Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

materials

Article
Structural, Morphological, Optical and Photocatalytic
Properties of Y, N-Doped and Codoped TiO2
Thin Films
Zeineb Hamden 1 , David Conceição 2 , Sami Boufi 3 , Luís Filipe Vieira Ferreira 2, * and
Soraa Bouattour 1, *
1 Faculty of Science, LCI, University of Sfax, BP1171-3018 Sfax, Tunisia; zeineb.hamden@yahoo.fr
2 Centro de Química-Física Molecular and Institute of Nanoscience and Nanotechnology,
Instituto Superior Técnico, University of Lisbon, 1049-001 Lisbon, Portugal;
david.conceicao@tecnico.ulisboa.pt
3 Faculty of Science, LSME, University of Sfax, BP1171-3018 Sfax, Tunisia; sami_boufi@yahoo.com
* Correspondence: lfvieiraferreira@tecnico.ulisboa.pt (L.F.V.F.); soraa.boufi@yahoo.com (S.B.);
Tel.: +351-21-841-9252 (L.F.V.F.); +216-98-660-535 (S.B.)

Academic Editor: Klara Hernadi


Received: 11 April 2017; Accepted: 25 May 2017; Published: 31 May 2017

Abstract: Pure TiO2 , Y-N single-doped and codoped TiO2 powders and thin films deposited on glass
beads were successfully prepared using dip-coating and sol-gel methods. The samples were analyzed
using grazing angle X-ray diffraction (GXRD), Raman spectroscopy, time resolved luminescence,
ground state diffuse reflectance absorption and scanning electron microscopy (SEM). According
to the GXRD patterns and micro-Raman spectra, only the anatase form of TiO2 was made evident.
Ground state diffuse reflectance absorption studies showed that doping with N or codoping with N
and Y led to an increase of the band gap. Laser induced luminescence analysis revealed a decrease
in the recombination rate of the photogenerated holes and electrons. The photocatalytic activity of
supported catalysts, toward the degradation of toluidine, revealed a meaningful enhancement upon
codoping samples at a level of 2% (atomic ratio). The photocatalytic activity of the material and its
reactivity can be attributed to a reduced, but significant, direct photoexcitation of the semiconductor
by the halogen lamp, together with a charge-transfer-complex mechanism, or with the formation of
surface oxygen vacancies by the N dopant atoms.

Keywords: titanium dioxide; photochemistry; band-gap engineering; photocatalysis

1. Introduction
Nowadays, organic pollutants produced by some industries are harmful to human health and
living creatures. Owing to the urgent need for a clean and comfortable environment, photocatalysis
offers great potential for the elimination of toxic chemicals in the environment through its efficiency
and broad applicability.
Among many semiconductors and photocatalysts, titanium dioxide TiO2 , in the form of thin films,
powders or nanostructured layers, is close to be an ideal bench mark photocatalyst in the environmental
photocatalysis applications, due to its many desirable properties such as inexpensive, readily available,
biologically and chemically inert, and good photoactivity [1,2]. However, the application of TiO2
is limited by problems associated with the fast charge (electron/hole) recombination phenomenon,
and the band gap belonging to the UV region, since the absorption behavior and the separation
efficiency of electron–hole pairs are two essential factors on which the photodegradation of TiO2
strongly depends [3]. Thus, several approaches have been used in order to overcome these difficulties,

Materials 2017, 10, 600; doi:10.3390/ma10060600 www.mdpi.com/journal/materials


Materials 2017, 10, 600 2 of 13

including doping with other species [4,5]. Between many chemical elements used as dopants for TiO2 ,
rare earth elements such as (Y, Eu, Er, Nd . . . ) have been widely studied [6,7]. Doping TiO2 with rare
earth elements is employed to enhance the photochemical activity regarding degradation of organic
pollutants in aqueous media and to shift the irradiation wavelength from the UV to the visible range.
Narayan et al. [8] have studied the effect of rare earth (Y, Yb, Gd) ions on TiO2 properties using a
co-precipitation/hydrolysis method and its photocatalytic activity was evaluated for the degradation
of Congo red under visible light irradiation. They demonstrated that Y modified TiO2 provided the
best photocatalytic activity due to the smaller particle size of photocatalyst and effective separation of
electron–hole pairs. Wang et al. [9] synthesized also yttrium doped TiO2 through the sol–gel method
and reported that Y-TiO2 has about 1.50 times greater photocatalytic activity compared to undoped
TiO2 for the methyl orange degradation. Moreover, in a previous research work, Bouattour et al. [10],
studied the photocatalytic activity of nanopowders of yttrium doped titania using sol–gel, solid state
process, and 2-naphthol as a pollutant model, under sunlight irradiation. Results showed a great
enhancement of the photocatalytic efficiency with the incorporation of Y in samples prepared by solid
grinding. Kallel et al. [11] has studied Rb-Y codoped TiO2 . They demonstrated that, according to
the XPS and XRD analysis, the Y3+ dopant did not enter the TiO2 crystal lattice to substitute for Ti4+ .
It was dispersed uniformly onto TiO2 nanoparticles. Wu et al. [12] concluded in their studies that the
incorporation of interstitial boron dopants would create oxygen vacancies (Ov¨) and reduce Ti4+ to
Ti3+ to form the [Ov¨-Ti3+ ]+ complex, which would then trap the excited charge carriers and prolong
carrier lifetime. Moreover, Y3+ ions could also trap the photo-excited electrons to form Y2+ ions, which
would then react with the absorbed O2 on TiO2 surface to generate reactive species. On the other hand,
research on the incorporation of non-metal ions (N, C, S, F) into TiO2 has increased since few years [13],
for the reason that doping with these atoms can largely enhance the photoactivity efficiency of TiO2 .
However, there are still many controversies on the modification of TiO2 by these species, especially
the nitrogen, since different hypotheses concerning the state of nitrogen in the N-TiO2 lattice and the
mechanism of band gap modification have been derived. For example, Asahi et al. [4] proposed that
substitutional-type doping using N was effective for the band gap narrowing of TiO2 due to the mixing
of N 2p with O 2p states in the valence band based on spin-restricted local density approximation
calculations on the anatase phase. However, Irie et al. [14] confirmed that interstitial-type doping of
N atoms was related to the photothreshold energy decrease, which induced localized N 2p states within
the band gap just above the top of the valence band, facilitating the production of oxygen vacancies.
A third research group suggest that the band edge narrowing can occur when doping levels reach a
critical concentration that exceeds 20% in anatase TiO2 [13]. From the above-mentioned references,
it appears that the effect of dopants on the structural properties and the photocatalytic activity of TiO2
is a complex problem that depends not only on the dopant nature but also on the preparative procedure
of the materials and the pollutant to degrade. Thus, we have considered that studying codoped TiO2
thin films with N as one of the dopants should be interesting to try to identify the effect of the each
dopant on the properties and TiO2 performance. In this context, we have prepared a series of Li-,
N-doped and Li-N codoped TiO2 thin films and powders [15]. We have demonstrated that the energy
gap increases and the electron/hole recombinations are delayed for all the samples. An Li-N codoped
sample exhibits interesting photoactivity compared to undoped and Li-N monodoped TiO2 under
visible light irradiation using aromatic amines as pollutants models. In the present work, Y-N doped
or codoped titania thin films with controlled composition were prepared by the sol-gel process using
the dip-coating method. The effect of Y and N dopants on the structure and the phase stability was
studied by X-ray diffraction and Raman spectroscopy. The efficiency of these samples as photocatalysts
was investigated for the degradation of toluidine as organic compound models under a halogen lamp
used as an irradiation source.
Materials 2017, 10, 600 3 of 13

