You are on page 1of 10

Optik - International Journal for Light and Electron Optics 169 (2018) 137–146

Contents lists available at ScienceDirect

Optik
journal homepage: www.elsevier.com/locate/ijleo

Original research article

Synthesis of TiO2 nanoparticles and TiO2-Zeolite composites and


T
study of optical properties and structural characterization

K.M. Alvareza, , J. Alvaradoa, B.S. Sotoa, M.A. Hernandezb
a
Centro de Investigación en Dispositivos Semiconductores, Benemérita Universidad Autónoma de Puebla, Puebla, 72570, Mexico
b
Departamento de Investigación en zeolitas, Benemérita Universidad Autónoma de Puebla, 72570, Puebla, Mexico

A R T IC LE I N F O ABS TRA CT

Keywords: TiO2 is a semiconductor used in photocatalysis due to its optical and electrical properties, while
Titanium oxide Zeolites are used as photocatalysts due to its porous structure that allows adequate adsorption by
Zeolite combining these compounds advanced materials are created. In this work, TiO2 nanoparticles
Optical properties and TiO2-Zeolie compounds are synthesized by the sol-gel method. The composite structure is
Sol-gel
confirmed by using XRD, FTIR and Raman. UV–vis test is used for investigation of TiO2 and TiO2-
Zeolite optical properties. The EDS, SEM and TEM techniques are used for synthesized nano-
composite morphology investigation. The results exhibit that TiO2-Zeolite optical properties in-
creases with the incorporation of Zeolite in TiO2.

1. Introduction

Titanium dioxide (TiO2) is one of the most studied semiconductors due to its non-toxicity, high chemical stability (particularly to
photocorrosion) and photocatalytic efficiency [1]. It is well known that TiO2consists of three crystal forms: anatase, brookite, and
rutile [2]. The anatase phase of titanium dioxide is known to offer a better photocatalyst performance than the brookite or rutile
phases. This is because of its improved charge-carrier mobility and the higher number of surface hydroxyl groups [3]. Also, TiO2 has a
band gap of 3.2 eV in the anatase phase, which limits its photo-excitation to only UV wavelengths [4,5]. It is also one of the most
commonly used metal oxides in industrial applications, being present in sunscreens, paints, ointments, toothpastes, sensors, ceramics
and photocatalysts [6,7]. However, one of the major challenges with TiO2 is its limited ability to adsorb the pollutants on its surface
[8]. The combination of adsorbents and photocatalysts, such as TiO2 nanoparticles, has attracted much attention for the enhanced
catalytic performance of photocatalysts [9].
On the other hand, Zeolites materials are of immense fundamental and industry importance, being widely used, for instance, as
commercial adsorbents and catalysts [10]. Currently, there is a database of more than 201 structures that has been created by the IZA
(International Zeolite Association), which is the agency in charge of assigning a three-letter code to all Zeolite structures that are
being discovered [11]. Then, in order to describe these structures, secondary building units are used, which are composed by clusters
of smaller tetrahedrons.
These secondary building units allow the formation of pores and cavities, and according to the number of tetrahedra that make up
the pores, we can classify the Zeolites as: 1) small pore with channels delimited by rings of 8 tetrahedra (i.e. Erionite Code ERI,
Mordenite Code MOR and Clinoptilolite Code HEU), 2) medium pore with channels delimited by rings of 10 tetrahedra (i.e.
Tetrapropylammonium ZSM-5 Code MFI, ITQ-1 and 3) large pore with channels delimited by rings of 12 tetrahedra (i.e. Faujasite
Code FAU) [12], among which we find Erionite [13], Mordenite [14] and Clinoptilolite [15].


Corresponding author.
E-mail address: wiki.ecko@gmail.com (K.M. Alvarez).

https://doi.org/10.1016/j.ijleo.2018.05.028
Received 20 March 2018; Accepted 8 May 2018
0030-4026/ © 2018 Elsevier GmbH. All rights reserved.
K.M. Alvarez et al. Optik - International Journal for Light and Electron Optics 169 (2018) 137–146

In this work we want to show that it is possible to incorporate the TiO2 into the Zeolites to form hybrid compounds. The Zeolites
used in ERI; MOR and HEU, were chosen because they are small materials in which small size molecules can be captured (e.g.
dimethyl ether, CO2, etc.) and incorporated, which can allow the use of these composites as photocatalysts.

