kinetic curve comparable to
(Belver et al. 2015; Weon et al. 2017), self-cleaning
surfaces (Pinho et al. 2015), electronic devices (Kim
G. S. Falk (*) : M. Borlaf : J. B. Rodrigues Neto et al. 2016), and dye-sensitized solar cells (Liang et al.
Department of Mechanical Engineering, Federal University of
Santa Catarina, Florianópolis, SC 88040-900, Brazil
2016; Hou et al. 2016; Shen et al. 2016; Cui et al. 2017),
e-mail: gilbertofalk@outlook.com to name a few, have been reported in the literature.
In particular, TiO2 has been extensively studied in
M. J. López-Muñoz different photocatalytic systems, including the
Department of Chemical and Energy Technology, Chemical and
Environmental Technology, Mechanical Technology and
photodegradation of volatile organic compounds, hy-
Analytical Chemistry, Rey Juan Carlos University, C/Tulipán s/n, drogen production, and other environmental applica-
28933 Móstoles, Madrid, Spain tions due to its n-type semiconductor properties with a
band gap of 3.0 to 3.2 eV (dependent on crystallograph-
J. C. Fariñas : R. Moreno
Instituto de Cerámica y Vidrio, CSIC, Kelsen 5, 28049 Madrid,
ic variants), low toxicity potential, good chemical and
Spain thermal stabilities, surface acidity, and high selectivity,
as well as cheapness and availability (Fan et al. 2016; Li
et al. 2016; Cortéz-Lorenzo et al. 2017; Sun et al. 2017).
Present Address:
M. Borlaf
Different methods have been reported for the synthe-
Laboratory for High Performance Ceramics, Empa, sis of titanium dioxide, including hydrothermal methods
Überlandstrasse 129, 8600 Dübendorf, Switzerland and sol-gel processes using a variety of precursors such
23 Page 2 of 10 J Nanopart Res (2018) 20:23

as titanium alcoxides, titanium tetrachloride (Kobayashi the easier recovery of the catalyst after the reaction in
et al. 2007; Adjimi et al. 2014; Elsellami et al. 2017). In comparison to slurry systems.
most cases, however, long periods of times are usually
required to achieve the crystallization of titania in a
specific phase (mainly anatase or rutile). In this context, Experimental
the microwave-assisted hydrothermal method is a rapid,
energy-saving, and efficient synthesis methodology for Synthesis and characterization of the TiO2 nanoparticles
the preparation of functional nanomaterials (Meng et al.
2016). It has received great attention in the technological TiO2 nanoparticles were synthesized combining con-
field because of its suitability for reducing the reaction ventional colloidal sol-gel with microwave-assisted hy-
time from days to hours or minutes, demonstrating drothermal method. The preparation started as in any
advantages that include tailored microstructure, unique colloidal sol-gel process with the quick hydrolysis, un-
properties, and an improved product yield, leading to a der continuous stirring, of titanium (IV) isopropoxide
reduction in manufacturing costs and enhanced versatil- (TIPO, 97%, Sigma-Aldrich, Steinheim, Germany) in a
ity for the synthesis of new materials as compared to mixture of deionized water (18.2 MΩ cm−1, ultra-pure
conventional methods (Jiang et al. 2015; Esquivel- Milli-Q, France) and nitric acid (65%, Panreac, Barce-
Escalante et al. 2016; Mirzaei and Neri 2016). In addi- lona, Spain) in H2O:Ti4+ = 100:1 and H+:Ti4+ = 0.2 mo-
tion, colloidal sol-gel is performed in water so that the lar ratios. Nitric acid was used as a dispersing agent/
synthesis method is more economic and efficient than catalyst.
other synthesis methods used elsewhere. The solution was stirred continuously at 700 rpm for
The reaction process for microwave-assisted hy- 30 min, and then transferred to a 100-mL sealed vessel
drothermal synthesis is dependent on the interac- made of high-purity TFM (modified teflon) placed in a
tions of the microwaves with the materials and the commercial microwave digestion system (Milestone
reaction medium, that is, under the action of the ETHOS One, Sorisole, Italy) where hydrothermal reac-
electromagnetic field, the particles can produce dif- tions occurred at a microwave frequency of 2.45 GHz.
ferent types of polarization (electron, atom, orienta- The container was surrounded by a safety shield. Tem-
tion, and space charge), being the microwave energy perature and pressure during the synthesis were moni-
transmitted directly to the material in terms of mo- tored and controlled with a shielded thermocouple di-
lecular interactions with the electromagnetic field, rectly inserted into the vessel and with a pressure trans-
leading to a fast and homogenous heating process ducer sensor connected to the vessel, respectively. Built-
in short periods of time and generating a homoge- in magnetic stirring (Teflon-coated stirring bar) was
neous nucleation and a uniform growth of the nano- used. For each synthesis, the heating ramp was set in
particles (Zhu and Hang 2013; Bhattacharya and 10 min with holding times of 20 min at maximum
Basak 2016; Mirzaei and Neri 2016). temperatures of 180 and 200 °C. After cooling at room
Although different routes of TiO2 syntheses have temperature, the obtained powders were carefully
been extensively described in the literature, documenta- washed for at least three times with deionized water
tion of microwave-assisted synthesis in short periods of and centrifuged every time at a rotation speed of
time and in water is still scarcely reported. In this con- 2400 rpm using a multi-purpose bench top centrifuge
text, this research had the objective of both developing a (Núve NF800, Ankara, Turkey).
rapid, energy-saving, and promising clean method for Particle size distributions (PSD) before and after the
the synthesis of TiO2 nanoparticles by using a colloidal microwave process were measured by laser diffraction
sol-gel approach accelerated by microwave-assisted hy- (LD, Mastersizer, Malvern, UK) and dynamic light scat-
drothermal route, and evaluating the structural, morpho- tering (DLS) in a Nanozetasizer ZEN 3600 equipment
logical, and photocatalytic properties of the resulting (Malvern, UK).
nanopowders as well as thin films obtained by dip- The nanoparticles were obtained by freeze-drying
coating from those nanopowders. With the purpose of process as it has been shown that, in contrast to conven-
photocatalytic applications, the use of titania films tional drying treatments, this procedure allows to obtain
immobilized on supports transparent to UV radiation quite homogeneous particle size distributions
can be of great interest, as it offers as main advantage, (Eggenhuisen et al. 2013). This technique consists of
J Nanopart Res (2018) 20:23 Page 3 of 10 23