Materials 2017, 10, 600 3 of 13


2. Results and Discussion
2. Results and Discussion
2.1. SEM Analysis
2.1. SEM Analysis
Surface morphology of pure TiO2 , single doped and Y-N codoped TiO2 thin film, calcined at
400 ◦ C,Surface
is shownmorphology
in Figure of pureThe
1a–d. TiOundoped
2, single doped and Y-N codoped TiO2 thin film, calcined at 400 °C,
and N doped TiO2 was already fully characterized [14].
is shown in Figure 1a–d. The undoped
As can be seen from these FE-SEM micrographs, and N doped TiO2 was
relatively already
dense, fullyand
uniform, characterized [14]. As
nano-granular film
can be seen from these FE-SEM micrographs, relatively dense, uniform, and nano-granular film was
was obtained for all samples. Higher magnification revealed a nanostructured films composed with
obtained for all samples. Higher magnification revealed a nanostructured films composed with
relatively uniform shape nanoparticles with average grain size between 20 and 30 nm as estimated
relatively uniform shape nanoparticles with average grain size between 20 and 30 nm as estimated
from scratched film (see inset Figure 1d). The thickness of film was in the range of 300–500 nm as
from scratched film (see inset Figure 1d). The thickness of film was in the range of 300–500 nm as
estimated from the edge as shown in Figure 1d.
estimated from the edge as shown in Figure 1d.

Figure 1. FE-SEM images of: (a) undoped TiO2 (b): 2% N dopedTiO2 (c): 2% Y doped TiO2 and (d): 2%
Figure 1. FE-SEM
Y-N codoped images
TiO of: (a) undoped TiO (b): 2% N dopedTiO2 (c): 2% Y doped TiO2 and (d): 2%
2 thin films annealed at 4002°C.
Y-N codoped TiO2 thin films annealed at 400 ◦ C.
2.2. GXRD Characterization
2.2. GXRD Characterization
Grazing angle X-ray diffraction (GXRD) was carried out to investigate the effect of nitrogen and
Grazing
yttrium doping angleonX-ray diffraction
the crystal structure (GXRD)
of TiO2was carried
. Figure outthe
2 shows to investigate
GXRD patterns the ofeffect of nitrogen
undoped TiO2,
and yttrium
N-TiO doping
2, Y-TiO 2, andonY-Nthe TiO
crystal structure
2 thin of TiO2at. Figure
films calcined 400 °C.2From
shows the patterns,
these GXRD patterns of undoped
it is clear that the
TiOdiffractogram recorded
2 , N-TiO2 , Y-TiO 2 , andforY-N
the undoped
TiO2 thinTiO 2, is slightly
films at 400to◦ C.
calcinedsimilar thatFrom
of TiO 2 doped
these with 2%
patterns, it N.
is For
clear
both, only one weak reflection associated to the anatase phase was observed,
that the diffractogram recorded for the undoped TiO2 , is slightly similar to that of TiO2 doped with indicating their low
2%crystallinity.
N. For both,However,
only one introducing
weak reflection Y asassociated
dopant improves the crystallinity
to the anatase phase wasofobserved,
TiO2. Indeed, five
indicating
reflections
their are observed
low crystallinity. for Y-TiO
However, 2, 2% Y-N
introducing Y TiO 2 and 4%
as dopant Y-N TiO
improves 2 samples.
the Theyofappear
crystallinity at 2θ
TiO2 . Indeed,
values:
five 25.3°, are
reflections 37.9°, 48.03°, 53.9°,
observed 55.1°2 and
for Y-TiO , 2%62.7°,
Y-N TiO corresponding to the (101), (004), (200), (105), (211)
2 and 4% Y-N TiO2 samples. They appear at 2θ
and (204) ◦planes, ◦ respectively.
◦ ◦All of ◦them are associated
◦ to
values: 25.3 , 37.9 , 48.03 , 53.9 , 55.1 and 62.7 , corresponding to the (101), the anatase phase.(004),
No reflection related
(200), (105), (211)
andto (204)
a secondary
planes,phase due to YAll
respectively. or of
N dopants
them areisassociated
distinguished.
to theThis resultphase.
anatase is in agreement with related
No reflection those
to reported
a secondary by Khan
phase anddueCao to[16]
Y orand Zhang etisal.distinguished.
N dopants [17], showing thatThisonly theisanatase
result phase iswith
in agreement obtained
those
for TiO2 powders doped TiO2 using various Y concentrations. On the other hand, a careful analysis
Materials 2017, 10, 600 4 of 13

reported by Khan and Cao [16] and Zhang et al. [17], showing that only the anatase phase is obtained
Materials
for TiO 2017, 10, 600 4 of 13
2 powders doped TiO2 using various Y concentrations. On the other hand, a careful analysis
of the GXRD patterns reveal a slight refection broadening with increasing Y-doping concentration,
of the GXRD patterns reveal a slight refection broadening with increasing Y-doping concentration,
suggesting a systematic decrease in the grain size. Applying Scherrer formula [18] and taking into
suggesting a systematic decrease in the grain size. Applying Scherrer formula [18] and taking into
account the full width at half maximum (FWHM) for the (101) reflection, typical values of crystallite
account
Materials the
2017,full width at half maximum (FWHM) for the (101) reflection, typical values of
10, 600 crystallite
4 of 13
sizes sizes
are calculated from XRD patterns (Figure 2). The average sizes of 2% Y doped TiO
are calculated from XRD patterns (Figure 2). The average sizes of 2% Y doped TiO2,2 2% Y-N
, 2% Y-N
codoped TiO and 4% Y-N codoped TiO particles, calcinated at 400 ◦ C, are found to be 18, 16 and
codoped 2TiO2 and
of the GXRD patterns reveal
4% Y-N a slight TiO
codoped 2 2 particles,
refection broadening with increasing
calcinated at 400 °C,Y-doping
are foundconcentration,
to be 18, 16 and 14
14 nm,nm,
demonstrating that doping TiO with 3+
YY3+size.
or
suggesting a systematic
demonstrating that decrease
doping in2 the
TiO grain
2 with or N contributed
Applying
N Scherrer
contributed tolowering
to lowering
formula [18]thesize
the sizetaking
and ofofthe
theinto
crystallite,
crystallite,
account
therefore
therefore the
inhibiting fullthe
inhibiting width
theatgrowth
growthhalfof
maximum
TiO (FWHM)Furthermore,
2 particle.
of TiO 2 particle.
for the (101) reflection,
Furthermore, typical
ititisisworth
worth to values
to note
note of crystallite
that
that thepresence
the presenceofof
sizes
Y enhanced are
the calculated from
crystallization XRDof patterns
the
Y enhanced the crystallization of the TiO TiO (Figure
phase. 2). The average
However, sizes
it is of 2%
importantY doped
to beTiO 2, 2% Y-N
precise thatthese
these
2 2 phase. However, it is important be precise that
codoped TiO2 and 4% Y-N codoped TiO2 particles, calcinated at 400 °C, are found to be 18, 16 and 14
two factors: the presence
two factors: the presenceof anatase phase
of anatase phaseandandthethedecrease
decreaseofofparticles
particles nanosize
nanosize areare favorable
favorabletotothe the
nm, demonstrating that doping TiO2 with Y3+ or N contributed to lowering the size of the crystallite,
photocatalytic
photocatalytic activity.
activity.
therefore inhibiting the growth of TiO2 particle. Furthermore, it is worth to note that the presence of
Y enhanced the crystallization of the TiO2 phase. However, it is important to be precise that these
(101)
two factors: the presence of anatase
 phase and the decrease of particles nanosize are favorable to the
: Anatase
photocatalytic activity.
TiO2+4%Y-N
(004) (200)
(101)   (105) (211) (204)
  
: Anatase

TiO +2%Y-N
Intensity (a.u.)