2. Experimental

2.1. Material and methods

2.1.1. Synthesis of TiO2 nanoparticle


Nanoparticles of TiO2 were synthesized by the Sol-Gel method which consists of five steps: i) Mixing: the precursor titanium
isopropoxide (Sigma-Aldrich, Cas N° 546-68-9) was mixed with the solvent isopropanol (J.T. Baker, Cas N° 67-63-0), ii) Gelification:
ammonium hydroxide (EM SCIENCE, Cas N° 1336-21-6) was added to generate the gelation, iii) Aging: The sample was allowed to
stand for about 24 h, vi) Drying: The sample was kept at 80 °C for approximately 6 h in order to remove organic solvents, and v)
Sintering: The nanoparticles were left at a temperature of 500 °C for an approximate time of 1 h for the restructuring of their
networks.

2.1.2. Synthesis of TiO2 nanocomposite


The synthesis of TiO2 with zeolites was carried out in situ, meaning that natural Zeolites were added during the mixing stage of
the TiO2 synthetic process.
In the present work, 6 samples were made, which are constituted by TiO2-Zeolites (ERI, MOR or HEU) and nomenclature 75–25
and 25–75, which indicates its percentage by weight.

3. Results and discussion

3.1. Characterization by X-ray diffraction (XRD)

The intrinsic activity of the Zeolite is a function of the density and strength of the acid sites located in the micropores and the
external area, which are the parameters that can be controlled in the catalyst by varying the Zeolite content in a given amount of
matrix or adjusting the size of the cell unit and the crystal dimensions of the Zeolite.
The nanoparticles of TiO2 obtained by the Sol-Gel method were characterized by XRD as shown in Fig. 1a, where the X-ray
diffraction pattern corresponding to the reference 00-021-1272 (ICCD, 2015), which indicates the presence of the anatase phase in

Fig. 1. X-ray diffraction pattern of a) TiO2, b) TiO2-Erionite, c) TiO2-Clinoptilolite and d) TiO2-Mordenite.

138
K.M. Alvarez et al. Optik - International Journal for Light and Electron Optics 169 (2018) 137–146

TiO2.
Fig. 1b shows the X-ray diffraction pattern of TiO2-Zeolite composite with the presence of Erionite for two different concentra-
tions. It is possible to observe that Ti is always present with the stronger peak at 2θ = 25° in both cases, which can be an indication
that the anatase phase of TiO2 is kept. It is observed that, when adding a small amount of Erionite (25%), the intensity of the pattern
is very small. However, when the concentration of Erionite is increased (75%), the intensity of the pattern increases as well. For both
concentrations a hexagonal phase of the Zeolite is observed and has the reference 00-088-1223 (ICCD,2015) [16].
Furthermore, Fig. 1c shows the difractogram for TiO2-Zeolite composite with Clinoptilolite, which is indicated on the peaks with
the letter C and has the reference 04-016-1257 (ICCD, 2015) for the monoclinic phase. The TiO2-Clinoptilolite composites, as well as
the TiO2-Erionite compounds, have the same behavior; by increasing their Zeolite concentration (75% Clinoptilolite) an increase in
the intensity of their diffraction peaks will be obtained, and vice versa. As the concentration decreases (25%), its intensity decreases,
however, unlike the samples with Erionite, the Clinoptilolite presents both a lower quantity and less intensity of pixels, which
indicates that it is both less polycrystalline and less crystalline.
In the case of Mordenite, which is shown in Fig. 1d as the letter M, the intensity of the peaks also increases in the presence of a
large amount of Zeolite Mordenite (75%). This increase makes the orthombic phase of the Zeolite (reference 029–1790 (ICCD, 2015))
more noticeable in the diffractogram. Mordenite and Erionite both have a large number of peaks, and so are more polycrystalline, in
contrast with Clinoptilite, presenting fewer peaks and, therefore, a lower degree of polycrystallinity. By means of XRD the crystalline
planes of the different phases of the Zeolite were obtained, it was observed that Erionite and Mordenite had more crystalline planes
unlike Clinoptilolite, this means that Erionite and Mordenite are more polycrystalline. However, it is worth mentioning that the
surface area of the amorphous particles is greater than the crystalline ones, this benefits the adsorption process in photocatalysis for
the Erionite and Mordenite composites, though the crystallinity gives stability to the samples.

