Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Journal Pre-proofs

A Study on Titanium Dioxide Nanoparticles Synthesized from Titanium Iso-


propoxide Under SILAR-Induced Gel Method: Transition from Anatase to
Rutile Structure

Agnes C. Nkele, Ugochi K. Chime, Leonard Asogwa, Assumpta C. Nwanya,


U. Nwankwo, K. Ukoba, T.C. Jen, M. Maaza, Fabian I. Ezema

PII: S1387-7003(19)30840-8
DOI: https://doi.org/10.1016/j.inoche.2019.107705
Reference: INOCHE 107705

To appear in: Inorganic Chemistry Communications

Received Date: 17 August 2019


Revised Date: 12 November 2019
Accepted Date: 27 November 2019

Please cite this article as: A.C. Nkele, U.K. Chime, L. Asogwa, A.C. Nwanya, U. Nwankwo, K. Ukoba, T.C. Jen,
M. Maaza, F.I. Ezema, A Study on Titanium Dioxide Nanoparticles Synthesized from Titanium Isopropoxide
Under SILAR-Induced Gel Method: Transition from Anatase to Rutile Structure, Inorganic Chemistry
Communications (2019), doi: https://doi.org/10.1016/j.inoche.2019.107705

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier B.V.


A Study on Titanium Dioxide Nanoparticles Synthesized from Titanium Isopropoxide Under
SILAR-Induced Gel Method: Transition from Anatase to Rutile Structure.

Agnes C. Nkelea, Ugochi K. Chimea, Leonard Asogwaa, Assumpta C. Nwanyaa,b,c, U. Nwankwoa,d, K.


Ukobae, T.C. Jene, M. Maazab,c, Fabian I. Ezemaa,b,c,f*
a
Department of Physics and Astronomy, University of Nigeria Nsukka, Nigeria
b
Nanosciences African Network (NANOAFNET), iThemba LABS-National Research Foundation, 1 Old
Faure Road, Somerset West, 7129, P.O. Box 722, Somerset West, Western Cape Province, South Africa
c
UNESCO-UNISA Africa Chair in Nanosciences/Nanotechnology, College of Graduate Studies,
University of South Africa (UNISA), Muckleneuk ridge, P.O. Box 392, Pretoria, South Africa
d
Department of Physics/Geology/Geophysics, Alex Ekwueme Federal University Ndufu-Alike, Ikwo,
Ebonyi State, Nigeria
e
Department of Mechanical Engineering Science, University of Johannesburg Auckland park
Johannesburg.
f
Department of Physics, Faculty of Natural and Applied Sciences, Coal City University, Enugu, Nigeria

ABSTRACT

Successive ionic layer adsorption and reaction (SILAR)-induced gelling method was adopted in
synthesizing titanium dioxide nanoparticles, annealed at various temperatures so as to investigate
their morphological, structural, optical, elemental, chemical bond and photoluminescence
properties. Characterizations of the as-prepared nanoparticles were carried out using scanning
electron microscope (SEM), X-ray diffractometer (XRD), transmission electron microscopy
(TEM), selected area electron diffractometer (SAED), energy dispersive X-ray spectroscopy
(EDX), UV-Visible spectrophotometer (UV-Vis), photoluminescence (PL) and fourier transform
infrared (FTIR) spectroscopies. Nanospherical balls revealing anatase and rutile crystal
structures at (101) and (110) planes respectively were observed. Agglomerations of chain-like
small particles manifested in Debye-Scherrer’s rings were evident from the TEM and SAED
patterns. EDX spectra revealed deposition of the major elemental constituents: Ti and O. High
transmittance of about 80% with a band gap energy ranges for the anatase and rutile phases was
obtained from the optical properties. Emission peaks arising from the PL spectra gave
information on the charge transport and recombination rates occurring at the TiO2 nanoparticles
while FTIR studies revealed the chemical vibrational bonds of the deposited TiO2 NPs. The
obtained results make the deposited nanoparticles suitable for solar cell applications.

Keywords: SILAR-induced gel; titanium dioxide; nanoparticles; band gap; solar cell.
1. INTRODUCTION

Titanium (iv) oxide commonly called titanium dioxide or titania with formula TiO2 is the
natural form of titanium. TiO2 nanoparticles amongst other metallic oxide nanoparticles like
ZrO2, Al2O3 etc., are industrially useful as electronic devices, catalysts, mechanical materials and
pigments [1]. Its large surface area is responsible for less atoms seen in the bulk material than at
the grain boundaries. Titania is an efficient photocatalyst, it maintains its photostability over
time, toxic-free and has high oxidizing power [2]. It exists naturally in brookite, rutile and
anatase mineral forms. The rutile crystallographic phase exhibits more stability and density than
the anatase or brookite phases. The interface between rutile and anatase improves photocatalytic
property by increasing the separation of charge carriers [3]. TiO2 as a thin film serves as a good
reflective optical coating because of its high refractive index, as electron transporting layer and
can be used in dye sensitized solar cells (DSSCs), perovskite solar cells etc [3]. TiO2 as a
multifunctional material has found applications in many areas due to its features like high
transmittance in the near-infrared and visible region, photocatalytic splitting of water and
outstanding electrical (high resistivity and dielectric constant) properties [2,4]. However, TiO2 in
nanoparticles form rather than in compact thin film are more advantageous because it minimizes
the recombining effects of electron-hole pairs, increases its active surface area and maintains the
band gap energy for particles greater than 3 nm [5]. Parameters such as the surface density,
specific surface area, stoichiometry, synthesis method and crystalline structure are considered
during synthesis in order to achieve optimum benefits from the nanoparticles. Titanium dioxide
films have been synthesized via various methods like atomic layer epitaxy [6], pulsed laser
deposition [7–9], electrophoretic method [10], thermal evaporation [11], dc magnetron sputtering
[5,12,13], chemical bath deposition (CBD) [14,15], reactive r.f. magnetron sputtering [16,17],
ion beam enhanced deposition [18], calcination process [19], liquid phase deposition [2,20],
spray pyrolysis [21,22], sol-gel [23–25], chemical vapor deposition [26,27], successive ionic
layer adsorption and reaction (SILAR) [28–30] while synthesizing titanium dioxide particles
have been possible via hydrolysis [1], alcohothermal [31], simple polymer gel technique [32],
combined sol-gel ball milling method [33], atmospheric pressure chemical vapor synthesis
process [34] etc . SILAR-induced gelling method was adopted because of its simplicity, no
special equipment required, relatively cheap and ease in film thickness control, especially as it
applies to optical devices. TiO2 in anatase and rutile form have been used in protective layers of
electronics devices [26], gas sensing material [25], photovoltaics [29], surface coatings to
minimize wears [15], as an electric insulator [14], anti-reflective layer [23] and in optical
coatings [23] applications.