TiO2+4%Y-N
2
(004) (200)
  (105) (211) (204)
 
TiO2+2%Y
TiO2+2%Y-N
Intensity (a.u.)

TiO2 +2%N
TiO2+2%Y

TiO2Undoped
+2%N TiO2

20 30 40 50 60 TiO
Undoped 70
2

2θ (°)
20 30 40 50 60 70
FigureFigure 2. GXRD
2. GXRD patterns
patterns of undoped
of the the undoped
TiOTiO2 and 2%, 4% Y and/or N doped TiO2 thin films annealed
2 and
2θ (°)2%, 4% Y and/or N doped TiO2 thin films annealed
at◦ 400 °C.
at 400 C.
Figure 2. GXRD patterns of the undoped TiO2 and 2%, 4% Y and/or N doped TiO2 thin films annealed
at 400 °C.
Figure 3 shows the GXRD patterns of 2% Y-N codoped TiO2 thin film calcined at different
Figure 3 shows
temperatures. Onlythe
theGXRD
anatasepatterns of 2% Y-N
phase is observed codoped
for the samplesTiO 2 thinatfilm
calcined calcined
500 °C, 550 °C andat different
650 °C.
temperatures.Figure 3 shows the GXRD patterns of 2% Y-Nfor codoped TiO2 thincalcined
film calcined at ◦different ◦ C and
This result suggests that the typical anatase-rutile phase transformation [19] is delayed550
Only the anatase phase is observed the samples at 500 C, to higher
temperatures. Only the anatase phase is observed for the samples calcined at 500 °C, 550 °C and 650 °C.
650 ◦ C. This resultThe
temperature. suggests
anatase that theistypical
phase probably anatase-rutile
stabilized by phase transformation
the surrounding [19] ions
is delayed
throughto
This result suggests that the typical anatase-rutile phase transformation [19] is rare earth
delayed to higher
higher temperature.
the formation of The anatase
Ti-O-rare earth phase
bonds is probably
[20]. stabilized by the surrounding
temperature. The anatase phase is probably stabilized by the surrounding rare earth ions through
rare earth ions
throughthethe formation of Ti-O-rare earth bonds
formation of Ti-O-rare earth bonds [20]. [20].
: Anatase
(101) : Anatase

(101)

(a.u.)
Intensity(a.u.)

(004) (200)(200)
(004)  (105)(211)
Intensity

 650°C
 (204) (204)
(105)(211) 650°C
 (110)(220)(215)
  (110)(220)
(215)
 
  

550°C550°C

500°C500°C

10
10 20
20 3030 40 40 50 50 60 60 70 70 80 80
2
2
Figure 3. GXRD patterns of the 2% Y-N codoped TiO2 thin films annealed at different temperature.
Figure
Figure 3. GXRD
3. GXRD patterns
patterns of the
of the 2% 2%
Y-NY-N codoped
codoped TiO
TiO 2 thin
2 thin filmsannealed
films annealed at
at different
different temperature.
temperature.
Materials 2017, 10, 600 5 of 13

Materials 2017, 10, 600 5 of 13

2.3. Raman Spectroscopy


2.3. Raman Spectroscopy
The structural characteristics and phase composition of the samples were further investigated by
The structural characteristics and phase composition of the samples were further investigated
micro-Raman spectroscopy. Figure 4a,b shows the Raman spectra of undoped TiO2 and Y-N-doped
by micro-Raman spectroscopy. Figure 4a,b shows the Raman spectra of undoped TiO2 and
or codoped TiO2 , films and powders, respectively. Only the anatase phase, which is characterized
Y-N-doped or codoped TiO2, films and powders, respectively. Only the anatase phase, which is
by six Raman active modes (A1g + 2B1g + 3Eg) [21], is observed for all samples. The dominant Eg
characterized by six−1Raman active modes (A1g + 2B1g + 3Eg) [21], is observed for all samples. The
observed at 144 cm is associated with the Ti-O bending vibration. It is also attributed to the two
dominant Eg observed at 144 − cm −1 is associated with the Ti-O bending vibration. It is also attributed
1
weak bands at 197 and 638 cm . However, B1g and the (A1g + B1g ) bands are shown around 395 cm−1
to the two weak
− 1
bands at 197 and 638 cm−1. However, B1g and the (A1g + B1g) bands are shown around
and 517−1cm , respectively. Furthermore, data in Figure 4a indicated a Raman band broadening for
395 cm and 517 cm , respectively. Furthermore, data in Figure 4a indicated a Raman band
−1
Y-N codoped TiO2 . These results can be ascribed to phonon confinement and surface strain effects,
broadening for Y-N codoped TiO2. These results can be ascribed to phonon confinement and surface
usually observed in nanostructured materials [22]. These outcomes are in good agreement with values
strain effects, usually observed in nanostructured materials [22]. These outcomes are in good
of crystallite size determined using Scherrer’s formula and XRD studies.
agreement with values of crystallite size determined using Scherrer’s formula and XRD studies.

a) b)
145
TiO2 + 4% Y-N
TiO2 + 2% Y-N

397 517 638


TiO2 + 2% Y-N
Ram an Intensity (a.u.)

Raman Intensity (a.u.)

TiO2 + 2% Y

TiO2 +2%Y

TiO2 + 2% N

TiO2 + 2% N

TiO2 undoped 144


TiO2 undoped

395 517 638

200 300 400 500 600 700 200 300 400 500 600 700
-1
Wavenumber cm Wavenumber cm-1

Figure 4.
Figure 4. Raman
Raman spectra
spectra of
of undoped
undoped and
and 2%
2% Y-N doped and
Y-N doped and codoped
codoped TiO
TiO22 thin
thin films
films (a)
(a) and
and powder
powder
(b) annealed
(b) annealed at
at 400 ◦
400 °C.
C.

Additionally, micro-Raman analyses for the nanopowders in Figure 4b show that the
Additionally, micro-Raman analyses for the nanopowders in Figure 4b show that the introduction
introduction of N decreases the anatase crystallinity by means of a broadening effect mainly
of N decreases the anatase crystallinity by means of a broadening effect mainly observed on the 395,
observed on the 395, 517 and 638 cm−1 peaks of the Raman signal. The anatase structure is greatly
517 and 638 cm−1 peaks of the Raman signal. The anatase structure is greatly affected by the codoping
affected by the codoping with 2% Y-N and 4% Y-N, and it does not change with the introduction of
with 2% Y-N and 4% Y-N, and it does not change with the introduction of Yttrium as a single dopant.
Yttrium as a single dopant.
2.4. Ground State Diffuse Reflectance Spectra
2.4. Ground State Diffuse Reflectance Spectra
In Figure 5, one can see the ground state diffuse reflectance spectra of samples with different
In Figure
dopants. 5, one cannanoparticles
TiO2 undoped see the ground state were
spectra diffuse
alsoreflectance
included spectra of samples
for comparison. with different
A significant tail
dopants. TiO 2 undoped nanoparticles spectra were also included for comparison. A significant tail
was observed in all diffuse reflectance spectra, reaching the visible range of wavelengths. The cut-off
was observedwere
wavelengths in allobtained
diffuse reflectance spectra, of
via the intersection reaching the visibleextrapolations
the straight-line range of wavelengths.
below andThe cut-off
above the
wavelengths were obtained via the intersection of the straight-line extrapolations below
small photon energy knee, in the Tauc plots from the curves presented in Figure 5a, which provided and above
the gap
the small photon
energy energy
values knee, in
presented inTable
the Tauc plots5bfrom
1. Figure the the
presents curves presented
calculation in band
of the Figure 5a,energies
gap which
provided
in the casethe gappure
of the energyTiOvalues presented in Table 1. Figure 5b presents the calculation of the band
2 and TiO2 + N samples.
gap energies in the case of the pure TiO2 and TiO2 + N samples.
Materials 2017, 10, 600 6 of 13