3.2. Characterization by retro-disperse energy spectroscopy (EDS)

Using the EDS technique, it was possible to determine the elements present in each sample, where we observed that TiO2 (a
semiconductor used for photocatalysis) was present and the components of the Zeolites (mainly contained in aluminislicates and
other salts). TiO2 has been widely utilized due to its ability to adsorb by its active center for catalysis.
Table 1 shows the weight % and atomic % of the different elements found in the TiO2 and TiO2-Zeolite compounds. The Table is
divided into seven categories to show the distribution of Oxygen, Titanium, and all of the additional elements found in the various
TiO2-Zeolite compounds. Aluminisilicates (Al and Si) were detected in each compound in addition to Oxygen, Titanium, and varying
degrees of Na, Ca, Mg, and K. It is also possible to observe that when the amount of TiO2 is increased or decreased, there is a
correlation of Ti and O elements in the EDS. In the particular case of TiO2-Erionite it is observed that when we decrease the amount of
this Zeolite, the presence of Mg and K salts is so small that it cannot be detected. In the case of the TiO2-Clinoptilolite the dis-
appearance of C and Ca is observed when the weight % of the Zeolite decreases, but Mg appears, this is due to the area in which the
sample is analyzed, which is related with the small amount of Zeolite in the composite. In the case of the TiO2-Mordenite sample, it is
seen how the amount of carbon decreases when the Zeolite is reduced, this is because the carbon is present in the Zeolite and by
reducing the amount of Zeolite the detector cannot longer identify the small amounts of C.

3.3. Raman

Raman scattering is a fundamental form of molecular spectroscopy. Together with infrared (IR) absorption, Raman scattering is
used to obtain information about the structure and properties of molecules from their vibrational transitions. The theory of Raman
scattering is more complex than the theory of IR absorption [17]. Some vibrations are only Raman active while others are only IR-
active. Typical examples are found in molecules having a center of symmetry for which the mutual exclusion rule holds. However,
totally symmetric vibrations are always Raman active [18].
The vibrational modes of TiO2 nanoparticles observed in Fig. 2a can find the peaks 143 (symmetrical OeTieO Stretch), 181
(symmetrical OeTieO Stretch), 396 (symmetrical OeTieO Bending Vibration), 516 (anti-symmetrical bending vibration in the
bonds) and 639 nm (symmetrical OeTieO Stretch) [19].
In Fig. 2b the Raman spectrum of the TiO2-Eri composites is shown, where it can be observed that when there is an increase in the
content of Erionite Zeolite the intensity of the peaks decreases. This is due to the fact that Erionite does not possess a vibrational
symmetry resulting in not active in Raman.
Also, the TiO2-Clinoptilolite composites can be observed in Fig. 2 c, which like the TiO2-Erionite composites, it is possible to
observe greater intensity in the sample that contains more TiO2, but in these composites it is possible to distinguish a noise shot, this
is due to fluorescence, unfortunately, the measured Raman spectrum is masked by a strong fluorescence background which might be
useful for some potential applications. The reason for this is that the probability of Raman (cross-sectional) scattering is much lower
than that of fluorescence. In others words, Raman scattering and fluorescence emission are two competing phenomena and the
spectrum is dominated by the most likely phenomenon, which is typically fluorescence, and thus it will induce a continuous back-
ground to the residual spectrum and especially it will increase the photon shot noise degrading the signal-to-noise ratio, which will
result in uncertainly in the case of both material identification and concentration measurements [20]. Among the materials that exist
with fluorescence, we can find uranium (U), fluorite (CaF) and calcite (CaCO3) [21,22].
It is possible to observe in the EDS that the samples that contain TiO2-Clinoptilolite, have elements such as Ca, C and O and that
are characteristic to the calcite, which is a flouroscent compound and that causes shot noises, that is why the Raman spectra of these

139
K.M. Alvarez et al. Optik - International Journal for Light and Electron Optics 169 (2018) 137–146

Table 1
Weight % and atomic % of the composites.
Composites % Weight concentration in the synthesis Element Weight % Atomic %

