You are on page 1of 8

Applied Surface Science 302 (2014) 11–18

Contents lists available at ScienceDirect

Applied Surface Science


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

Enhancing the visible light absorption of titania nanoparticles by


S and C doping in a single-step process
M. Scarisoreanu a , I. Morjan a,∗ , R. Alexandrescu a,1 , C.T. Fleaca a , A. Badoi a , E. Dutu a ,
A.-M. Niculescu a , C. Luculescu a , E. Vasile b , J. Wang c , S. Bouhadoun c , N. Herlin-Boime c
a
National Institute for Lasers, Plasma and Radiation Physics, 409 Atomistilor, POB MG-36, Magurele, Bucharest 077125, Romania
b
Metav, Research and Development, 31C.A. Rosetti, Bucharest 020011, Romania
c
IRAMIS/SPAM/LFP, CEA-CNRS URA 2453, CEA de Saclay, Gif sur Yvettes 91191, France

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

Article history: We report the synthesis of carbon coated and sulfur doped titania nanoparticles using a continuous,
Received 5 July 2013 single-step laser pyrolysis technique. We employed air as oxidant and C2 H4 as laser energy transfer agent
Received in revised form 2 December 2013 (sensitizer)/carbon donor, both carrying the TiCl4 vapors as a titania precursor. The volatile (CH3 )2 S2
Accepted 22 January 2014
was used to introduce sulfur as dopant in the nanopowders. The incorporation of C and S atoms in
Available online 3 February 2014
nanopowders with anatase dominant phase and with average particle diameter between 18 and 25 nm
was performed through the addition of S2 (CH3 )2 and C2 H4 to the reactive precursor mixtures. The samples
Keywords:
were characterized by: EDX, XRD, TEM, XPS and UV–Vis spectroscopy. By the introduction of the sulfur
Laser pyrolysis
Nanoparticles
precursor, the anatase-to-rutile ratio within the resulted TiO2 -based nanoparticles decreased, as well as
C their bandgap energy values which are also lower than those of commercial TiO2 Degussa P25.
S-modified TiO2 © 2014 Elsevier B.V. All rights reserved.
Band gap energy
UV–Vis spectroscopy

PACS:
81.16.Mk
61.46.Df
78.67.Bf

1. Introduction Studies have shown that it is preferable non-metal element doping


of TiO2 due to the formation of various secondary impurity phases
Nowadays, is a major interest for photocatalytic semiconduc- (such as metal titanates and/or metal oxides) in the case of dop-
tors in green chemistry in the context of increasing pollution and ing with metals, which reduce the photoactvity and phase stability
consumption of non-renewable resources [1]. Because of its very of anatase titania [11]. Recent reports have shown that TiO2 dop-
good properties (chemical stability, non-toxic and cheap), titania is ing with N, C, S proved high photocatalytic activity under visible
considered one of eco-friendly materials which are used to a wide light due to the fact that the band gap of TiO2 can be narrowed by
range of applications: solar cells [2], wastewater treatment [3], self- these elements doping. Asahi et al. reported that the photocatalytic
cleaning coatings [4] and gas sensors [5]. Unfortunately, anatase, activity in visible light of titania could be enhanced by nitrogen
which is believed to be the most reactive phase of TiO2 , can be doped into the substitutional sites of TiO2 [12]. In our previous
activated only under UV light of wavelengths <387 nm irradiation study we demonstrated an enhancement of the absorption in the
(about 4% of the solar light spectrum) because of its wide band gap visible spectrum for C—coated titania nanoparticles [13]. Regarding
(3.2 eV). From the need to decrease the band gap and to reduce the sulfur doping studies have shown that sulphur could either add to
recombination of photo-generated TiO2 electrons and holes, vari- the TiO2 matrix as a cation or an anion leading to visible light activ-
ous metal and non-metal ions have been doped into titania [6–10]. ity. Umebayashi et al. have reported photocatalytic degradation of
methylene blue in visible-light for S-doped TiO2 as an anion (replac-
ing the lattice oxygen in TiO2 ) [14,15]. On the other hand, Ohno
∗ Corresponding author at: National Institute for Lasers Plasma and Radia- et al. reported that S4+ substitutes some Ti4+ ions of the lattice in
tion Physics, Laboratory of Laser Photochemistry, 409, Atomistilor Str., Bucharest the sulfur-doped TiO2 photocatalyst [16,17]. Recently, the research
077125, Romania. Tel.: +40 21 4574489; fax: +40 21 4574243.. studies who has aimed band gap narrowing by the introduction of
E-mail address: ion.morjan@inflpr.ro (I. Morjan).
1
the different ions co-doping in TiO2 nanopowders have been shown
Dr. Rodica Alexandrescu, a major contributor to these experiments, has deceased
during the preparation of the manuscript.
that the doping of TiO2 with two or three elements could further