Materials 2017, 10, 600 6 of 13

Figure 5. (a) ground-state diffuse reflectance absorption spectra of undoped, doped and codoped
Figure 5. (a) ground-state diffuse reflectance absorption spectra of undoped, doped and codoped TiO2
TiO2 with N and Y, annealed at ◦400 °C; (b) Tauc plots of pure and N-doped TiO2 samples, evidencing
with N and Y, annealed at 400 C; (b) Tauc plots of pure and N-doped TiO2 samples, evidencing the
the calculation of the band gap energies for these two samples.
calculation of the band gap energies for these two samples.

Table 1. Energy gaps and λabs for TiO2 undoped, doped and codoped with N and Y.
Table 1. Energy gaps and λabs for TiO2 undoped, doped and codoped with N and Y.
Sample λabs (nm) Eg (ev)
Sample λabs (nm) Eg (ev)
Undoped TiO2 396 3.13
Undoped
N dopedTiO2 396
392 3.133.17
2% YNdoped
doped 392
400 3.17
3.10
2% Y doped 400 3.10
2%2%
Y-N codoped
Y-N codoped 385
385 3.223.22
4%4%
Y-N codoped
Y-N codoped 396
396 3.133.13

Ground
Groundstatestatediffuse
diffusereflectance
reflectanceabsorption
absorptionstudies
studiesfor the
for nanopowders
the nanopowders showshowthat, while
that, thethe
while N
dopant
N dopant causes deviations
causes of of
deviations thethe
band
bandgap
gaptotohigher
higherenergies,
energies,doping
dopingTiOTiO2 2with
withYY suggests
suggests the
the
opposite effect. Indeed, the value of energy gap ∆E increased from 3.13 eV (undoped)
opposite effect. Indeed, the value of energy gap ∆E increased from 3.13 eV (undoped) to about to about 3.16
3.16 eV
eV in TiO 2 doped with N and to 3.22 eV with 2% Y-N, and it reached 3.13 eV for 4% Y-N (Table 1).
in TiO2 doped with N and to 3.22 eV with 2% Y-N, and it reached 3.13 eV for 4% Y-N (Table 1). On the
On thehand,
other otherwith
hand,
2%with 2% Y ∆E
Y dopant, dopant, ∆E decreased
decreased to 3.10 eV,toshowing
3.10 eV,that
showing
Yttriumthat Yttrium
shifts shifts the
the wavelength
wavelength
absorption range to higher values and this also explains the difference between 2% Y-N and2%
absorption range to higher values and this also explains the difference between 4% Y-N
Y-N.
and 4% Y-N. The same trend was found in the work of Zhao and co-workers [23], which showed that
The same trend was found in the work of Zhao and co-workers [23], which showed that Y-doping
Y-doping also induced a red shift of the absorption edge. Concretely, the absorption threshold
also induced a red shift of the absorption edge. Concretely, the absorption threshold shifted from
shifted from 395 nm (undoped) to 405 nm (3% Y-TiO2). The most probable reason was that Y doping
395 nm (undoped) to 405 nm (3% Y-TiO2 ). The most probable reason was that Y doping induced a
induced a part of the Y 4d states to extend into the TiO2 conduction band, resulting in a narrowing of
part of the Y 4d states to extend into the TiO2 conduction band, resulting in a narrowing of the band
the band gap of anatase. Therefore, these results suggest that the introduction of these two elements
gap of anatase. Therefore, these results suggest that the introduction of these two elements (N, Y) can
(N, Y) can modify the optical behavior of the material in different ways, presumably due to opposite
modify the optical behavior of the material in different ways, presumably due to opposite functions
functions related with the modification of the energy band gap and/or the creation of surface traps
related with the modification of the energy band gap and/or the creation of surface traps for the
for the semiconductor’s charge carriers.
semiconductor’s charge carriers.

2.5.
2.5. Laser
LaserInduced
InducedLuminescence
LuminescenceStudies
Studies
Figure
Figure 6a–e
6a–e shows
shows the the laser
laser induced
induced luminescence spectra (λ
luminescence spectra (λexc
exc =
=337337nm,
nm, ~1.3
~1.3mJ mJper
perexcitation
excitation
pulse),
pulse), at 77 K of the powdered samples. The initial curve was obtained immediately after laser
at 77 K of the powdered samples. The initial curve was obtained immediately after laser
pulse,
pulse, and
and the
the other
other curves
curves are are separated
separated byby steps
steps ofof 20
20 ns.
ns. The
The used
used time
time gate
gate width
width was
was oneone
microsecond.
microsecond.The Thelaser
laserinduced
inducedemission
emissionspectra
spectraofofundoped
undopedTiO TiO2 is is
2
presented
presented in Figure
in 6a,
Figure with
6a, witha
maximum at 520 nm. It exhibits a decay with a lifetime of aproximately
a maximum at 520 nm. It exhibits a decay with a lifetime of aproximately 59 ns. This luminescent59 ns. This luminescent
emission
emission isis attributed
attributedtoto thethe typical
typical andand well-known
well-known greengreen luminescent
luminescent band ofband of anatase.
anatase. This
This emission
emission
occurs due occurs
to thedue to the recombination
recombination of thecharge
of the excited excited chargethe
carriers, carriers, theand
electron electron and hole
hole pair, frompair,
trap
from trap states located either at the surface or within the bulk phase of the crystalline
states located either at the surface or within the bulk phase of the crystalline material [24]. Figure 6b material [24].
Figure 6b presents the luminescence spectra of TiO2, doped with N. The fluorescence lifetime of the
sample changes, increasing to approximately 65 ns. It was previously mentioned that, with the
introduction of N, the energy band gap increased, compared with the undoped TiO2. It seems that,
Materials 2017, 10, 600 7 of 13

presents the luminescence spectra of TiO2 , doped with N. The fluorescence lifetime of the sample
Materials
changes, 2017, 10, 600
increasingto approximately 65 ns. It was previously mentioned that, with the introduction 7 of 13