TiO2 100 % TiO2 O 47.54 73.06


Ti 52.46 26.94
Total 100.00 100.00
TiO2-Erionite (25–75) 25 % TiO2 C 13.11 25.93
75% Erionite Ti 45.49 22.55
O 25.53 37.89
Na 1.38 1.42
Mg 0.49 0.47
Al 2.97 2.61
Si 10.18 8.61
K 0.85 0.52
Total 100.00 100.00
TiO2-Erionite (75–25) 75 % TiO2 C 6.7 15.29
25% Erionite Ti 61.56 35.24
O 24.88 42.64
Na 0.33 0.4
Al 1.26 1.28
Si 5.27 5.14
Total 100.00 100.00
TiO2-Clinoptilolite (25–75) 25 % TiO2 C 9.15 019.18
75% Clinoptilolite Ti 44.94 23.63
O 24.62 38.75
Al 3.39 3.16
Si 14.91 13.37
K 2.05 1.32
Ca 0.95 0.6
Total 100.00 100.00
TiO2-Clinoptilolite (75–25) 75 % TiO2 Ti 45.93 26.07
25% Clinoptilolite O 29.99 50.95
Mg 0.24 0.27
Al 4.19 4.22
Si 17.69 17.12
K 1.96 1.36
Total 100.00 100.00
TiO2-Mordenite (25–75) 25 % TiO2 C 6.83 14.19
75% Mordenite Ti 42.62 22.21
O 27.7 43.21
Al 2.33 2.15
Si 20.53 18.24
Total 100.00 100.00
TiO2-Mordenite (75–25) 75 % TiO2 Ti 69.00 45.69
25% Mordenite O 22.56 44.72
Al 0.95 1.12
Si 7.50 8.47
Total 100.00 100.00

samples have the presence of noise shot.


The shifting of peaks towards lower or higher wave number is specific to the chemical bonds and symmetry of molecules. So, if
chemical bond length of molecules changes due to any internal or external effects (Dopant incorporation), then it may shift the wave
number [23].
Furthermore, in Fig. 2c, the Raman spectrum of the TiO2- Mordenite composites is shown, in which the decrease in the intensity of
the peak and the registration of the bands is observed when the content of the Zeolite is due to the Mordenite not it has a vibrational
symmetry.
This type of analysis helped us identify the phase in which TiO2 (anatase) is found by observing its bands, it is well known that
TiO2 has three (rutile, brokite and anatase), the anatase phase is the most efficient in photocatalytic processes.

3.4. Characterization by Fourier transformed infrared

FTIR absorption spectra of the TiO2 nanoparticles and TiO2-Zeolite composites were performed in order to evaluate the char-
acteristics of the nanoparticles of TiO2 and composites obtaining by Sol-Gel.
The analysis of FTIR spectra for the nanoparticles of TiO2, is depicted in Fig. 3a. Bands can be seen in the wavelengths of 2900,
2120 and 1890 cm−1 which correspond to the links CeH, C^C y CeH. The vibration C]O stretching belongs to 1700 cm−1, the
vibration C]Cstreching to 1500 cm−1, the vibration CeHbending corresponds to the wavelength 1390 cm−1, CeOstrenching corresponds
to vibration in the wavelength of 1020 cm−1[24]. A well-known very strong band of TiO2 can be observed in the range of

140
K.M. Alvarez et al. Optik - International Journal for Light and Electron Optics 169 (2018) 137–146

Fig. 2. Raman Spectroscopy of a) TiO2, b) TiO2-Erionite, c) TiO2-Clinoptilolite and d) TiO2Mordenite.

Fig. 3. FTIR spectra of the TiO2-Zeolite composites.