The structural phase transformation of anatase-rutile requires more studies for


technological advancements [35,36]. Both phases have peculiar properties, the anatase phase is a
potential material for photocatalytic applications while the rutile phase is used in medicine [10].
Chae et al. produced TiO2 nanoparticles by hydrothermal method under an ethanol-rich
condition. Varying the Ti precursor concentration produced particles within the sizes of 7-25 nm.
The resulting nanoparticles exhibited anatase and rutile phases, maintained long-term stability
and was used to fabricate ultra-transparent films of TiO2 [37]. Dreesen et al. synthesized TiO2
nanoparticles by reactive DC magnetron sputtering on carbon substrates. It was observed that the
deposition time affected the crystalline structure, surface coverage and the size of the
nanoparticles. Electron diffraction analysis revealed rutile and anatase phases having a mean
diameter of 19 nm as seen from the SAED results. The particles deposited at lower times
exhibited lower surface coverage with interplanar distances corresponding to the diffraction
planes [5]. Mahshid, Askari and Ghamsari synthesized TiO2 nanoparticles through hydrolyzing
and peptizing titanium isopropoxide solution at 60-70 oC for about 20 hrs. The obtained powder
had anatase and rutile structures with an average nanoparticle size of 28 nm. Spherical
morphology with minimal agglomeration was observed with the TEM and SAED results
corresponding to the SEM and XRD results. Increasing the annealing temperature, increases the
particle size [38]. Karthick et al. synthesized TiO2 nanoparticles via simple polymer gel method
and obtained anatase structure with particle sizes within 7-12 nm range. Below 400 nm,
maximum absorption was observed with an optical band gap energy of 3.55 eV and a maximum
photoluminescence peak of 468 nm [32]. In the current studies, we have used different synthesis
method and deposition conditions. We reported an optical band gap modification mediated by
structural transformation and deposition method adopted in this studies, which were not taken
into account in the previous reports.

This work presents the morphological, structural, elemental and optical features of
titanium dioxide nanoparticles synthesized via SILAR-induced gelling method at annealing
temperatures of 300 oC, 500 oC and 600 oC for 2 hours. The synthesis method was adopted to
regulate its sizes and encourage large scale production. It does not require the addition of extra
solvent was introduced unlike in the previous reports so as to avoid contaminating the deposited
layers and altering desired results. Alternatively, TiO2 nanoparticles instead of thin films have
been used so as to minimize recombination effects, increase surface area activity and increase
direct transition of the anatase phase.

2. EXPERIMENTAL DETAILS

TiO2 nanoparticles were synthesized via SILAR-induced gelling method on glass substrates and
annealing the obtained films at 300 oC, 500 oC and 600 oC respectively for 2 hours. The room-
temperature SILAR-induced method comprised of four (4) steps;

Step 1: In synthesizing the cationic solution, 4 ml of titanium isopropoxide (Ti[OCH(CH3)2]4)


was mixed in 25 ml of isopropanol (CH3CHOHCH3) in a beaker. 0.35 ml of hydrochloric (HCl)
acid was mixed with 25 ml of isopropanol in another beaker. The solutions in both beakers were
mixed together and stirred for about 10 minutes.

Step 2: 100 ml of distilled water was poured in a beaker for rinsing film obtained from step 1

Step 3: In synthesizing the anionic solution, 0.01 M of sodium hydroxide (NaOH) solution was
prepared and put in one beaker.

Step 4: 100 ml of distilled water in a beaker for rinsing film obtained from step 3

Cyclic Process: The titanium dioxide nanoparticles were synthesized following the described
SILAR-induced procedures. The substrate was dipped in the cationic solution for 20 seconds for
adsorption of the titanium species where the gelation process starts. Rinsing of substrate in
distilled water was done for 5 seconds to remove electrons that were loosely bound. The
substrate was dipped in the anionic solution for 20 seconds and afterward rinsed in another
beaker of distilled water for 5 seconds. The process continued until after 23 cycles where the
titanium solution congealed and later dried in a closed chamber for some hours. The synthesized
nanoparticles were annealed respectively at 300 oC, 500 oC and 600 oC for 2 hours. The synthesis
procedure is illustrated in figure 1.
Figure 1: Synthesis processes involved in forming the TiO2 nanoparticles.