of N, the energy band gap increased, compared with the undoped TiO2 . It seems that, as the energy
as the energy bang gap is higher, so is the fluorescence lifetime. These results suggest that, together
bang gap is higher, so is the fluorescence lifetime. These results suggest that, together with the
with the broadening effect of N observed in Raman spectra, the element could serve as surface
broadening effect of N observed in Raman spectra, the element could serve as surface defects, which
defects, which can decrease the material’s crystallinity but simultaneously enhance the electron-hole
can decrease the material’s crystallinity but simultaneously enhance the electron-hole separation,
separation, functioning in this case as trap sites for the separated charged states of the
functioning in this case as trap sites for the separated charged states of the semiconductor. The results
semiconductor. The results presented in Figure 6c prove that, as opposed to N-doped sample,
presented in Figure 6c prove that, as opposed to N-doped sample, Yttrium should enable the creation of
Yttrium should enable the creation of intraband gap states that do not alter the crystallinity but
intraband gap states that do not alter the crystallinity but actually change the electronic recombination
actually change the electronic recombination mechanism of the semiconductor. In this case, a
mechanism of the semiconductor. In this case, a significant quenching of the fluorescence signal is
significant quenching of the fluorescence signal is observed, together with a small decrease of the
observed, together with a small decrease of the fluorescence lifetime, which is now approximately
fluorescence lifetime, which is now approximately 58 ns. This points out to a slight increase of the
58 ns. This points out to a slight increase of the recombination rate of the electron-hole pair. Figure 6d,e
recombination rate of the electron-hole pair. Figure 6d,e confirm the trend, presenting a significant
confirm the trend, presenting a significant change in the fluorescence lifetime from the codoped 2% Y-N
change in the fluorescence lifetime from the codoped 2% Y-N sample to the codoped 4% Y-N
sample to the codoped 4% Y-N sample, which infers a decrease of approximately 59 ns to around 44 ns,
sample, which infers a decrease of approximately 59 ns to around 44 ns, respectively. Therefore, the
respectively. Therefore, the photoluminescence of these samples is, as expected, directly correlated
photoluminescence of these samples is, as expected, directly correlated with the recombination
with the recombination process of the electron-hole pair [25]. Indeed, by increasing the energy gap and
process of the electron-hole pair [25]. Indeed, by increasing the energy gap and while delaying the
while delaying the recombination carrier, the radiative recombination lifetime should be higher, leading
recombination carrier, the radiative recombination lifetime should be higher, leading also to higher
also to higher luminescence lifetimes. The doping of Yttrium into titania quenches the fluorescence
luminescence lifetimes. The doping of Yttrium into titania quenches the fluorescence signal because
signal because the excited electron can transfer from the valence band to the new levels that exist
the excited electron can transfer from the valence band to the new levels that exist beneath the
beneath the conduction band, which can decrease the photoluminescence intensity. In this way, TiO2
conduction band, which can decrease the photoluminescence intensity. In this way, TiO2 doped with
doped with N and co-doped with 2% Y-N could have a higher photo-catalytic activity by delaying the
N and co-doped with 2% Y-N could have a higher photo-catalytic activity by delaying the
electron-hole recombination, despite their lower absorption in the visible range, when compared with
electron-hole recombination, despite their lower absorption in the visible range, when compared
the sample doped with 2% Y. However, it is important to be precise that the value of the lifetime of
with the sample doped with 2% Y. However, it is important to be precise that the value of the
photoexcited charge in 2% Y-N is much smaller than the observed one in the 2% Li-N codoped sample,
lifetime of photoexcited charge in 2% Y-N is much smaller than the observed one in the 2% Li-N
which is 72 ns [15].
codoped sample, which is 72 ns [15].

Figure 6. Cont.
Materials 2017, 10, 600 8 of 13
Materials 2017, 10, 600 8 of 13

Figure 6.
Figure 6. Time
Time resolved
resolved laser
laser induced
induced luminescence
luminescence spectra,
spectra, at
at 77
77 K,
K, of:
of: anatase
anatase undoped
undoped (a);
(a); doped
doped
with N
with N (b);
(b); doped
doped with
with 2%
2% YY (c);
(c); codoped
codoped with
with 2%
2% (d)
(d) and
and 4%
4% ofof Y-N
Y-N (e).
(e). The
The excitation
excitation wavelength
wavelength
was 337 nm. The initial curve was obtained immediately
immediately after
after the laser pulse, and, for all of other
curves, the time step was 20 ns annealed at 400 ◦°C.
C.

2.6. Photocatalytic
2.6. Photocatalytic Degradation
Degradation
The photocatalytic
The photocatalytic activity
activity of allofprepared
all prepared
samples samples
was evaluated was by evaluated
measuringbythemeasuring
decomposition the
decomposition
rate of toluidinerate underof toluidine underfrom
photoexcitation photoexcitation
a halogen light from a halogen
source. Figure light source.
7 shows theFigure 7 shows
degradation of
the degradation of the organic solute model without any catalyst
the organic solute model without any catalyst and in the presence of undoped TiO2 , N or Y doped and in the presence of undoped
TiO2Y-N
and , N or Y doped
codoped TiOand Y-N codoped TiO2 at a level of 2%. As shown in Figure 7, the residual
2 at a level of 2%. As shown in Figure 7, the residual concentration of toluidine
concentration
after of toluidine
8 h of irradiation after
in the 8 h of irradiation
presence of undoped in the
TiOpresence of undoped TiO2 thin ◦ films calcined
2 thin films calcined at 400 C is about 64%,
at 400 °C is about 64%, attesting the low photocatalytic activity of the pure
attesting the low photocatalytic activity of the pure TiO2 . Doping TiO2 with N or Y notably affects the TiO 2. Doping TiO2 with N

or Y notably affects
photocatalytic theResults
activity. photocatalytic
reportedactivity.
in Figure Results reported
7 revealed in Figure
different trends7 depending
revealed different trends
on the doping
depending on the doping element. The largest catalytic effect is observed
element. The largest catalytic effect is observed in the presence of 2% Y-N, for which the residual in the presence of 2% Y-N,
for which the residual concentration of toluidine attained 12% after
concentration of toluidine attained 12% after 8 h of light irradiation. One should note that, although 8 h of light irradiation. Oneall
should note that, although all of the samples crystallize in the anatase
of the samples crystallize in the anatase form of TiO2 , their photocatalytic activities differ considerably. form of TiO 2, their