141
K.M. Alvarez et al. Optik - International Journal for Light and Electron Optics 169 (2018) 137–146

1000–600 cm−1 [25]. Vibrational bands can also be observed around the wavelength of 3430 cm−1 due to eOH groups [26].
In Fig. 3b the bands at 720 and 775 cm−1 are observed at symmetrical stretching vibrations, whereas the bands at 1030 and
1150 cm−1 are related to asymmetric stretching vibrations of TiO2-Erionite composites [27].
The infrared spectra of the TiO2-Clinoptilolite were analyzed in a frequency range of 650–4000 cm−1 and are shown in Fig. 3c.
The vibration frequency peaks at 800, 1090, 1640, 2350 and 3450 cm−1 can be recognized from the absorption spectra of containing
composites Clinoptilolite, which are the typical vibrations of clinoptilolite-based samples. Based on the literature assignments, the
absorption bands appearing around 800 and 1090 cm−1 are assigned to the typical vibration peaks of the aluminisilicates. These
bands are attributed to the symmetrical and asymmetric vibrations of TeOeT [28]. It was observed that when the samples have more
Zeolite, their intensity increased, while the TiO2 peaks decreased.
Furthermore, TiO2-Mordenite composites are shown in Fig. 3d where the absorption bands are observed in 1228, 1080 and
792 cm−1, which are due to TeO flexion, symmetrical internal and external stretching, asymmetric internal and external stretching
vibrations corresponding to Silica materials [29] The characteristic bands with very strong TiO2 signals are found in the range of
1000-600 cm−1 [25]. It is observed that the more TiO2 is aggregated the more intensity peaks belonging to this material are obtained,
and viceversa.
Through the FTIR, the functional groups of the TiO2 and Zeolite samples are determined, which indicate the presence of the
compounds that are important for photocatalysis.

3.5. UV–vis

The importance of the use of the Kubelka-Munk function, F(R), is that allows the use of the measured reflectance, R, to calculate
the optical absorbance. This function is expressed as [29]:
(1−R)2
F (R) =
2R (1)

From the last expression, semi-conducting material can be analyzed with a Tauc-Plot, which Allows the calculation of the bandgap
by absorbance. Whereby [29]:
(F (R)* hV )n (2)

This equation against hV is plotted and n = 0.5 (direct band gap) and 2 (indirect band gap). On the Tauc plot, a linear region just
above the optical absorption edge is fitted and extrapolated to photon energy axis, which is possible to observe in Fig. 4. The intercept
on this axis represents the band gap of the material under investigation [29].
In Fig. 4 we can observe the energies of the band that is obtained in the TiO2 and in the different composites, where in order to the
intercept with the x axis, it is observed that the energies are between the values of 3 eV, in Table 2 are summarized these energies.
In Table 2 the energies of the different composites obtained by the Kubelka-Munk equation are show.
Table 2 shows that the bandgap for TiO2 is 3.05 eV, when Zeolite is added, there is an increase in the bandgap, in the case of the
samples containing the Clinoptilolite and Erionite samples, the more the amount is increased of %weight, the more bandgap increase,
with the exception of Mordenite that decreases, this is due to the impurities presented in the samples.
This study allows determining the bandwidth of the different composites where it was possible to calculate that it is greater than
3 eV, which is in the range of the UV spectrum. This is important because the photocatalytic processes use the energy provided by this
UV wavelength to produce a photocatalytic oxidation and destroy the pollutants.

3.6. Scanning electronic microscopy (SEM)

Fig. 5a shows the micrograph of nano particles of TiO2, where a presence of clusters is observed. Furthermore, Fig. 5b-g show
TiO2-Zeolite composites where the clusters of TiO2 nano particles and the presence of Zeolites with a larger size compared to TiO2 are

Fig. 4. Band gap diagram of the different TiO2-Zeolite composites.

142
K.M. Alvarez et al. Optik - International Journal for Light and Electron Optics 169 (2018) 137–146

Table 2
Band Gap of TiO2 and the different composites.
Sample Energy
(eV)

TiO2 3.05
TiO2-Erionite(25-75) 3.22
TiO2-Erionite(75-25) 3.07
TiO2-Clinoptilolite(25-75) 3.51
TiO2-Clinoptilolite(75-25) 3.10
TiO2-Mordenite (25-75) 3.30
TiO2-Mordenite(75-25) 3.47

Fig. 5. Composites of a) TiO2, b) TiO2-Erionite (25–75), c) TiO2-Clinoptilolite (25–75), d) TiO2-Mordenite (25–75), e) TiO2-Erionite (75–25), f) TiO2-
Clinoptilolite (75–25) y g) TiO2-Mordenite (75–25).