two steps: freezing and drying. In the first one, the In the meantime, prior to irradiation, the 20 mg L−1 MB
solutions were frozen by cooling at − 50 °C, and in the aqueous solution containing 0.25 g L−1 of photocatalyst
second, the sublimation of the solvent and other volatile was magnetically stirred for 30 min in the dark to reach
compounds was carried out with a vacuum pump at − an equilibrated MB adsorption. The amount of dye
50 °C (CRYODOS-50, Telstar, Spain). Subsequently, adsorbed onto the catalyst was calculated by the differ-
the obtained nanoparticles were submitted to thermal ence between the initial MB concentration and that
treatments in air atmosphere using heating and cooling remaining in the solution after the equilibration interval.
rates of 10 °C min−1 and a dwell time of 1 h. Photocatalytic reactions were carried out at pH = 5.
The specific surface area (SSA) was measured with a Samples were periodically taken from the reactor for
one-point BET method (Monosorb Surface Area analysis. The MB degradation was quantified by mea-
Analyser, MS-13, Quantachrome Corporation, USA) suring the absorbance at 665 nm (the maximum of the
after degassing at 150 °C for 2 h. absorption band of MB) in an UV-Vis spectrophotome-
X-ray diffraction (XRD) analyses were conducted ter (JASCO V-630, JASCO Inc., Japan). The commer-
with a PW-1830 diffractometer (Philips model X’Pert - cial TiO2 (anatase TiO2 nanopowder < 25 nm, 99.7%,
NL, Eindhoven, the Netherlands) with CuKα radiation SSA = 45–55 m2 g−1, Sigma-Aldrich, Germany) and
(k = 1.5406 Å) as X-ray source over 2-h range (10–80). TiO2 P25 (SSA ~ 50 m2 g−1, Evonik, Essen, Germany),
Raman spectroscopy analyses were performed with a which are customarily used as references due to their
spectrometer (inVia, UK) coupled to an Olympus mi- excellent photocatalytic properties, have also been eval-
croscope (BX41 TM). All spectra were obtained using uated for comparative purposes.
an argon laser (λ = 514.5 nm).
Transmission electron microscopy (TEM) analyses Thin film preparation and characterization
were performed on a Philips Technai 20 T electron
microscope operating at 200 kV and with a 2.7-Å reso- TiO2 thin films supported on glass slides were obtained
lution. The samples were dispersed in acetone applying by dipping technique using a withdrawal rate of 1 mm s−1
1 min of sonication and then a droplet of the suspension at room temperature. Glass substrates were used after
was deposited onto a carbon-coated copper grid. washing with deionized water, acetone, and finally eth-
The UV-Vis diffuse reflectance spectra were collect- anol. A suspension was prepared by dispersing 1 g of
ed using a spectrophotometer PerkinElmer Lambda 950 TiO2 nanoparticles (obtained by microwave-assisted
precisely (USA). From the spectra, the optical band gap hydrolysis after centrifugation and subsequent freeze-
was determined using the Tauc plot (Tauc 1970) for an drying) in a mixture of 1 mL deionized water
indirect semiconductor: (18.2 MΩ cm−1, ultra-pure Milli-Q, France) and 9 mL
of ethanol absolute (EtOH, ≥ 99.8%, Sigma-Aldrich,
ðαhν Þ0:5 vs:hν ð1Þ Germany). The pH = 3 was adjusted with nitric acid
and ultrasonication (UP 400s, Hielscher Ultrasonics
where α is the absorption coefficient and hν is the GmbH, Teltow, Germany; power = 400 W) was applied
photon energy in electronvolt. for 1 min in order to improve the dispersing conditions.
The photocatalytic activity of the synthesized sam- The thin films were dried at room temperature for
ples was evaluated through the degradation of methy- 30 min and subsequently stabilized in a muffle at
lene blue (MB) using a batch reactor with an effective 200 °C for 1 h to remove the solvent and to induce the
solution volume of 1 L. The light source was a medium condensation process. Stabilized films were then ther-
pressure Hg lamp (150 W, TQ-150, Heraeus Instru- mally treated at 500 °C for 1 h in air. The same proce-
ments, Germany; maximum intensity at λ = 365 nm) dure was followed to obtain coatings by dipping of a
axially immersed inside the reactor and provided with commercial nanosized TiO2 powder (P25, Evonik, Es-
a cooling tube. An aqueous copper sulphate solution sen, Germany) which is customarily used as a reference
(0.01 M) kept at 25 °C was circulated through the due to its excellent photocatalytic properties.
cooling tube both to prevent the overheating of the lamp The thickness and the refractive index (RI) of the thin
and the solution and to avoid photolytic processes. To films were determined by ellipsometry (GES5E Spec-
achieve a stabilized radiation emission, the lamp was troscopic Ellipsometer from SOPRALAB, Courbevoie,
switched on 20 min before being fitted into the reactor. France). The measurements were done from 1.22 to
23 Page 4 of 10 J Nanopart Res (2018) 20:23
35
3.00 eV at an incidence angle of 75°. The obtained After microwave process
curves were analyzed with the help of the WINSE 30
180 °C
software.
25
Atomic force microscopy (AFM; 200 °C Before microwave process