photocatalytic
This rules out activities differ considerably.
the assignment of the difference Thisinrules
the out the assignment
degradation performanceof the difference in the
to the structural
degradation performance to the structural properties. However, this
properties. However, this result is probably due to the synergetic role of the dopants dispersed on result is probably due to the
the
synergetic role of the dopants dispersed on the TiO 2 surface (as demonstrated by the decrease of
TiO2 surface (as demonstrated by the decrease of crystallite sizes of the samples), yielding low average
crystallite between
distances sizes of the thesamples), yielding
dopant centers low as
acting average
traps. distances
However,betweencompared thetodopant
2% Y-N centers
codoped acting
TiOas2,
traps. However,
increasing compared
the Y level leads toto huge
2% Y-N codoped
decreases TiOphotocatalytic
in the 2, increasing the Y level leads to huge decreases
performance. Indeed, the residual
in the photocatalytic
concentration performance.
of toluidine reachedIndeed,
39% and the44%
residual
in theconcentration
case of 3% Y-N of toluidine
and 4% Y-N, reached 39% and
respectively.
We argue that lower Y doping is more beneficial to the degradation process compared toishigher
44% in the case of 3% Y-N and 4% Y-N, respectively. We argue that lower Y doping more
beneficial
Y doping. to Inthe degradation
addition, with theprocess
increasecompared
of the Y todoping
higher Y doping.
level, In addition,
the distance with the
between increase
trapping of
sites
the Y doping level, the distance between trapping sites decreases. The
decreases. The Y site at the surface of TiO2 nanoparticles can act as the charge carrier recombination2 Y site at the surface of TiO
nanoparticles
center [17], whichcan act as the
results in charge
a decrease carrier recombination
of the center [17], which
number of photogenerated results
charge, in a contributes
which decrease of
thethenumber
in degradation of process.
photogenerated charge, which
Photoluminescence analysis contributes
corroborates in wellthewith degradation process.
these experimental
observations, since the lifetime significantly decreases with increasing the doping level of Y from the
Photoluminescence analysis corroborates well with these experimental observations, since 2%
lifetime
Y-N to 4% significantly
Y-N. decreases with increasing the doping level of Y from 2% Y-N to 4% Y-N.
In summary,
In summary, careful
careful analysis
analysis of of the
the experimental
experimental results
results reveals
reveals energy
energy gap gap values,
values, which
which onlyonly
allows the absorption of irradiation in the UV region. However,
allows the absorption of irradiation in the UV region. However, interesting performance was evidenced interesting performance was
evidenced
for for the photodegradation
the photodegradation of toluidineofpollutant
toluidineunder
pollutanthalogenunderlamp, halogen lamp,inespecially
especially the presence in the
of
presence of 2% Y-N codoped TiO 2. Both sunlight and W-Hal lamp have a small tail in the UV region,
2% Y-N codoped TiO2 . Both sunlight and W-Hal lamp have a small tail in the UV region, and this
and explain
may this may theexplain
observed the observed photocatalytic
photocatalytic activity.
activity. To explain thisTo trend,explain
numerousthis trend,
hypothesesnumerous
have
hypotheses have been proposed, some relevant ones that claimed that
been proposed, some relevant ones that claimed that the N dopant atoms resulted in the creation of the N dopant atoms resulted
in the creation
surface of surface
oxygen vacancies oxygen
(OVs) [26]. vacancies
On the other (OVs)
hand, [26]. On the
Y dopant seemsother to hand,
promote Y the
dopant seems to
recombination
promote the recombination process, thereby decreasing the photocatalytic
process, thereby decreasing the photocatalytic efficiency. This reactivity was also attributed by other efficiency. This reactivity
was alsotoattributed
authors by other authors to
a charge-transfer-complex a charge-transfer-complex
mechanism, in which neither mechanism,the photocatalyst in which neither
nor the the
organic
photocatalyst nor the organic compounds
compounds absorb visible light by themselves [27,28]. absorb visible light by themselves [27,28].
Materials 2017, 10, 600 9 of 13
Materials 2017, 10, 600 9 of 13
Materials 2017, 10, 600 9 of 13

Figure 7. Evolution of the residual concentration of toluidine in the presence of undoped, Y-N doped
Figure 7.
7. Evolution
Evolution of
ofthe
the residual
residual concentration
concentration of toluidine in the presence of undoped, Y-N doped
Figure
and codoped TiO2 thin films annealed at 400 °C. of toluidine in the presence of undoped, Y-N doped
and codoped TiO 2 thin films annealed at 400 °C.
and codoped TiO thin films annealed at 400 C. ◦
2
2.7. Reuse of the Photocatalyst
2.7. Reuse of the Photocatalyst
2.7. Reuse of the Photocatalyst
In view of practical application, the photocatalyst should be chemically and optically stable
In view of practical application, the photocatalyst should be chemically and optically stable
afterIn view repeated
several of practical application,
cycles. the photocatalyst
To investigate the reusability should
of 2% beN,chemically
2% Y andand optically
2% Y-N doped stable
and
after several repeated cycles. To investigate the reusability of 2% N, 2% Y and 2% Y-N doped and
after several repeated cycles. To investigate the reusability of 2% N, 2%
codoped TiO2 in the photocatalytic degradation of toluidine experiment was repeated three times. Y and 2% Y-N doped and
codoped TiO2 in the photocatalytic degradation of toluidine experiment was repeated three times.
codoped TiO2 in the
The concentration photocatalytic
of the degradation
organic pollutant of toluidine
was measured afterexperiment
8 h exposedwas repeated
to lamp three times.
irradiation. After
The concentration of the organic pollutant was measured after 8 h exposed to lamp irradiation. After
The concentration of the organic pollutant was measured after 8 h exposed
each decomposition reaction, the beads were rinsed with water, dried and used again in the same to lamp irradiation. After
each decomposition reaction, the beads were rinsed with water, dried and used again in the same
each decomposition
conditions. As shown reaction,
in Figurethe beads
8, the were
betterrinsed
results with
werewater, dried and
observed in theused again in
presence ofthe
2%same
Y-N
conditions. As shown in Figure 8, the better results were observed in the presence of 2% Y-N
conditions.
codoped TiO As2,shown in Figure
for which, after8,three
the better results
cycles, weredecrease
a slight observedininphotodegradation
the presence of 2%efficiency
Y-N codoped was
codoped TiO2, for which, after three cycles, a slight decrease in photodegradation efficiency was
TiO ,
observed.
2 for which, after three cycles, a slight decrease in photodegradation efficiency
In fact, in this case, the residual concentration of toluidine increased from 12% to 16%. was observed.
observed. In fact, in this case, the residual concentration of toluidine increased from 12% to 16%.
In fact,
This may in this case, the by
be explained residual concentration
the formation of someof intermediate
toluidine increasedspeciesfrom 12% to 16%.
that remained This may
adsorbed be
at the
This may be explained by the formation of some intermediate species that remained adsorbed at the
explained
surface of by the formation
catalyst particles.of some intermediate
Judging species2%
from these results, that
Y-Nremained
codoped adsorbed
TiO2 canatbetheregarded
surface ofas
surface of catalyst particles. Judging from these results, 2% Y-N codoped TiO2 can be regarded as
catalyst particles. Judging from these results, 2%
stable and owing a considerable reproducibility of photodegradationY-N codoped TiO can be regarded as stable
2 of toluidine, which makes its and
stable and owing a considerable reproducibility of photodegradation of toluidine, which makes its
owing
use in thea considerable reproducibility
practical process possible. Of ofcourse,
photodegradation
further workofistoluidine,
required towhich
bettermakes its usefactors
understand in the
use in the practical process possible. Of course, further work is required to better understand factors
practical
that affect process possible. Of course,rate
the photodegradation further workY-N
of 2% is required
codoped to better
TiO2 andunderstand factors
to further that affect
improve the
that affect the photodegradation rate of 2% Y-N codoped TiO2 and to further improve the
the photodegradation
photocatalyst stability. rate of 2% Y-N codoped TiO 2 and to further improve the photocatalyst stability.
photocatalyst stability.

Figure 8.
Figure 8. Evolution
Evolution ofof the
the residual
residual concentration
concentration of
of toluidine
toluidine after
after three
three degradation
degradation cycles
cycles in
in the
the
Figure 8. Evolution of the residual concentration of toluidine after three degradation cycles in the
presence
presence ofof 2%
of 2% N,
2% N, 2%
N, 2%
2% YY and
Y and 2%
and 2% Y-N
2% Y-N codoped
Y-N codoped TiO 2 photocatalyst annealed at 400◦°C.
presence codoped TiO photocatalyst annealed
TiO22 photocatalyst annealed atat 400 C.
400 °C.
Materials 2017, 10, 600 10 of 13

3. Materials and Methods

3.1. Preparation of Films and Powders


Y-N co-doped TiO2 samples were deposited using the sol–gel chemical process and dip-coating
techniques with Ti(OBu)4 as precursor and acetylacetone (C5 H8 O2 ) as solvent. To achieve the
doping, YCl3 ·6H2 O and CH3 (CH2 )4 CH2 NH2 were used as sources of yttrium and nitrogen dopants,
respectively. As a first step, an appropriate amount of YCl3 ·6H2 O was dissolved in acetylacetone.
Three different solutions corresponding to three atomic ratios Y/Ti = 0.02, 0.03 and 0.04 were elaborated.
The hexylamine was then slowly added to the limpid solution under vigorous stirring (Atomic ratio
N/Ti = 0.02). Afterwards, Ti(OBu)4 were introduced. Then, the activated substrate of glass was
immersed in the prepared sol for 2 min. The obtained transparent thin films were then calcined at
different temperatures: 400, 500, 550 and 650 ◦ C for 2 h by heating at a rate of 5 ◦ C/min followed
by a free cooling. The remaining solutions were kept under stirring for several hours to obtain the
powdered samples. For comparison purposes, undoped and Y or N monodoped TiO2 were prepared
using the same synthetic protocol.