observed.
Fig. 5b shows the TiO2-Erionite (25–75) sample. It is possible to observe some fibrous blocks characteristic of the Erionite Zeolite,
which are quite predominant in comparison to the TiO2 conglomerates being smaller than the Zeolite, in the opposite case when the
sample has more TiO2 as shown in Fig. 5e TiO2-Erionite (75–25), the conglomerates are more visible than the fibrous portions of
Zeolites.
In Fig. 5c related to the sample TiO2-Clinoptilolite (25–75), it is possible to observe some hexagonals blocks characteristic of the
Zeolite Clinoptilolite and they are quite predominant in comparison to the TiO2 conglomerates, which are very small that Zeolite, in
the opposite case when the sample has more TiO2 as shown in Fig. 5f TiO2-Clinoptilolite (75–25), the conglomerates are more visible
than the hexagonal blocks of Zeolites.
In the case of the compound with the highest content of Mordenite (Fig. 5d) TiO2-Mordenite (25–75), small non-uniform particles
are observed and in some parts of the micrograph a mixture of oval –shaped particles is observed and spherical (indicate presence of
TiO2), while in Fig. 5g TiO2-Mordenite (75–25) which presents more amount of TiO2 than Mordenite, a more uniform microscopy is

143
K.M. Alvarez et al. Optik - International Journal for Light and Electron Optics 169 (2018) 137–146

Fig. 6. TEM images of TiO2-Zeolite composites, a) TiO2, b) TiO2-Erionite (75–25), c) TiO2-Clinoptilolite (75–25) and d) TiO2-Mordenite (75–25).

144
K.M. Alvarez et al. Optik - International Journal for Light and Electron Optics 169 (2018) 137–146

observed, i.e. spherical particles of a determined.

3.7. Transmission electron microscopy (TEM)

Electron Transmission Microscopy was performed on the samples containing the least amount of Zeolite, because in SEM it was
observed that the zeolites had a considerable size in comparison to TiO2. Fig. 6a shows the TiO2 nanoparticles where an approximate
size of 10 nm can be distinguished, while in Fig. 6b a mixture of these nanoparticles with blocks higher than 200 nm belonging to the
Erionite Zeolite are observed.
In Fig. 6c belonging to the TiO2-Clinoptilolite (75–25) sample, it is possible to observed conglomerates of the TiO2 nanoparticles
with sizes approximating 10 nm and blocks superior to more than 100 nm, which belongs to the clinoptilolite.
TiO2-Mordenite (75–25) composites are observed in Fig. 6d, where the TiO2 conglomerates and the blocks larger than 100 nm
belonging to the Mordenite can be seen, and the pores of this Zeolite were also observed.

4. Conclusions

In this study, TiO2 nanoparticles was selected due to its due to its optical properties and because TiO2 is known as famous
phtocatalyst. After synthesizing TiO2 nanoparticles by sol-gel method, composites were made in situ with natural zeolites. SEM and
TEM images reveal that TiO2 nanoparticles have been placed on the surface of the block of Zeolites that agree with the XRD and EDS
which determine the presence of these componentes in the sample. Raman and FTIR evaluated the presence of TiO2 in the anatase
phase and TiO2-Zeolite composite and their different phases. The results of UV–vis are related to the Kubelka-Munk method to
determine band interval value of the different compounds that Zeolite is included TiO2 the bandgap increase. Among the applications
of TiO2 and Zeolites according ti their optical properties, photocatalysis can be mentioned.

Acknowledgements

The authors thank the Benemérita Universidad Autónoma de University of Puebla for the flared supports, the center for na-
nosciences and micro and nanotechnologies of the Instituto Politécnico Nacional by the measurements made and financial support to
CONACyT for the realization of this work is appreciated.