The as-synthesized nanoparticles were characterized using the following techniques; the
morphological properties were gotten using a Zeiss scanning electron microscope while the
structures were recorded with X-ray diffractometer (Bruker AXS D8 diffractometer) at a
scanning angle range from 10–90 o with Cu-Kα radiation of λ = 1.5406 Ǻ. The average particle
sizes were measured using Image J software. The elemental compositions were studied by
energy dispersive X-ray (EDX) spectrometer while the optical properties were obtained using a
QEP00503 spectrometer within a wavelength range of 300–2500 nm and 300-1200 nm
respectively. Transmission electron microscopy (TEM) and selected area electron diffraction
(SAED) studies were carried out using Tecnai G2 F20 S-Twin High Resolution Transmission
Electron Microscope (HRTEM) operated at 200 kV, the chemical vibrational bonds of the
nanoparticles were studied from the fourier transform infrared (FTIR) spectrum using Thermo-
Nicolet 8700 FTIR spectrometer while the photoluminescence (PL) properties were obtained
from F-7000 FL spectrophotometer model at a chopping speed of 40 Hz.
3.0 RESULTS AND DISCUSSION

3.1 Morphological Studies

Figure 2(a-c) shows the morphological surface of TiO2 nanoparticles synthesized and
annealed at different annealing temperatures for 2 hours. The surface nature of the samples
reveals spherical balls of different sizes with dimensions in the nanometer range.
Agglomerations were also observed from the SEM images. The sample annealed at 300 °C looks
more aggregated than that at 500 °C from SEM because of agglomeration which increased the
grain size and reduces the aggregation at higher temperatures. The nanoparticles of different
sizes spread uniformly across the surface for the purely anatase phase. This was due to the
nucleation and coalescence encountered during the particles formation process [29]. The
observed surface structures encourages photocatalytic activity as more species would be
adsorbed on the surface per unit area [5].
Figure 2: (a-c) SEM images and (d-f) Grain size distribution of the TiO2 nanoparticles produced
respectively at a) 300 oC b) 500 oC c) 600 oC

The grain size distribution obtained in figure 2(d-f) shows that majority of the particle sizes
ranged between 40 to 60 nm, however, the sample annealed at 600 oC (figure 2(f)) reveals a
wider range of particle sizes. The average grain sizes of the particles were also found to increase
with annealing temperature due to improved crystallinity and reduction of interfacial surface
energy during the annealing process. It is observed that increased annealing temperature
produced larger grain sizes and smaller voids present. The small voids obtained at higher
temperatures increase the oxidation rate of the particles by allowing direct passage of oxygen
molecules inward from the ambient surrounding [12]. Annealing the as-synthesized TiO2
enhances the adsorption of oxygen so that the rutile and anatase phases of the TiO2 nanograins
could be formed [11].

3.2 Structural Studies

Figure 3a reveals the X-ray diffractograms of the TiO2 particles produced at various annealing
temperatures with the corresponding indices of the different diffraction planes where R and A
represent rutile and anatase respectively. The particles exhibited two phases: anatase and rutile.
However, the particles synthesized at 300 oC manifested the anatase phase, the particles annealed
at 500 oC exhibited both anatase and rutile phases while the particles annealed at 600 oC showed
a purely rutile phase. An increase in the diffraction intensity as the annealing temperature
increased is an indication of crystallite growth and improved crystalline nature of the samples
[2]. An increase in the magnitude of the diffraction peaks at increasing temperature could be
because the hydroxyl radicals (OH-) were adsorbed better into the materials synthesized at higher
temperatures and with larger grain sizes [21]. At 300 oC, the low intensities of the diffraction
peaks were due to the transformation from amorphous to anatase phase that required sufficient
temperature in initiating the crystallization process [2,39].

Figure 3: (a) XRD patterns of the TiO2 particles at various annealing temperatures (b) Zoomed
XRD patterns showing the most prominent rutile (R(110)) anatase (A(101)) peaks.
The XRD patterns in Figure 3 reveal tetragonal crystal structures of the anatase and rutile
TiO2 phases having the JCPDS card numbers: 00-021-1276 and 00-021-1272 respectively. Upon
increasing the temperature beyond 300 oC, (110) plane was prominently observed. The (110)
plane of the stable rutile tetragonal phase at 600 oC exhibited the most intense peak. The anatase
phase of TiO2 has been reported to have better stability and is therefore more suitable for
photocatalytic activity [40].

From Figure 3b, as there is no apparent rutile phase for the particles annealed at 300 oC,
hence its rutile structure content is assumed to be 0%. Similarly, for the particles annealed at 600
o
C, there is no anatase phase present in the xrd patterns and the rutile structure content is
assumed to be 100%. However, sample annealed at 500 oC, the anatase and rutile contents (fA)
and (fR) in the materials were obtained to be 50.8% and 49.2% respectively using a formula
proposed by Spurr and Myers [41] in equation (1),

1
fA  (1a)
1  1.26I R I A 

f R 1  f A (1b)

where IA and IR are the XRD intensities for A(101) and R(110) structures respectively.

This indicates that by increasing the annealing temperature of TiO2, the amorphous-to-anatase-
to-rutile transition was observed. This observed transitions are in agreement with Porter et al.
studies on calcination temperature effect on TiO2 structure [42].