Volume (%)
NanotecElectronica, Madrid, Spain) was used to ob- 20
serve the surface topography of selected thin films.
15
The films were also characterized by Raman spectros-
copy using the same procedure previously cited for the 10
analysis of the synthesized nanosized powder.
5
The photocatalytic activity of the nanocrystalline
titania coatings was evaluated in a batch reactor
consisting of a Pyrex cylinder of similar size to that 0 20 40 60 2000 4000 6000 8000
Particle size (nm)
described above for the reactions carried out with the
powdered samples, provided with the same lamp as Fig. 1 Particle size distribution before and after the microwave
process
source of irradiation. The initial concentration of MB
was set at 10 mg L−1. Six glass slides, each one covered
formation of uniform nanoparticles in just few minutes
with an area of 3 × 7 cm2 of titania, were placed around
whereas the conventional methods require longer pe-
the lamp in a symmetrical arrangement to ensure they
riods of time to obtain the nanoparticles (Delekar et al.
got the maximum UV radiation. The total amount of
2012; Borlaf et al. 2014; Zhang et al. 2016).
catalyst was calculated to be approximately 0.05 g L−1.
The colloidal stability of the TiO2 nanoparticles ob-
To keep the slides at upright position along the reaction
tained after the microwave process was evaluated by
time throughout the continuous magnetic stirring of the
measuring the zeta potential as a function of pH (Fig. 2).
solution, the slides were hold by a polymer rack.
It can be observed that at either acid pH values ≤ 5 or
basic pH ≥ 9, the zeta potential absolute values vary
between 30 and 40 mV, which are considered sufficient
Results and discussion
to maintain the stability of the nanoparticles (Moreno
2012). For pH values between 6 and 8, the zeta potential
Synthesis and characterization of the TiO2 nanoparticles
values are lower than 20 mV, not considered enough for
keeping the nanoparticles stable in the system. The
Measurements of PSD, displayed in Fig. 1, were per-
isoelectric point is observed at pH ~7.5, within the range
formed in order to evaluate the effectiveness of the
reported in the literature for TiO2 anatase (Kosmulskil
microwave-assisted hydrothermal method in the prepa-
1992; Song et al. 2008; Borlaf et al. 2014).
ration of TiO2 nanoparticles. As it can be observed, the
Figure 3 shows the XRD patterns of the as-prepared
fast hydrolysis of TIPO initially generates large agglom-
samples obtained after the microwave-assisted synthesis
erates ranging from 2 to 7 μm. Subsequently, after the
microwave process, nanoparticles with a size ranging
50
from 7 to 28 and 13 to 52 nm were obtained for the
40
samples synthesized at 180 and 200 °C/20 min, respec-
30
tively. This increase in particle size with temperature
20
Zeta Potential (mV)

could be associated to an excess of energy, which pro-


10
vokes the re-agglomeration of the nanoparticles (Falk
et al. 2016; Ashok and Rao 2017). In addition, the lower 0
pH
2 4 6 8 10
temperature led to a narrower monomodal distribution, -10

so that keeping in mind the goal to get homogeneous -20

titania particles with high specific surface area, 180 °C -30

was chosen as temperature for microwave processing. -40


In general, this demonstrates that this method -50
coupled with the rapid and homogeneous heating ob- Fig. 2 Variation of zeta potential of the synthesized nanoparticles
tained by microwave radiation is capable of inducing the as a function of pH
J Nanopart Res (2018) 20:23 Page 5 of 10 23

R
R 143 600 °C
600 °C
A R

R A R
R R
AR A R
R R A
Intensity (a.u.)