References

[1] C.B.D. Marien, T. Cottineau, D. Robert, P. Drogui, TiO2 nanotube arrays: influence of tube length on the photocatalytic degradation of paraquat, Appl. Catal. B
Environ. 194 (2016) 1–6, http://dx.doi.org/10.1016/j.apcatb.2016.04.040.
[2] X. Yao, R. Zhao, L. Chen, J. Du, C. Tao, F. Yang, L. Dong, Selective catalytic reduction of NOx by NH3 over CeO2 supported on TiO2: comparison of anatase,
brookite, and rutile, Appl. Catal. B Environ. 208 (2017) 82–93, http://dx.doi.org/10.1016/j.apcatb.2017.02.060.
[3] R. Fagan, D.E. McCormack, S. Hinder, S.C. Pillai, Improved high temperature stability of anatase TiO2 photocatalysts by N, F, P co-doping, Mater. Des. 96 (2016)
44–53, http://dx.doi.org/10.1016/j.matdes.2016.01.142.
[4] H. Choi, S. Khan, J. Choi, D.T.T. Dinh, S.Y. Lee, U. Paik, S.H. Cho, S. Kim, Synergetic control of band gap and structural transformation for optimizing TiO2
photocatalysts, Appl. Catal. B Environ. 210 (2017) 513–521, http://dx.doi.org/10.1016/j.apcatb.2017.04.020.
[5] E.M. Samsudin, S.B.A. Hamid, Effect of band gap engineering in anionic-doped TiO2photocatalyst, Appl. Surf. Sci. 391 (2017) 326, http://dx.doi.org/10.1016/j.
apsusc.2016.07.007.
[6] I. Zama, C. Martelli, G. Gorni, Preparation of TiO2 paste starting from organic colloidal suspension for semi-transparent DSSC photo-anode application, Mater.
Sci. Semicond. Process. 61 (2017) 137–144, http://dx.doi.org/10.1016/j.mssp.2017.01.010.
[7] H. Choi, E. Stathatos, D.D. Dionysiou, Photocatalytic TiO2 films and membranes for the development of efficient wastewater treatment and reuse systems,
Desalination 202 (2017) 199–206, http://dx.doi.org/10.1016/j.desal.2005.12.055.
[8] R. Nagarjuna, S. Challagulla, N. Alla, R. Ganesan, Synthesis and characterization of reduced-graphene oxide/TiO2/Zeolite-4A: a bifunctional nanocomposite for
abatement of methylene blue, Mater. Des. 86 (2015) 621–626, http://dx.doi.org/10.1016/j.matdes.2015.07.116.
[9] J. Shi, Y. Kuwahara, T. An, H. Yamashita, The fabrication of TiO2 supported on slag-made calcium silicate as low-cost photocatalyst with high adsorption ability
for the degradation of dye pollutants in water, Catal. Today 281 (2017) 21–28, http://dx.doi.org/10.1016/j.cattod.2016.03.039.
[10] R. de Andrade, L. de Sousa, C. Pinto, F. Bohn, R. Ferreira, J.M. Sasaki, A. Rodrigues, Kaolin-based magnetic zeolites A and P as water softeners, Microporous
Mesporous Mater. 246 (2017) 64–72, http://dx.doi.org/10.1016/j.micromeso.2017.03.004.
[11] Database of Zeolite Structures (IZA structure commission, Switzerland, (2013). http://www.iza-structure.org/ (Ch. Baerlocher and L.B. McCusker).
[12] P. Guo, N. Yan, L. Wang, X. Zou, Database mining of zeolite structures, Cryst. Growth Des. 17 (2017) 6821–6835, http://dx.doi.org/10.1021/acs.cgd.
7b01410?src=recsys.
[13] P. Ballirano, A. Pacella, C. Cremisini, E. Nardi, M. Fantauzzi, D. Atzei, A. Rossi, G. Cametti, Fe (II) segregation at A specific crystallographic site of fibrous
erionite: a first step toward the understanding of the mechanisms inducing its carcinogenicity, Microporous Mesporous Mater. 211 (2015) 49–63, http://dx.doi.
org/10.1016/j.micromeso.2015.02.046.
[14] F. de S. Vilhena, R.M. Serra, A.V. Boix, G.B. Ferreira, J. Walkimar de M. Carneiro, DFT study of Li+ and Na+ positions in mordenites and hydration stability,
Comp. Theor. Chem. 1091 (2016) 115–121, http://dx.doi.org/10.1016/j.comptc.2016.07.017.
[15] A. Dziedzicka, B. Sulikowski, M. Ruggiero-Mikolajczyk, Catalytic and physicochemical properties of modified natural clinoptilolite, Catal. Today 259 (2016)
50–58, http://dx.doi.org/10.1016/j.cattod.2015.04.039.
[16] K. Quiroz, M.A. Hernández, R. Portillo, F. Rojas, E. Rubio, V. Petranovskii, Comparación de la capacidad de secuestro de CO2 en Clinoptilolita, Epistilbita,
Erionita, Mordenita y Caolinita como constituyentes del suelo, Superficies y Vacío 29 (2016) 55–61 http://www.redalyc.org/pdf/942/94246523005.pdf.
[17] F. Siedert, P. Hildebrandt, Vibrational Spectroscopy in Life Science, WILEY-VCH. Weinheim, 2008.
[18] J.R. Ferraro, K. Nakamoto, Introductory Raman Spectroscopy, Academic Press, London, 1994.
[19] P.O. Ifeacho, Semi-Conducting Metal Oxide Nanoparticles from a Low-Pressure Premixed H2/O2/Ar Flame: Synthesis and Charaterization, Cuvillier Verlag
Göttingen Internationales wissenschaftlicher Fachverlag, Göttingen, 2008.
[20] I. Nissinen, J. Nissinen, P. Keränen, J. Kostamovaara, On the effects of the time gate position and width on the signal-to-noise ratio for detection of Raman
spectrum in a time-gated CMOS single-photon avalanche diode based sensor, Sens. Actuators B Chem. 241 (2017) 1145–1152, http://dx.doi.org/10.1016/j.snb.
2016.10.021.