The XRD results corresponds with the SEM results that as the annealing temperature is
increased the better is the crystalline quality and hence increases the crystallite number. The
average crystallite sizes of the nanoparticles for the prominent peaks were calculated using
Scherrer’s formula [43] in equation (2):

0.94
D (2)
 cos

Where β is the full width at half maximum (FWHM), θ is Bragg’s diffraction angle, λ is the
wavelength of the X-ray target.
Crystal defects in the form of grain boundaries dislocations existed due to the polycrystalline
nature of the titania nanoparticles. The lattice constants: a and c of the TiO2 nanoparticles with
other structural parameters are listed in Table 1. The inter-planar distance, d, between parallel
atomic planes and the dislocation density, δ, were obtained using equations (3) and (4) [44]:

d  (3)
2 sin 
1
 (4)
D2

Table 1: Summary of some crystal parameters obtained from the TiO2 nanoparticles

Annealing Crystallite FWHM D (nm) d (nm) δ (nm-2) a (Ǻ) c (Ǻ)


Temperatures (oC) Phases (o)

300 A(101) 1.4266 5.96 0.3502 0.0282 0.4044 0.7005


500 A(101) 0.5887 14.44 0.3502 0.0048 0.4044 0.7005
500 R(110) 0.3217 26.57 0.3263 0.0014 0.3768 0.6526
600 R(110) 0.2506 34.05 0.3275 0.0009 0.3781 0.6549

From Table 1, nanoparticles annealed at 300 oC possess the lowest calculated average crystallite
size using equation (2), and this is in agreement with observed SEM results. The dislocation
density of a material is inversely proportional to the crystallite size and is confirmed from the
obtained results. The lattice constants correspond to the anatase and rutile phases of TiO2
nanoparticles.

Transmission electron microscope (TEM) images and selected area electron diffraction
(SAED) modes were studied for the nanoparticles synthesized at different temperatures. TEM
and SAED results help to understand the morphological and crystallographic structures of the
nanoparticles. Figure 4(a-b, d-e) and figure 4(c, f) describe the anatase and rutile phase of the
TiO2 nanoparticles respectively. The TEM patterns revealed agglomerations of small particles
and chain-like crystallite aggregates viewed at the nanometer range. The different structures
observed from the TEM images account for the different phases: anatase and rutile observed in
the XRD results. The interplanar spacing of the different samples revealed the anatase and rutile
structures and are indexed in the TEM patterns. The large surface area encompassing the
nanospherical grains make the titania particles potential photoelectrochemical electrodes [29].
The SAED patterns in the figure revealed Debye-Scherrer rings with each ring representing a
given (h k l) reflection of either the anatase or rutile crystalline phases of the titania
nanoparticles. The rings confirm the nanocrystalline state of TiO2 particles. Evolution of
crystalline grain phases due to annealing is evident from the diffracted rings arrangements.
Transformation from anatase to rutile phase was evident from the spots which became brighter
with increasing temperature. The bright spots show preferential diffraction features while the
high intensity of the spots show that it came from the nanoparticle core [5]. The spots also reveal
the random orientation of the crystallites which correspond to different diffraction planes. The
interplanar distances are indicated by rings which correspond to the lattice planes observed from
the XRD patterns.
a d

(101) , d= 3.52 Å

b e

(101) , d= 3.52 Å
(111) , d= 2.19 Å

c f

(111) , d= 2.19 Å

Figure 4: (a-c) TEM images and (d-f) SAED patterns of the TiO2 nanoparticles produced
respectively at a) 300 oC b) 500 oC c) 600 oC

3.3 Elemental Composition

Energy dispersive X-ray spectroscopy (EDX or EDS) was adopted in determining the constituent
elements of the synthesized TiO2 nanoparticles. The EDX spectra for the TiO2 nanoparticles
annealed at different temperatures are displayed in Figure S1. The major elements: titanium (Ti)
and oxygen (O) were obtained and confirmed the deposition of the expected film. The EDX
investigations confirmed the formation of TiO2 nanoparticles upon deposition on the substrate
surfaces.

3.4 Optical Studies

Figure 5 reveals the absorbance, transmittance and reflectance optical spectra of the TiO2
particles annealed at different temperatures within the wavelength region 300-2500 nm. The
spectra showed interference fringes throughout the wavelength range. Low absorbance observed
within the visible electromagnetic region in Figure 5(a) is the characteristic of TiO2. Maximum
transmittance of about 80% was obtained in Figure 5(b) at the highest annealing temperature.
The lower transmittance property of the particles at annealing temperatures of 300 oC and 500 oC
could be due to sufficient oxygen vacancies that absorbed the incident light [18]. Exciting
electrons from the valence to conduction band of titania after light absorption contributed to the
low transmittance exhibited below 400 nm. These transmittance properties could be attributed to
the annealing carried out at high temperatures and transformations between anatase and rutile
phases [45]. The rutile titania phase exhibited better optical transmittance. Figure 5(c) shows the
particles deposited at lower temperatures recorded reduced reflectance than that synthesized at
600 oC. Changes in the optical features could be due to slight changes in the crystallite size,
phase structure and surface morphologies of the particles.
T R A N S MIT T A N C E (% )

(a
A B S O R B A N C E (a . u . )

0.04 O 80
(b
R E F L E C T A N C E (a . u . )

300 C 0.1 5 (c
O
500 C 60

O 0.1 0
600 C 40
O
0.02 300 C O
O
300 C
20 500 C 0.05 O
O
500 C
600 C O
600 C
800 1 600 2400 800 1 600 2400
800 1 600 2400
W A V E L E NG T H (n m ) W A V E L E NG T H (n m )
W A V E L E NG T H (n m )

Figure 5: Plots showing the a) absorbance b) transmittance c) reflectance of the TiO2


nanoparticles synthesized at 300 oC, 500 oC and 600 oC.