3.2. Physical and Chemical Characterization of the Catalysts


The surface morphology of the Y-, N-doped or codoped TiO2 thin films was studied by
field-emission scanning electron microscopy (FE-SEM) using ZEISS SUPRA40 (ZEISS, Germany) fully
controlled from a computer workstation. The electron source, a hot cathode producing electrons
by Schottky effect, is a tungsten filament coated with a ZrO layer. Images are created by the
software SMARTSEM.
The crystal structure of the obtained films was characterized by the grazing angle X-ray diffraction
GXRD technique using an X-ray diffractometer Panalytical Xpert with CuKα radiation. The scanning
range (2θ) was from 20◦ to 75◦ with a scanning rate of 0.05◦ /min and an incidence angle of 0.5◦ .
The Raman spectra of the powders were obtained in a back-scattering micro-configuration, with
532 nm excitation (Cobolt Samba CW DPSSL, 300 mW, Stockholm, Sweden) and by the use of a
SuperHead 532 from Horiba JobinYvon (JY) (Villeneuve-d’Ascq, France) with a 50× Edmund long
working distance objective. The Raman probe was coupled to a Shamrock 163 with a 100 µm entrance
slit) and a Newton DU 971P-BV camera from Andor (Belfast, UK) was used as a detector for the
Raman signals, working at −60 ◦ C. Data acquisition was performed with the Andor software and data
processing, namely, the baseline corrections, when needed, were made with the LabSpec software from
JY (version 5.25.15, Villeneuve-d’Ascq, France). The spectral resolution of this Raman spectrometer
was ~2 cm−1 .
For the thin films, Raman spectra were recorded on a JYT64000 Raman spectrometer. A signal
was obtained on excitation of the samples by an ArKr laser (514.5 nm). Measurement was done with
the beam path set at (50×) and an exposure time of 2 × 300 s.
Ground-state absorption studies were performed using a homemade diffuse reflectance laser
flash photolysis setup, with a 150 W tungsten-halogen lamp as monitoring lamp, triggering the system
in the normal way but without the laser fire, and, in this way, recording the lamp profile for all
samples under study and also for two standards, barium sulfate and magnesium oxide powders.
A fixed monochromator coupled to an intensified charge coupled device (ICCD) with time gate
capabilities was used for detecting the reflectance signals. The reflectance, R, from each sample was
obtained in the UV–Vis-NIR (UV–Ultraviolet; Vis–visible; NIR-near infrared) spectral regions and the
remission function, F(R), was calculated using the Kubelka–Munk equation for optically thick samples.
The remission function is F(R) = (1 − R)2 /2R.
The set-up for time resolved luminescence LIL (laser induced fluorescence, LIF, and laser induced
phosphorescence, LIP) is presented in Refs. [29,30]. Time-resolved emission spectra were performed in
the nanosecond to second time range with an N2 laser (Photon technology international-PTI, model
Materials 2017, 10, 600 11 of 13

2000, ca. 600 ps, full width half maximum (FWHM), about 1 mJ per pulse, 337 nm of excitation, PTI,
ON, Canada).

3.3. Photocatalytic Activity Measurements


The photocatalytic performance of the Y-N codoped TiO2 was quantified by measuring the rate of
degradation of organic solute model under halogen light irradiation. In a typical measurement, 25 mL
of 5.10−4 mol L−1 solution of the organic solute was prepared. Titanium dioxide films deposed on beads
of glass were introduced as photocatalysts. Prior to illumination, the suspension was stirred in the dark
for 1 h to establish an adsorption/desorption equilibrium between the photocatalyst and pollutant
molecules. Then, the photocatalytic degradation of pollutant model was initiated. The degradation
test was carried out using a 47 W halogen lamp power (230 V), as a source of irradiation. The use of
a lamp assures a constant light output in opposition to the fluctuation in solar intensity with season
and time of day. To evaluate the organic solute concentration during the photodegradation reaction
under constant irradiation, 1 mL of the suspension was taken out at a given time interval and the UV
absorbance at their corresponding λmax (at 280 nm) was measured. The concentration was evaluated
from the calibration curve previously established with a known concentration of aliquot.

4. Conclusions
In this work, Y-, N-doped or codoped TiO2 powders and thin films deposited on glass beads were
successfully prepared using the sol-gel method, Ti (OBu)4 as Ti precursor.
Structurally, the samples were found to be anatase with uniform shape nanoparticles (20–30 nm)
as evidenced by FE-SEM observation. Ground state diffuse reflectance absorption studies showed
that doping with N or codoping with N and Y led to an increase of the band gap. Laser induced
luminescence analysis revealed a decrease in the recombination rate of the photogenerated holes
and electrons.
The photocatalytic activity of the prepared catalysts, under halogen light irradiation, was evaluated
using toluidine as pollutant model. Results showed a great enhancement in the photocatalytic efficiency
following the corporation of Y-N simultaneously, which is probably due to an electron-transfer
mechanism, in which neither the photocatalyst nor the organic compounds absorb visible light
by themselves. The photocatalyst was reusable for several degradation cycles with only a small
reduction in the degradation efficiency. This advantage is important in the continuous photocatalytic
degradation process.

Acknowledgments: Thanks are due to FCT (Portugal’s Foundation for Science and Technology), for the funding
project UID/NAN/50024/2013 and for the PhD fellowships (SFRH/BD/95358/2013) of David Conceição.
Author Contributions: S.B., S.B. and L.F.V.F. conceived and designed the experiments; Z.H. and D.C.
performed the experiments; D.C., S.B., S.B. and L.F.V.F. analyzed the data; S.B. and L.F.V.F. contributed
reagents/materials/analysis tools; and S.B. and L.F.V.F. wrote the paper.
Conflicts of Interest: We declare that we do not have any commercial or associative interests that represent
a conflict of interest in connection with the work submitted.