145
K.M. Alvarez et al. Optik - International Journal for Light and Electron Optics 169 (2018) 137–146

[21] M. Havelcová, V. Machovic, J. Mizera, I. Sýkorová, M. Rene, L. Borecká, L. Lapcák, O. Bicaková, O. Janecek, Z. Dvorák, Structural changes in amber due to
uranium mineralization, J. Environ. Radioact. 158-159 (2016) 89–101, http://dx.doi.org/10.1016/j.jenvrad.2016.04.004.
[22] T. Omairi, M. Wainwright, Fluorescent minerals – a potential source of UV protection and visible light for the growth of green algae and cyanobacteria in
extreme cosmic environments, Life Sci. Space Res. 6 (2015) 87–91, http://dx.doi.org/10.1016/j.lssr.2015.07.004.
[23] B. Raina, S. Verma, V. Gupta, K.K. Bamzai, Structural, spectroscopic and optical properties of pure and gadolinium doped neodymium phosphate nanoparticles
prepared using wet chemical route, Integr. Ferroelectr. Int. J. 184 (2017) 199–209, http://dx.doi.org/10.1080/10584587.2017.1368661.
[24] M. Gharagozlou, S. Naghibi, Preparation of vitamin B12–TiO2 nanohybrid studied by TEM, FTIR and optical analysis techniques, Mater. Sci. Semicond. Process.
35 (2015) 166–173, http://dx.doi.org/10.1016/j.mssp.2015.03.009.
[25] Y. Zhang, A. Barber, J. Maxted, C. Lowe, R. Smith, T. Li, The depth profiling of TiO2 pigmented coil coatings using step scan phase modulation photoacoustic
FTIR, Prog. Org. Coat. 76 (2013) 131–136, http://dx.doi.org/10.1016/j.porgcoat.2012.08.021.
[26] A. Adamczyk, M. Rokita, The structural studies of Ag containing TiO2–SiO2 gels and thin films deposited on steel, J. Mol. Struct. 1114 (2016) 17–180, http://dx.
doi.org/10.1016/j.molstruc.2016.02.054.
[27] I.R. Lewis, H.G.M. Edwards, Handbook of Raman Spectroscopy From the Research Laboratory to the Process Line, CRC press, U.S.A, 2001.
[28] S. Yang, N.P. Evmiridis, Synthesis and characterization of an offretite/erionite type zeolite, Microporous Mater. 6 (1996) 19–26, http://dx.doi.org/10.1016/
0927-6513(95)00077-1.
[29] S. Ren, C. Gong, P. Zeng, Q. Guo, B. Shen, Synthesis of flammulina-like mordenite using starch as template and high catalytic performance in crack of wax oil,
Fuel 166 (2016) 347, http://dx.doi.org/10.1016/j.fuel.2015.11.010.

146

You might also like