A
446 640
514 608
500 °C
A 195 398
A AA A
A A A
B
A

Intensity (a.u.)
as-prepared 500 °C
A A 144
A A
B A A

20 30 40 50 60 70 80
2 (degrees)
398 517 640
Fig. 3 XRD patterns of the as-prepared and calcined powders at 195
500 and 600 °C/1 h. (A—anatase, B—brookite, and R—rutile)

and after subsequent calcination at different tempera- 150 as-prepared


tures. It is possible to observe that the microwave-
assisted hydrothermal method without further heat treat-
ments was sufficient for the successful crystallization of
titania, with well-defined peaks typical of the anatase 398 517 640
TiO2 phase (JCPDS No. 01-071-1166), although a mi-
nor contribution of brookite (JCPDS No. 01-076-1936)
is also present as evidenced by the small peak centered 200 400 600 800
-1
at ca. 31°. Up to 500 °C/1 h, the crystallinity of the Raman Shift (cm )
anatase phase increased, according to the narrowing of Fig. 4 Raman spectra of the as-prepared and calcined TiO2
diffraction peaks, and at 600 °C/1 h, the anatase-to-rutile nanoparticles
(JCPDS No. 01-087-0710) transformation is observed
while brookite phase completely disappears. Even in good agreement with the XRD analysis. Neverthe-
though at 600 °C/1 h the rutile fraction is higher than less, it should be noticed that according to the XRD
anatase one, the full conversion of the anatase-to-rutile patterns, brookite is present in much lower proportion in
phase requires higher temperatures. comparison to anatase. Considering that the most in-
The Raman spectra of the samples obtained by the tense Raman band of brookite is centered at ca.
microwave processing as well as after thermal treat- 153 cm−1 (Tompsett et al. 1995) but also that all the
ments at different temperatures are shown in Fig. 4. It bands are less intense than the anatase or rutile ones, if
can be seen that the as-prepared sample shows Raman present, it must be overlapped by the band at 150 cm−1
bands centered at 150, 195, 398, 517, and 640 cm−1 ascribed to Eg(1) mode of anatase and the rest of the
which are assigned to Eg(1), Eg(2), B1g, B1g, and Eg(3) bands would be very difficult to detect.
modes in anatase phase, respectively (Tompsett et al. High-resolution TEM micrographs of the samples are
1995; Huang et al. 1997; Chen and Mao 2007). The shown in Fig. 5. A disordered three-dimensional num-
spectrum of the sample calcined at 500 °C/1 h exhibits ber of nanoparticles with different morphologies and
similar features. The main difference is related to the unit dimensions lower than 7 nm are clearly observed
shift of the band Eg(1) to 144 cm−1, which, together with in the sample obtained by the microwave-assisted hy-
the decrease of the width of the band, is associated to the drothermal method (Fig. 5a). After calcining at 500 °C/
larger particle size reached after the thermal treatment 1 h (Fig. 5b), the crystals are larger, as it is expected after
(Borlaf et al. 2015). At 600 °C/1 h, the bands associated a thermal treatment, with sizes larger than 10 nm. No
to the rutile phase are observed at 446 and 608 cm−1, significant difference is observed in the selected area
attributed to the Eg and A1g modes (Tompsett et al. electron diffraction (SAED) analysis between the two
1995; Chen and Mao 2007). These results are therefore samples. In both cases, it shows the typical diffuse
23 Page 6 of 10 J Nanopart Res (2018) 20:23

a)
5


 h 
2

1 as-prepared
500 °C
600 °C
0
2 3 4
h (eV)
b) Fig. 6 Tauc representation of the UV-Vis spectra obtained from
the as-prepared and calcined samples at 500 and 600 °C/1 h. The
arrows indicate the band gap value calculated by extrapolation

The photocatalytic efficiency of the synthesized ma-


terials under UV radiation was evaluated by the oxida-
tion of MB. The evolution of normalized MB concen-
tration as a function of the irradiation time is displayed
in Fig. 7. Adsorption experiments carried out in the dark
prior to irradiations evidenced a scarce adsorption of the
dye (less than 0.8% of initial MB concentration) on all
materials. The extent of direct photolysis of the dye was
evaluated by irradiation of the MB solutions without any
Fig. 5 TEM images of the as-prepared TiO2 nanoparticles (a) and catalyst. As can be seen in Fig. 7, in these conditions,
the calcined ones at 500 °C/1 h (b) only 10% of initial MB was degraded after 150 min of
reaction, thus giving a reference for the catalyst effect in
polycrystalline pattern for a TiO2 anatase phase, with the all other experiments.
main planes identified in (101), (004), (200), and (105), As it can be seen, for equal catalyst loading, the
being the first one of greater intensity and the rest with sample obtained by microwave-assisted hydrother-
similar intensity. mal method without subsequent heat treatments
Figure 6 shows the Tauc plot of the samples obtained
by microwave process and as a function of the calcina- 1,0
tion temperature. The average band gap was estimated
from the intercept of the linear portion of the (νhν)0.5 vs. 0,8