References
1. Sreekanta, S.; Ahazan, R.; Lockman, Z. Photoactivity of anatase-rutile TiO2 nanotubes formed by anodization
method. Thin Solid Films 2009, 518, 16–22. [CrossRef]
2. Hamdi, A.; Ferraria, A.M.; Rego, A.M.B.; Ferreira, D.P.; Conceição, D.S.; Ferreira, L.F.V.; Bouattour, S.
Bi–Y doped and codoped TiO2 nanoparticles: Characterization and photocatalytic activity undervisible light
irradiation. J. Mol. Catal. A Chem. 2013, 380, 34–42. [CrossRef]
3. Nosaka, Y.; Matsushita, M.; Nishino, J.; Nosaka, A.Y. Nitrogen-doped titanium dioxide photocatalysts for
visible response prepared by using organic compounds. Sci. Technol. Adv. Mater. 2005, 6, 143–148. [CrossRef]
4. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-Light Photocatalysis in Nitrogen-Doped
Titanium Oxides. Science 2001, 293, 269–271. [CrossRef] [PubMed]
Materials 2017, 10, 600 12 of 13

5. Morikawa, T.; Ohwaki, T.; Suzuki, K.; Moribe, S.; Kubota, S.T. Visible-light-induced photocatalytic oxidation
of carboxylic acids and aldehydes over N-doped TiO2 loaded with Fe, Cu or Pt. Appl. Catal. B Environ. 2008,
83, 56–62. [CrossRef]
6. Chen, X.; Luo, W. Optical spectroscopy of rare earth ion-doped TiO2 nanophosphors. J. Nanosci. Nanotechnol.
2010, 10, 1482–1494. [CrossRef] [PubMed]
7. Hamden, Z.; Ferreira, D.P.; Ferreira, L.V.; Bouattour, S. Li–Y doped and codopedTiO2 thin films: Enhancement
of photocatalytic activity under visible light irradiation. Ceram. Int. 2014, 40, 3227–3235. [CrossRef]
8. Narayan, H.; Alemu, H.; Setofolo, L.; Macheli, L. Visible Light Photocatalysis with Rare Earth Ion-Doped
TiO2 Nanocomposites. ISRN Phys. Chem. 2012. [CrossRef]
9. Wang, Y.; Lu, K.; Feng, C. Photocatalytic degradation of methyl orange by polyoxometalates supported on
yttrium-doped TiO2 . J. Rare Earth 2011, 29, 866–877. [CrossRef]
10. Bouattour, S.; Rego, A.M.B.; Ferreira, L.F.V. Photocatalytic activity of Li+ –Rb+ –Y3+ doped or codoped TiO2
under sunlight irradiation. Mater. Res. Bull. 2010, 45, 818–825. [CrossRef]
11. Kallel, W.; Bouattour, S.; Kolsi, A.W. Structural and conductivity study of Y and Rb co-doped TiO2 synthesized
by the sol-gel method. J. Non-Cryst. Solids 2006, 352, 3970–3978. [CrossRef]
12. Wu, Y.; Gong, Y.; Liu, J.; Zhang, Z.; Xu, Y.; Ren, H.; Li, C.; Niu, L. B and Y co-doped TiO2 photocatalyst with
enhanced photodegradation efficiency. J. Alloys Compd. 2017, 695, 1462–1469. [CrossRef]
13. Kuvarega, A.T.; Krause, R.W.M.; Mamba, B.B. Nitrogen/Palladium-Codoped TiO2 for Efficient Visible Light
Photocatalytic Dye Degradation. J. Phys. Chem. C 2011, 115, 22110–22120. [CrossRef]
14. Irie, H.; Watanabe, Y.; Hashimoto, K. Nitrogen-Concentration Dependence on Photocatalytic Activity of
TiO2 -xNxPowders. J. Phys. Chem. B 2003, 107, 5483–5486. [CrossRef]
15. Hamden, Z.; Boufi, S.; Conceicão, D.S.; Ferraria, A.M.; Rego, A.M.B.; Ferreira, D.P.; Ferreira, L.F.V.;
Bouattour, S. Li–N doped and codoped TiO2 thin films deposited by dip-coating: Characterization and
photocatalytic activity under halogen lamp. Appl. Surf. Sci. 2014, 314, 910–918. [CrossRef]
16. Khan, M.; Cao, W. Preparation of Y-doped TiO2 by hydrothermal method and investigation of its visible
light photocatalytic activity by the degradation of methylene blue. J. Mol. Catal. A Chem. 2013, 376, 71–77.
[CrossRef]
17. Zhang, H.; Tan, K.; Zheng, H.; Gu, Y.; Zhang, W.F. Preparation, characterization and photocatalytic activity
of TiO2 codoped with yttrium and nitrogen. Mater. Chem. Phys. 2011, 125, 156–160. [CrossRef]
18. Birks, L.S.; Friedman, H. Particle Size Determination from X-Ray Line Broadening. J. Appl. Phys. 1946, 17,
687–692. [CrossRef]
19. Shi, J.; Chen, J.; Feng, Z.; Chen, T.; Lian, Y.; Wang, X.; Li, C. Photoluminescence characteristics of TiO2 and
their relationship to the photoassisted reaction of water/methanol mixture. J. Phys. Chem. C 2007, 111,
693–699. [CrossRef]
20. Hughes, A.E.; Sexton, B.A. XPS study of an intergranular phase in yttria-zirconia. J. Mater. Sci. 1989, 24,
1057–1061. [CrossRef]
21. Choudhury, B.; Borah, B.; Choudhury, A. Ce–N codoping effect on thestructural and optical properties of
TiO2 nanoparticles. Mater. Sci. Eng. B 2013, 178, 239–247. [CrossRef]
22. Zhu, K.; Zhang, M.; Chen, Q.; Yin, Z. Size and phonon-confinement effects on low-frequency Raman mode
of anatase TiO2 nanocrystal. Phys. Lett. A 2005, 340, 220–227. [CrossRef]
23. Zhao, B.; Wang, J.; Li, H.; Wang, H.; Jia, X.; Su, P. The influence of yttrium dopant on the properties of
anatase nanoparticles and the performance of dye-sensitized solar cells. Phys. Chem. Chem. Phys. 2015, 17,
14836–14842. [CrossRef] [PubMed]
24. Mercado, C.; Seeley, Z.; Bandyopadhyay, A.; Bose, S.; McHale, J. Photoluminescence of dense nanocrystalline
titanium dioxide thin films: Effect of doping and thickness and relation to gas sensing. ACS Appl.
Mater. Interfaces 2011, 3, 2281–2288. [CrossRef] [PubMed]
25. Liu, G.; Wang, X.; Wang, L.; Chen, Z.; Li, F.; Lu, G.; Cheng, H.M. Drastically enhanced photocatalytic activity
in nitrogen doped mesoporous TiO2 with abundant surface states. J. Colloid Interface. Sci. 2009, 334, 171–175.
[CrossRef] [PubMed]
26. Li, D.; Ohashi, N.; Hishita, S.; Kolodiazhnyi, T.; Haneda, H. Origin of visible-light-driven photocatalysis:
A comparative study on N/F-doped and N–F-codoped TiO2 powders by means of experimental
characterizations and theoretical calculations. J. Solid State Chem. 2005, 178, 3293–3302. [CrossRef]
Materials 2017, 10, 600 13 of 13

27. Agrios, A.G.; Gray, K.A.; Weitz, E. Narrow-band irradiation of a homologous series of chlorophenols on
TiO2 : Charge-transfer complex formation and reactivity. Langmuir 2004, 20, 5911–5918. [CrossRef] [PubMed]
28. Wang, N.; Zhu, L.; Huang, Y.; She, Y.; Yu, Y.; Tang, H. Drastically enhanced visible-light photocatalytic
degradation of colorless aromatic pollutants over TiO2 via a charge-transfer-complex path: A correlation
between chemical structure and degradation rate of the pollutants. J. Catal. 2009, 266, 199–206. [CrossRef]
29. Rego, A.M.B.; Ferreira, L.F.V. Photonic and electronic spectroscopies for the characterization of organic
surfaces and organic molecules adsorbed on surfaces. Exp. Methods Phys. Sci. 2001, 38, 269–354.
30. Ferreira, L.F.V.; Machado, I.L.F. Surface Photochemistry: Organic Molecules within Nanocavities of
Calixarenes. Curr. Drug Discov. Technol. 2007, 4, 229–245. [CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like