Figures 6(a) and (b) illustrate the absorption coefficient (α) and (αhν)2 versus photon energy)
plots of the particles synthesized at different temperatures. The plots were gotten from diffuse
reflectance values using Kubelka-Munk theory. The incident photon energy dependence on
absorption coefficient is obtainable from Tauc equation helps in determining the transition nature
of the optical absorption. The particles synthesized at 300 oC and 500 oC demonstrated lower
absorption coefficients than that synthesized at 600 oC. All the particles recorded higher
absorption coefficients at the higher photon energy regions are shown in Figure 6(a).

The Kubelka-Munk (K-M) equation is given in equation (5a) while the optical band gap energy
(Eg) and absorption coefficient is related by equation (5):

F(R) = (1-R)2 / 2R (5a)

(h )1 n  A(hv  Eg ) (5b)

Where F(R) is the K-M function,  is the absorption coefficient, R is the diffuse reflectance,
h is photon energy, A is a factor that is dependent on the transition probability, h is Planck’s
constant, while the power n is dependent on the transition during the absorption process. n has
values of 1/2, 3/2, 2, and 3 for the direct allowed, direct forbidden, indirect allowed, and indirect
forbidden transition respectively. The nanoparticles annealed at 300 oC, 500 oC and 600 oC
recorded optical band gap energies of 2.62 eV for the anatase phase and 3.05 eV and 2.98 eV for
the rutile phase as seen in Figures 6(b-c). The anatase phase recorded lower band gap energies
than the rutile phase because of its structural composition. The anatase and rutile phases
exhibited the indirect allowed and direct allowed transitions respectively. The band gap energies
of the anatase and rutile phases of the nanoparticle synthesized at 500 oC were different owing to
their different phase structures. Reduction of band gap energies with annealing temperature has
been attributed to a reduction in interatomic spacing and structural modifications [12,16]. Similar
band gap energies have also been obtained [14,29]. The higher band gap energies obtained could
be due to quantum size effect [29] and the thermal stress encountered during the annealing
process [7]. The reduction of the band gap energies at increasing temperatures justifies the
increase in grain sizes as seen from the SEM images.
Figure 6: Optical plots showing the a) absorption coefficient and band gap energy of the b)
anatase and c) rutile phase of the TiO2 nanoparticles synthesized at various temperatures.

3.5 Fourier transform infrared spectroscopy (FTIR)

FTIR studies were carried out to investigate the vibrational bonds and chemical structure existing
on the TiO2 nanoparticles. The FTIR spectra of the TiO2 nanoparticles annealed at 300 oC, 500 oC
and 600 oC are illustrated in Figure 7. The nanoparticles display main bands at 400-700 cm-1 that
belong to Ti-O stretching and Ti-O-Ti bridging stretching modes [2,46]. The 1400-1750 cm-1
regions belong to O-H bending vibrations [2]. The broad bands at 2800-3800 cm-1 represent the
O-H vibration for water molecules while the broad band at 3100-3600 cm-1 belongs to Ti-OH [2].
The O-H group was evident because of water molecule absorption from the aqueous solution
during the synthesis process. The nanoparticles synthesized at higher temperature recorded
higher transmittance percentages. The different phases: anatase and rutile were evident from the
spectrum displayed by each sample. These results agree with the X-ray diffractogram. The
intensities of O-H vibration band reduced while that of the Ti-O band increased with annealing
temperature. The results confirm the formation of titanium nanoparticles.
T R A NS MIT T A NC E (% )
60

50

300 o C
500 o C
40
600 o C
1 000 2000 3000 4000
W A V E NU MB E R (c m -1 )

Figure 7: FTIR spectra of the TiO2 nanoparticles synthesized at300 oC, 500 oC and 600 oC.

3.6 Photoluminescence (PL) Studies

The PL emission spectrum is an important tool in determining efficient charge transport and the
rate of recombination of carriers in semiconductors [47]. Some reports on the PL spectra of
titania nanoparticles have been presented in literatures [14,48,49]. Figure 8(a) reveals the
photoluminescence spectra at room temperature for the TiO2 particles annealed at 300 oC, 500 oC
and 600 oC with respect to photon energy. Three major emission peaks were observed at 2.78
eV, 1.77 eV and 1.47 eV for the sample synthesized at 600 oC while the particles synthesized at
300 oC and 500 oC had emission peaks at approximately 2.90 eV and 1.77 eV. The variations in
the photoluminescence intensity were due to a change in defect state of the titania surface and
the different phase structures that arose from different annealing temperatures [50]. The particle
synthesized at 600 oC recorded the maximum emission intensity as compared to the other
samples synthesized at lower temperatures.

An illustration of the proposed excitation and emission process for the sample annealed at 600 oC
is shown in figure 8(b). When light of sufficient energy strikes TiO2 in the valence band, the
electrons get ionized and move to the conduction band leaving holes behind in the valence band.
The ionized electrons in the conduction band drift around and may be trapped by defects
resulting from Ti or O vacancies present inside and below the conduction band. As a result of
these defects, the lattice structure becomes relaxed and this leads to non-radiative relaxation
(NRR) transitions of the electrons from the conduction bands into the electron traps before
finally recombining with the holes in the valence band. This mechanism is accompanied by
emission of photons of energies 2.78 eV, 1.77 eV and 1.47 eV.

Figure 8: (a) Photoluminescence spectra showing the emission peaks (green arrow) and (b)
Illustration of active energy level for the TiO2 nanoparticles annealed at 600 oC.