hν plots on energy axis at (νhν)0.5 equal to 0. The band


0,6
gap calculated for the as-prepared samples was ~ 3.1 eV
C/C0

and decreased at higher calcination temperatures up to ~ 0,4


3.0 eV. This fact can be explained by the transformation
of the anatase phase into rutile (600 °C/1 h), as rutile is 0,2 Photolysis
as-prepared
characterized by a lower theoretical band gap value in 500 °C
600 °C
comparison to anatase (3.2 and 3.0 eV for anatase and 0,0 TiO2 Aldrich
TiO2 P25
rutile, respectively) (Henderson 2011). Nevertheless,
0 20 40 60 80 100 120 140 160
not only the crystalline composition but other factors Irradiation Time (min)
can affect to the observed decrease of the band gap value
Fig. 7 Kinetic curves of the methylene blue photodegradation
as, for example, different defects such as oxygen vacan- using the as-prepared and calcined TiO2 samples as a function of
cies, which are typical in TiO2 (Diebold 2003). the exposure time under UV radiation
J Nanopart Res (2018) 20:23 Page 7 of 10 23

shows the highest photocatalytic activity, followed Thin film preparation and characterization
by the sample calcined at 500 °C/1 h (anatase) and
600 °C/1 h (anatase + rutile). In the first two cases, Once confirmed the photocatalytic activity of titania
the photoactivity was higher compared to commer- nanopowders obtained by the microwave-assisted
cial anatase TiO2 nanoparticles (from Sigma-Aldrich) hydrothermal synthesis, the next goal was to prepare
used as a reference. The SSA of the samples was titania films on glass supports. In comparison to
measured, obtaining 220, 65, and 43 m2 g−1 for the powdered samples, the main interest of using
samples prepared by freeze-drying process without immobilized titania is the possibility of reusing the
further thermal treatment and calcined at 500 and photocatalyst.
600 °C/1 h, respectively. It is well-known that the The suspension was prepared under acidic conditions
SSA influences the photocatalytic activity, that is, the at pH = 3 due to the higher stability of the particles
larger the specific surface area, the larger number of according to zeta potential results (Fig. 2). In these
active sites, and consequently, an enhanced photocat- conditions, the nanoparticles have sizes ranging from
alytic activity of the material should be expected. In 10 to 35 nm (measured by DLS, not shown), demon-
this case, we can observe this direct relation for the strating that there is almost no agglomeration of the
materials synthesized in the present work. Moreover, particles as compared with the distribution shown in
it has been reported that the phase structure and Fig. 1, despite the use in that case of a mixture of water
composition, and crystallite size have great influence and ethanol as dispersion medium for the nanoparticles.
on the photocatalytic activity and In order to optimize the dip-coating experiments, a
photoelectrochemical properties of semiconductors preliminary study of the withdrawal rate was performed
(Aguado et al. 2006). The kinetic profiles shown in ranging from 0.5 to 3 mm s−1. Although the experimen-
Fig. 7 point out that crystallization to anatase also tal values are not shown for the sake of simplicity, it was
plays a crucial role, as the sample calcined at 500 °C/ appreciated that the homogeneity of the film was better
1 h consisting only in anatase phase shows a signif- at low withdrawal rates, between 0.8 and 1.1 mm s−1.
icantly higher activity than that calcined at 600 °C/ Accordingly, the thin films were prepared at a withdraw-
1 h despite their SSA is not so different. In this case, al speed of 1 mm s−1. After the drying process (200 °C/
the presence of rutile in the sample calcined at 1 h), they were calcined at 500 °C/1 h in order to get a
600 °C/1 h is presented in Fig. 3. It is worthy to note, dense film with good adhesion to the glass substrate.
however, that commercial TiO2 P25 also consists on The obtained films were characterized in terms of sur-
anatase and rutile phases and has a SSA close to that face texture through AFM observations, as displayed in
of the sample calcined at 600 °C/1 h. However, its Fig. 8. An overview of the thin film topography was
photocatalytic activity is significantly higher, what recorded, showing a roughness ranging from 10 to
some authors have explained as due to a synergy 35 nm. The topography confirms a coating of high
effect between both anatase and rutile phases leading homogeneity with nanoparticle features with granular
to an effective suppression of charge carrier recom- morphology.
bination (Ohno et al. 2001; Hurum et al. 2003). It is To verify the presence of the anatase phase, Raman
therefore plausible to propose that, by contrast to measurements were performed (Fig. 9). It can be seen
TiO 2 P25, in the case of the sample calcined at that the normalized Raman spectrum of the thin film is
600 °C/1 h, defects generated in the phase transition in accordance with that of the powders calcined at
can act as recombination centers of charge carrier 500 °C/1 h, presenting four major bands centered at
pairs, and therefore this might be responsible for its 145, 195, 398, 517, and 640 cm−1 associated to the
lower photocatalytic performance. anatase phase.
Despite the greater photocatalytic activity of the ma- The photocatalytic activity behavior of TiO2 thin
terials when used in the form of powder, the main films was evaluated by the MB degradation under UV
drawbacks of this system are the difficult recovery of light irradiation (Fig. 10). For comparison purposes, a
the photocatalyst. In this context, a plausible alternative thin film of commercial TiO2 P25 was prepared and
investigated in this work is the use of films, where the tested under the same conditions. It can be seen that
nanoparticles were impregnated in the form of a thin the percentages of MB degradation achieved with irra-
film on a glass substrate. diation time along the reaction were significantly higher
23 Page 8 of 10 J Nanopart Res (2018) 20:23