4.0 CONCLUSION

This work investigated the morphological, structural, elemental, optical and charge transport
properties of TiO2 nanoparticles synthesized by SILAR-induced gelling method, annealed at the
temperatures of 300 oC, 500 oC and 600 oC. SEM results revealed nanospherical grains with
majority of the grain sizes ranging from 40 to 60 nm. Anatase and rutile TiO2 phases were
confirmed from the XRD results with prominent peaks seen at (101) and (110) planes
respectively. The anatase phase produced an optical band gap energy of 2.62 eV while 2.98 eV
and 3.05 eV were obtained for the rutile phase nanoparticles. TEM images showed
agglomerations of spherical particles while SAED measurements confirmed the nanocrystalline
nature of the deposited samples. FTIR studies revealed the chemical vibrational bonds existing
on the TiO2 nanoparticles. PL measurements described the charge transport mechanism occurring
with maximum emission intensity observed for the nanoparticle synthesized at 600 oC. The
anatase and rutile TiO2 phases both exhibited good optical, structural and morphological
properties and would be useful as a potential photocatalyst.

Acknowledgments
ACN (90406558) and FIE (90407830) graciously acknowledge UNISA for Postdoc and VRSP
Fellowship awards respectively. We graciously acknowledge the grant by NCC under contract
number NCC/R&D/UNN/014. Also, we thank Engr. Emeka Okwuosa for the generous sponsorship
of April 2014, July 2016 and July 2018 Conference/Workshops on Applications of
Nanotechnology to Energy, Health & Environment conference.