Fig. 8 AFM images (topography channel) of the TiO2 thin film calcined at 500 °C/1 h

than the results of direct photolysis, thus demonstrating to an immobilized catalyst system. Despite this draw-
the photocatalytic efficiency of the thin titania films. In back, as indicated above, the use of supported titania
comparison with the reactions carried out with pow- films avoids the filtration of the catalyst.
dered photocatalysts, it is observed that the rate for As can be observed in Fig. 10, both titania samples
MB degradation with the films is slower, which is an were found to have fairly similar kinetics and, after
expected result since the active interface with the liquid 120 min, almost half of MB was degraded. Taking into
phase is higher in a slurry configuration in comparison account not only the high photocatalytic activity of the
TiO2 P25 sample, which is customarily taken as a
195
145

398

517

640

1.0

0.9
Intensity (a.u.)

TiO2 powders
0.8
C/C0

0.7

0.6
Photolysis
Thin Film TiO2
0.5 TiO2 P25

200 400 600 800 0 20 40 60 80 100 120


-1
Raman Shift (cm ) Irradiation Time (min)
Fig. 9 Raman spectra of the TiO2 thin film and TiO2 powders Fig. 10 Kinetic curves of the methylene blue degradation using
both calcined at 500 °C/1 h the TiO2 and thin films of prepared commercial TiO2 P25
J Nanopart Res (2018) 20:23 Page 9 of 10 23

reference in many photocatalytic experiments, but also References


the differences between both materials in terms of SSA,
particle size and structure, we can conclude that the TiO2 Adjimi S, Sergent N, Roux J et al (2014) Photocatalytic paper
thin film prepared in this work using the TiO2 synthe- based on sol–gel titania nanoparticles immobilized on porous
sized by a colloidal sol-gel enhanced by a microwave silica for VOC abatement. Appl Catal B Environ 154–155:
123–133. https://doi.org/10.1016/j.apcatb.2014.02.011
thermal process, has an excellent photocatalytic activity,
Aguado J, Vangrieken R, Lopez-Muñoz MJ, Marugan JJ (2006) A
comparable to that of titania P25. This evidences the comprehensive study of the synthesis, characterization and
great capabilities of the TiO2 nanoparticles obtained by activity of TiO2 and mixed TiO2/SiO2 photocatalysts. Appl
microwave-assisted green synthesis for applications in Catal A Gen 312:202–212. https://doi.org/10.1016/j.
apcata.2006.07.003
the field of photocatalysis, both in the form of nanopar-
Ashok CH, Rao KV (2017) Synthesis of nanostructured metal
ticles and as thin films. oxide by microwave-assisted method and its humidity sensor
application. Mater Today 4:3816–3824. https://doi.
org/10.1016/j.matpr.2017.02.279
Belver C, Bedia J, Rodriguez J (2015) Zr-doped TiO2 supported
on delaminated clay materials for solar photocatalytic treat-
Conclusions
ment of emerging pollutants. J Hazard Mater 322(Pt A):233–
242. https://doi.org/10.1016/j.jhazmat.2016.02.028
TiO2 nanoparticles with an average crystal size of ~ Bhattacharya M, Basak T (2016) A review on the susceptor
7 nm and a well-defined anatase crystalline phase assisted microwave processing of materials. Energy 97:
without additional heat treatments were successfully 306–338. https://doi.org/10.1016/j.energy.2015.11.034
Borlaf M, Moreno R, Ortiz AL, Colomer MT (2014) Synthesis
obtained by microwave-assisted hydrothermal syn- and photocatalytic activity of Eu3+-doped nanoparticulate
thesis in a simple, soft, fast, and energy-efficient TiO2 sols and thermal stability of the resulting xerogels.
manner in a process of 20 min at a temperature of Mater Chem Phys 144(1-2):8–16. https://doi.org/10.1016/j.
only 180 °C. matchemphys.2013.11.058
Borlaf M, Colomer T, Moreno R, De Andr A (2015) Structural and
Heat treatments at higher temperatures lead to an photoluminescence study of Eu3+/TiO2 xerogels as a function
increase of the average crystal size up to ~ 10 nm of the temperature using optical techniques. J Am Ceram Soc
as observed in the TEM analysis and also to a 345(1):338–345. https://doi.org/10.1111/jace.13299
higher crystallinity of the nanoparticles in the ana- Chen X, Mao SS (2007) Titanium dioxide nanomaterials: synthe-
sis, properties, modifications, and applications. Chem Rev
tase phase (500 °C/1 h). At higher temperatures 107(7):2891–2959. https://doi.org/10.1021/cr0500535
(600 °C/1 h), the transformation of the anatase- Cortéz-Lorenzo A, Escamilla-Perea L, Esquivel-Escalante K,
rutile phase occurs. Velázquez-Castillo R (2017) Modified gelcasting of micro-
The as-synthesized anatase TiO 2 nanoparticles wave assisted synthesized sulfur-doped anatase for photocat-
alytic degradation of organic compounds. Catal Today 282:
showed the highest photocatalytic activity which is 159–167. https://doi.org/10.1016/j.cattod.2016.10.015
comparable to the TiO2 P25. In this context, the pre- Cui Y, He X, Zhu M, Li X (2017) Preparation of anatase TiO2
pared TiO2 thin films (using the as-synthesized and the microspheres with high exposure (001) facets as the light-
TiO2 P25) calcined at 500 °C/1 h also presented a scattering layer for improving performance of dye-sensitized
solar cells. J Alloys Compd 694:568–573. https://doi.
similar kinetic curve for the MB degradation. org/10.1016/j.jallcom.2016.10.032
Delekar SD, Yadav HM, Achary SN, Meena SS, Pawar SH (2012)
Structural refinement and photocatalytic activity of Fe-doped
Acknowledgments The authors thank the resources provided anatase TiO2 nanoparticles. Appl Surf Sci 263:536–545.
by CAPES under the International Cooperation Program Science https://doi.org/10.1016/j.apsusc.2012.09.102
without Borders for Special Guest Researcher, PVE (MEC/MCTI/ Diebold U (2003) The surface science of titanium dioxide. Surf Sci
CAPES/CNPQ/FAP/71/2013), Project No. A011/2013. Rep 48(5–8):53–229. https://doi.org/10.1016/S0167-5729
(02)00100-0
Funding This work was supported by Ministerio de Economía, Eggenhuisen TM, Munnik P, Talsma H, Jongh PE, Jong KP
Industria y Competitividad (Government of Spain) and FEDER (2013) Freeze-drying for controlled nanoparticle distribution
Funds under the Grant No. MAT2015-67586-C3-2-R and in Co/SiO2 Fischer-Tropsch catalysts. J Catal 297:306–313.
CTM2015-69246-R. https://doi.org/10.1016/j.jcat.2012.10.024
Elsellami L, Dappozze F, Fessi N et al (2017) Highly photocata-
Compliance with ethical standards lytic activity of nanocrystalline TiO2 (anatase, rutile) powders
prepared from TiCl4 by sol–gel method in aqueous solutions.
Conflict of interest The authors declare that they have no con- Process Saf Environ Prot 113:109–121. https://doi.
flict of interest. org/10.1016/j.psep.2017.09.006
23 Page 10 of 10 J Nanopart Res (2018) 20:23

Esquivel-Escalante K, Nava-Mendoza R, Velázquez-Castillo R heterojunction perovskite solar cells. Appl Surf Sci 391:2–9.
(2016) Crystal structure determination of the S/TiO2 system https://doi.org/10.1016/j.apsusc.2016.06.171
and the correlation with its photocatalytic properties. J Meng LY, Wang B, Ma MG, Lin KL (2016) The progress of
Nanosci Nanotechnol 16(1):967–972. https://doi. microwave-assisted hydrothermal method in the synthesis
org/10.1166/jnn.2016.11766 of functional nanomaterials. Mater Today Chem 1–2:63–83.
Falk G, Borlaf M, Bendo T, Novaes de Oliveira AP, Rodrigues https://doi.org/10.1016/j.mtchem.2016.11.003
Neto JB, Moreno R (2016) Colloidal sol-gel synthesis and Mirzaei A, Neri G (2016) Microwave-assisted synthesis of metal
photocatalytic activity of nanoparticulate Nb2O5 sols. J Am oxide nanostructures for gas sensing application: a review.
Ceram Soc 99(6):1968–1973. https://doi.org/10.1111 Sensors Actuators B Chem 237:749–775. https://doi.
/jace.14217 org/10.1016/j.snb.2016.06.114
Fan Z, Meng F, Zhang M, Wu Z, Sun Z, Li A (2016) Solvothermal Moreno R (2012) Colloidal processing of ceramics and compos-
synthesis of hierarchical TiO2 nanostructures with tunable ites. Adv Appl Ceram 111(5–6):246–253. https://doi.
morphology and enhanced photocatalytic activity. Appl org/10.1179/1743676111Y.0000000075
Surf Sci 360:298–305. https://doi.org/10.1016/j. Ohno T, Sarukawa K, Tokieda K, Matsumura M (2001)
apsusc.2015.11.021 Morphology of a TiO2 photocatalyst (Degussa, P-25)
Henderson MA (2011) A surface science perspective on TiO2 consisting of anatase and rutile crystalline phases. J Catal
photocatalysis. Surf Sci Rep 66(6-7):185–297. https://doi. 203(1):82–86. https://doi.org/10.1006/jcat.2001.3316
org/10.1016/j.surfrep.2011.01.001
Pinho L, Rojas M, Mosquera MJ (2015) Ag-SiO2-TiO2 nanocom-
Hou Y, Yang J, Jiang Q, Li W, Zhou Z, Li X, Zhou S (2016)
posite coatings with enhanced photoactivity for self-cleaning
Enhancement of photovoltaic performance of perovskite so-
application on building materials. Appl Catal B Environ 178:
lar cells by modification of the interface between the perov-
144–154. https://doi.org/10.1016/j.apcatb.2014.10.002
skite and mesoporous TiO2 film. Sol Energy Mater Sol Cells
155:101–107. https://doi.org/10.1016/j.solmat.2016.05.004 Shen Z, Wang G, Tian H, Sunarso J, Liu L, Liu J, Liu S (2016) Bi-
Huang PJ, Chang H, Yeh CT, Tsai CW (1997) Phase transforma- layer photoanode films of hierarchical carbon-doped brook-
tion of TiO2 monitored by Thermo-Raman spectroscopy with ite-rutile TiO2 composite and anatase TiO2 beads for efficient
TGA/DTA. Thermochim Acta 297(1-2):85–92. https://doi. dye-sensitized solar cells. Electrochim Acta 216:429–437.
org/10.1016/S0040-6031(97)00168-8 https://doi.org/10.1016/j.electacta.2016.09.047
Hurum DC, Agrios AG, Gray KA, Rajh T, Thurnauer MC (2003) Song M, Bian L, Zhou T, Zhao X (2008) Surface potential and
Explaining the enhanced photocatalytic activity of Degussa photocatalytic activity of rare earths doped TiO2. J Rare