REFERENCES

[1] K.D. Kim, H.T. Kim, Synthesis of TiO2 nanoparticles by hydrolysis of TEOT and decrease of particle
size using a two-stage mixed method, Powder Technol. 119 (2001) 164–172. doi:10.1016/S0032-
5910(00)00420-4.
[2] J.-G. Yu, H.-G. Yu, B. Cheng, X.-J. Zhao, J.C. Yu, W.-K. Ho, The Effect of Calcination Temperature on
the Surface Microstructure and Photocatalytic Activity of TiO2 Thin Films Prepared by Liquid Phase
Deposition, J. Phys. Chem. B. 107 (2003) 13871–13879. doi:10.1021/jp036158y.
[3] Titanium dioxide, Wikipedia. (2019).
https://en.wikipedia.org/w/index.php?title=Titanium_dioxide&oldid=897416696 (accessed May
21, 2019).
[4] S. Yamabi, H. Imai, Crystal Phase Control for Titanium Dioxide Films by Direct Deposition in
Aqueous Solutions, Chem. Mater. 14 (2002) 609–614. doi:10.1021/cm010559d.
[5] L. Dreesen, J.-F. Colomer, H. Limage, A. Giguère, S. Lucas, Synthesis of titanium dioxide
nanoparticles by reactive DC magnetron sputtering, Thin Solid Films. 518 (2009) 112–115.
doi:10.1016/j.tsf.2009.06.044.
[6] M. Ritala, M. Leskelä, E. Nykänen, P. Soininen, L. Niinistö, Growth of titanium dioxide thin films by
atomic layer epitaxy, Thin Solid Films. 225 (1993) 288–295. doi:10.1016/0040-6090(93)90172-L.
[7] Y. Du, M.-S. Zhang, J. Wu, L. Kang, S. Yang, P. Wu, Z. Yin, Optical properties of SrTiO3 thin films by
pulsed laser deposition, Appl. Phys. A. 76 (2003) 1105–1108. doi:10.1007/s00339-002-1998-z.
[8] E. György, G. Socol, E. Axente, I.N. Mihailescu, C. Ducu, S. Ciuca, Anatase phase TiO2 thin films
obtained by pulsed laser deposition for gas sensing applications, Appl. Surf. Sci. 247 (2005) 429–
433. doi:10.1016/j.apsusc.2005.01.074.
[9] N. Koshizaki, A. Narazaki, T. Sasaki, Preparation of nanocrystalline titania films by pulsed laser
deposition at room temperature, Appl. Surf. Sci. 197–198 (2002) 624–627. doi:10.1016/S0169-
4332(02)00429-4.
[10] T. Moskalewicz, A. Czyrska-Filemonowicz, A.R. Boccaccini, Microstructure of nanocrystalline TiO2
films produced by electrophoretic deposition on Ti–6Al–7Nb alloy, Surf. Coat. Technol. 201 (2007)
7467–7471. doi:10.1016/j.surfcoat.2007.02.016.
[11] D. Manno, G. Micocci, R. Rella, A. Serra, A. Taurino, A. Tepore, Titanium oxide thin films for NH3
monitoring: Structural and physical characterizations, J. Appl. Phys. 82 (1997) 54–59.
doi:10.1063/1.365848.
[12] C.-C. Ting, S.-Y. Chen, D.-M. Liu, Structural evolution and optical properties of TiO2 thin films
prepared by thermal oxidation of sputtered Ti films, J. Appl. Phys. 88 (2000) 4628–4633.
doi:10.1063/1.1309039.
[13] K.-S. Lee, I.-S. Park, Anatase-phase titanium oxide by low temperature oxidation of metallic Ti thin
film, Scr. Mater. 48 (2003) 659–663. doi:10.1016/S1359-6462(02)00561-4.
[14] G. Govindasamy, P. Murugasen, S. Sagadevan, G. Govindasamy, P. Murugasen, S. Sagadevan,
Investigations on the Synthesis, Optical and Electrical Properties of TiO2 Thin Films by Chemical
Bath Deposition (CBD) method, Mater. Res. 19 (2016) 413–419. doi:10.1590/1980-5373-MR-2015-
0411.
[15] A.H. Mayabadi, V.S. Waman, M.M. Kamble, S.S. Ghosh, B.B. Gabhale, S.R. Rondiya, A.V. Rokade,
S.S. Khadtare, V.G. Sathe, H.M. Pathan, S.W. Gosavi, S.R. Jadkar, Evolution of structural and optical
properties of rutile TiO2 thin films synthesized at room temperature by chemical bath deposition
method, J. Phys. Chem. Solids. 75 (2014) 182–187. doi:10.1016/j.jpcs.2013.09.008.
[16] N. Martin, C. Rousselot, D. Rondot, F. Palmino, R. Mercier, Microstructure modification of
amorphous titanium oxide thin films during annealing treatment, Thin Solid Films. 300 (1997) 113–
121. doi:10.1016/S0040-6090(96)09510-7.
[17] M.D. Wiggins, M.C. Nelson, C.R. Aita, Phase development in sputter deposited titanium dioxide, J.
Vac. Sci. Technol. A. 14 (1996) 772–776. doi:10.1116/1.580387.
[18] F. Zhang, Z. Zheng, X. Ding, Y. Mao, Y. Chen, Z. Zhou, S. Yang, X. Liu, Highly oriented rutile-type
TiO2 films synthesized by ion beam enhanced deposition, J. Vac. Sci. Technol. A. 15 (1997) 1824–
1827. doi:10.1116/1.580798.
[19] P. de Almeida, J. van Deelen, C. Catry, H. Sneyers, T. Pataki, R. Andriessen, C. Van Roost, J.M.
Kroon, Microstructure characterization of titanium dioxide nanodispersions and thin films for dye-
sensitized solar cell devices, Appl. Phys. A. 79 (2004) 1819–1828. doi:10.1007/s00339-003-2242-1.
[20] S. Deki, Y. Aoi, O. Hiroi, A. Kajinami, Titanium (IV) Oxide Thin Films Prepared from Aqueous
Solution, Chem. Lett. 25 (1996) 433–434. doi:10.1246/cl.1996.433.
[21] L. Castañeda, J.C. Alonso, A. Ortiz, E. Andrade, J.M. Saniger, J.G. Bañuelos, Spray pyrolysis
deposition and characterization of titanium oxide thin films, Mater. Chem. Phys. 77 (2003) 938–
944. doi:10.1016/S0254-0584(02)00193-1.
[22] D.R. Acosta, A.I. Martínez, A.A. López, C.R. Magaña, Titanium dioxide thin films: the effect of the
preparation method in their photocatalytic properties, J. Mol. Catal. Chem. 228 (2005) 183–188.
doi:10.1016/j.molcata.2004.09.070.
[23] T. Houzouji, N. Saito, A. Kudo, T. Sakata, Electroluminescence of TiO2 film and TiO2:Cu2+ film
prepared by the sol-gel method, Chem. Phys. Lett. 254 (1996) 109–113. doi:10.1016/0009-
2614(96)00284-9.
[24] R.S. Sonawane, S.G. Hegde, M.K. Dongare, Preparation of titanium(IV) oxide thin film photocatalyst
by sol–gel dip coating, Mater. Chem. Phys. 77 (2003) 744–750. doi:10.1016/S0254-0584(02)00138-
4.
[25] C. Garzella, E. Comini, E. Tempesti, C. Frigeri, G. Sberveglieri, TiO2 thin films by a novel sol–gel
processing for gas sensor applications, Sens. Actuators B Chem. 68 (2000) 189–196.
doi:10.1016/S0925-4005(00)00428-7.
[26] H. Ha, M. Yoshimoto, H. Koinuma, B. Moon, H. Ishiwara, Open air plasma chemical vapor
deposition of highly dielectric amorphous TiO2 films, Appl. Phys. Lett. 68 (1996) 2965–2967.
doi:10.1063/1.116370.