P25 mixed-phase TiO2 using EPR. J Phys Chem B 107(19): Earths 26(5):693–699. https://doi.org/10.1016/S1002-0721
4545–4549. https://doi.org/10.1021/jp0273934 (08)60165-9
Jiang H, Liu Y, Zang S, Li J, Wang H (2015) Microwave-assisted Sun S, Zhang J, Gao P, Wang Y, Li X, Wu T, Wang Y, Chen Y,
hydrothermal synthesis of Nd, N, and P tri-doped TiO2 from Yang P (2017) Full visible-light absorption of TiO2 nano-
TiCl4 hydrolysis and synergetic mechanism for enhanced tubes induced by anionic S22− doping and their greatly en-
photoactivity under simulated sunlight irradiation. Mater hanced photocatalytic hydrogen production abilities. Appl
Sci Semicond Process 40:822–831. https://doi.org/10.1016 Catal B Environ 206:168–174. https://doi.org/10.1016/j.
/j.mssp.2015.07.069 apcatb.2017.01.027
Kim HK, Mhamane D, Kim MS, Roh HK, Aravindan V, Madhavi Tauc J (1970) Absorption edge and internal electric fields in
S, Roh KC, Kim KB (2016) TiO2-reduced graphene oxide amorphous semiconductors. Mater Res Bull 5(8):721–729.
nanocomposites by microwave-assisted forced hydrolysis as https://doi.org/10.1016/0025-5408(70)90112-1
excellent insertion anode for Li-ion battery and capacitor. J Tompsett GA, Bowmaker GA, Cooney RP, Metson JB, Rodgers
Power Sources 327:171–177. https://doi.org/10.1016/j. KA, Seakins JM (1995) The Raman spectrum of brookite,
jpowsour.2016.07.053 TiO2 (Pbca, Z = 8). J Raman Spectrosc 26(1):57–62.
Kobayashi M, Tomita K, Petrykin V, Yin S, Sato T, Yoshimura M, https://doi.org/10.1002/jrs.1250260110
Kakihana M (2007) Hydrothermal synthesis of nanosized Weon S, Choi J, Park T, Choi W (2017) Freestanding doubly open-
titania photocatalysts using novel water-soluble titanium ended TiO2 nanotubes for efficient photocatalytic degrada-
complexes. Solid State Phenom 124–126:723–726. tion of volatile organic compounds. Appl Catal B Environ
https://doi.org/10.4028/www.scientific.net/SSP.124-126.723 205:386–392. https://doi.org/10.1016/j.apcatb.2016.12.048
Kosmulskil M (1992) Zeta potential of anatase (TiO2) in mixed Zhang X, Li D, Wan J, Yu X (2016) Hydrothermal synthesis of
solvents. Colloids Surf 64(1):57–65. https://doi.org/10.1016 TiO2 nanosheets photoelectrocatalyst on Ti mesh for degra-
/0166-6622(92)80162-U dation of norfloxacin: influence of pickling agents. Mater Sci
Li H, Shen X, Liu Y, Wang L, Lei J, Zhang J (2016) Titanate Semicond Process 43:47–54. https://doi.org/10.1016/j.
nanowire as a precursor for facile morphology control of mssp.2015.11.020
TiO2 catalysts with enhanced photocatalytic activity. J Zhu XH, Hang QM (2013) Microscopical and physical character-
Alloys Compd 687:927–936. https://doi.org/10.1016/j. ization of microwave and microwave-hydrothermal synthesis
jallcom.2016.05.320 products. Micron 44:21–44. https://doi.org/10.1016/j.
Liang C, Wu Z, Li P, Fan J, Zhang Y, S G (2016) Chemical bath micron.2012.06.005
deposited rutile TiO2 compact layer toward efficient planar

You might also like