[27] I. Djerdj, A.M. Tonejc, M. Bijelić, V. Vranes˘a, A. Turković, Transmission electron microscopy
studies of nanostructured TiO2 films on various substrates, Vacuum. 80 (2005) 371–378.
doi:10.1016/j.vacuum.2005.06.015.
[28] H.M. Pathan, S.-K. Min, J.D. Desai, K.-D. Jung, O.-S. Joo, Preparation and characterization of
titanium dioxide thin films by SILAR method, Mater. Chem. Phys. 97 (2006) 5–9.
doi:10.1016/j.matchemphys.2005.04.009.
[29] U.M. Patil, K.V. Gurav, O.-S. Joo, C.D. Lokhande, Synthesis of photosensitive nanograined TiO2 thin
films by SILAR method, J. Alloys Compd. 478 (2009) 711–715. doi:10.1016/j.jallcom.2008.11.160.
[30] S. Park, E. DiMasi, Y.-I. Kim, W. Han, P.M. Woodward, T. Vogt, The preparation and
characterization of photocatalytically active TiO2 thin films and nanoparticles using Successive-
Ionic-Layer-Adsorption-and-Reaction, Thin Solid Films. 515 (2006) 1250–1254.
doi:10.1016/j.tsf.2006.01.068.
[31] Y. Li, X. Sun, H. Li, S. Wang, Y. Wei, Preparation of anatase TiO2 nanoparticles with high thermal
stability and specific surface area by alcohothermal method, Powder Technol. 194 (2009) 149–152.
doi:10.1016/j.powtec.2009.03.041.
[32] S.N. Karthick, K. Prabakar, A. Subramania, J.-T. Hong, J.-J. Jang, H.-J. Kim, Formation of anatase
TiO2 nanoparticles by simple polymer gel technique and their properties, Powder Technol. 205
(2011) 36–41. doi:10.1016/j.powtec.2010.08.061.
[33] M. Farbod, M. Khademalrasool, Synthesis of TiO2 nanoparticles by a combined sol–gel ball milling
method and investigation of nanoparticle size effect on their photocatalytic activities, Powder
Technol. 214 (2011) 344–348. doi:10.1016/j.powtec.2011.08.026.
[34] M. Rahiminezhad-Soltani, K. Saberyan, F. Shahri, A. Simchi, Formation mechanism of TiO2
nanoparticles in H2O-assisted atmospheric pressure CVS process, Powder Technol. 209 (2011) 15–
24. doi:10.1016/j.powtec.2011.01.021.
[35] G.B. Song, J.K. Liang, F.S. Liu, T.J. Peng, G.H. Rao, Preparation and phase transformation of
anatase–rutile crystals in metal doped TiO2/muscovite nanocomposites, Thin Solid Films. 491
(2005) 110–116. doi:10.1016/j.tsf.2005.05.035.
[36] A. Weibel, R. Bouchet, P. Bouvier, P. Knauth, Hot compaction of nanocrystalline TiO2 (anatase)
ceramics. Mechanisms of densification: Grain size and doping effects, Acta Mater. 54 (2006) 3575–
3583. doi:10.1016/j.actamat.2006.03.033.
[37] S.Y. Chae, M.K. Park, S.K. Lee, T.Y. Kim, S.K. Kim, W.I. Lee, Preparation of Size-Controlled TiO2
Nanoparticles and Derivation of Optically Transparent Photocatalytic Films, Chem. Mater. 15
(2003) 3326–3331. doi:10.1021/cm030171d.
[38] S. Mahshid, M. Askari, M.S. Ghamsari, Synthesis of TiO2 nanoparticles by hydrolysis and
peptization of titanium isopropoxide solution, J. Mater. Process. Technol. 189 (2007) 296–300.
doi:10.1016/j.jmatprotec.2007.01.040.
[39] S. Rajesh Kumar, C. Suresh, A.K. Vasudevan, N.R. Suja, P. Mukundan, K.G.K. Warrier, Phase
transformation in sol–gel titania containing silica, Mater. Lett. 38 (1999) 161–166.
doi:10.1016/S0167-577X(98)00152-9.
[40] L. Chu, Z. Qin, J. Yang, X. Li, Anatase TiO2 Nanoparticles with Exposed {001} Facets for Efficient Dye-
Sensitized Solar Cells, Sci. Rep. 5 (2015) 12143. doi:10.1038/srep12143.
[41] R.A. Spurr, H. Myers, Quantitative analysis of anatase-rutile mixtures with an X-ray diffractometer,
Anal. Chem. 29 (1957) 760–762.
[42] J.F. Porter, Y.-G. Li, C.K. Chan, The effect of calcination on the microstructural characteristics and
photoreactivity of Degussa P-25 TiO2, J. Mater. Sci. 34 (1999) 1523–1531.
[43] A.C. Nkele, U.K. Chime, A.C. Nwanya, D. Obi, R.U. Osuji, R. Bucher, P.M. Ejikeme, M. Maaza, F.I.
Ezema, Role of metallic dopants on the properties of copper (1) iodide nanopod-like structures,
Vacuum. 161 (2019) 306–313. doi:10.1016/j.vacuum.2018.12.049.
[44] O.O. Apeh, U. Chime, S.N. Agbo, S. Ezugwu, R. Taziwa, E. Meyer, P. Sutta, M. Maaza, F.I. Ezema,
Properties of nanostructured ZnO thin films synthesized using a modified aqueous chemical
growth method, Mater. Res. Express. (2018). doi:10.1088/2053-1591/aadcd6.
[45] M.H. Suhail, I.M. Ibrahim, G.M. Rao, Characterization and gas sensitivity of cadmium oxide thin
films prepared by thermal evaporation technique, J. Electron Devices. 13 (2012) 965–974.
[46] A.M. Peiró, J. Peral, C. Domingo, X. Domènech, J.A. Ayllón, Low-Temperature Deposition of TiO2
Thin Films with Photocatalytic Activity from Colloidal Anatase Aqueous Solutions, Chem. Mater. 13
(2001) 2567–2573. doi:10.1021/cm0012419.
[47] X.Z. Li, F.B. Li, C.L. Yang, W.K. Ge, Photocatalytic activity of WOx-TiO2 under visible light irradiation
- ScienceDirect, (2001).
https://www.sciencedirect.com/science/article/abs/pii/S1010603001004464 (accessed June 24,
2019).
[48] B. Zhou, L. Xiao, T.J. Li, Absorption redshift in TiO2 ultrafine particles with surfacial dipole layer,
(1991). https://aip.scitation.org/doi/abs/10.1063/1.106211 (accessed June 24, 2019).
[49] N. Bai, S. Li, H. Chen, W. Pang, Preparation, characterization and photoluminescence properties of
mesolamellar titanium dioxide films, (2001).
https://pubs.rsc.org/en/content/articlelanding/2001/jm/b103930j/unauth#!divAbstract (accessed
June 24, 2019).
[50] F.B. Li, X.Z. Li, Photocatalytic properties of gold/gold ion-modified titanium dioxide for wastewater
treatment, Appl. Catal. Gen. 228 (2002) 15–27. doi:10.1016/S0926-860X(01)00953-X.
GRAPHICAL ABSTRACT
RESEARCH HIGHLIGHTS

 TiO2 nanoparticles after at 300 oC, 500 oC and 600 oC by SILAR-induced gelling method.

 TEM and SAED patterns showed agglomerations of TiO2

 High transmittance percentage with band gap energy ranging from 3.41–3.60 eV.

 The chemical vibrational bonds existing between the nanoparticles and the band-to-band
transitions were evident from the FTIR and PL studies respectively.

You might also like