0169-4332/$ – see front matter © 2014 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.apsusc.2014.01.135
12 M. Scarisoreanu et al. / Applied Surface Science 302 (2014) 11–18

improve the light absorbance and photocatalytic activity in visible chemical reaction before encountering the laser radiation) which
spectral region [18–20]. is surrounded by a third concentric Ar flow (passing through the
Concerning the carbon-titania nanosystems, they can by classi- second annular nozzle) for the confinement of the reactive mix-
fied at least in two main categories: carbon coated/doped titania ture. Then, they meet the orthogonal laser beam, the reaction
and titania coated carbon. Our system belongs to first category [13] taking place in the volume defined by the intersection of pre-
as well as the photocatalytic C@TiO2 dyade structure [21], whereas cursors with laser radiation. In this case the ethylene molecules
the carbon nanotubes coated with titania nanoparticles (such as absorb laser radiation and then transfer energy to other precur-
those obtained from titanium isopropoxide deposed on the acid- sors by collisions. On leaving the reaction zone, the freshly formed
treated CNTs used for selective enrichment of phosphopeptides nanoparticles are rapidly cooled and collected at the exit of the
[22]) and TiO2 –graphene hybrid nanostructures (tested for Li-ion reaction chamber on a ceramic filter. By using electronic-assisted
batteries [23]) are two examples from the second category. equipments (such as mass flow controllers and feed-back driven
The flame aerosol combustion is the most resembling synthesis throttle valve pressure controller) the method allows a strict con-
technique with our laser pyrolysis for the obtaining of pure titania trol of synthesis parameters: pressure in the reaction chamber, gas
nanoparticles [24] and also for carbon coated TiO2 nanopowders flows and laser power density. In this study we varied the S precur-
[25]. Both techniques employ vapor/aerosol titanium precursors, sor flow, while the other synthesis parameters were kept constant
an oxygen source and various burner/injector devices for driving (470 W and 550 mbar were working values for laser power and
these reactants to the reaction zone. Yet, there are many differ- the pressure in the reaction chamber, respectively). Thus, by vary-
ences that make difficult a direct comparison between them. For ing the ethylene flow as carrier of the sulfur precursor through
example, while in flame aerosol the high temperature required for the first annular nozzle (5, 10 and 15 sccm) and keeping con-
the reaction is obtained from exotermic reaction between oxygen stant TiCl4 -saturated C2 H4 /air annular flow passing through the
and the combustible gas/vapors (hydrogen, hydrocarbons), in the inner central nozzle, we obtained a set of three doped samples
laser pyrolysis an important supplementary heat source provided (named TS-1, TS-2 and TS-3) whose structural, morphological and
by the excited sensitizer molecules due to the laser radiation optical properties were evaluated by comparison with standard
absorption is involved. sample TS-0 (synthesized under otherwise identical conditions in
In this paper, we prepared S, C modified–TiO2 nanoparticles the absence of ethylene-carried dimethyldisulfide). After that, in
by the one-step synthesis method, namely the laser pyrolysis. order to remove the amorphous carbon without inducing transfor-
The effect of sulphur and carbon modification was studied by mation of anatase to rutile phase, the samples were annealed at
using various characterization techniques such as: XRD, XPS, EDX, 450 ◦ C for 3 h, with the heating rate of 2 ◦ C min−1 , in a furnace with
TEM, HRTEM, SAED and UV–Vis diffuse reflectance spectroscopy. circulation of air. For the above mentioned samples, the experi-
Here, we report the further investigation of sulfur-doped TiO2 mental parameters and average S/Ti ratios (EDX measurements)
nanoparticles coated with carbon, analyzing to what extent sul- for the S-doped TiO2 samples, obtained from S2 (CH3 )2 precursor,
fur modification improves the band gap of titania. The optical are found in Table 1. S-doped TiO2 photocatalysts were prepared
properties for the samples showed an absorption shift to longer and were characterized by complementary methods: X-ray diffrac-
wavelengths, thus demonstrating an enhancement of the absorp- tion (XRD), transmission electron microscopy (TEM and HRTEM),
tion in the visible spectrum with respect to pure TiO2 . X-ray photoelectron spectroscopy (XPS), energy dispersive X ray
analysis (EDX) and UV–Vis diffuse reflectance spectroscopy in this
study. X-ray diffraction measurements were carried out at room
2. Experimental temperature on a PANalytical X’Pert PRO MPD X-ray diffractometer.
Incident X-ray radiation was line focused Cu X-ray tube providing
S-doped titania nanoparticles and coated with C were obtained a K␣ wavelength of 1.5418 Å. A curved graphite monochroma-
by laser pyrolysis method using gaseous or liquid with vapor tor and a variable divergence slit working in the fixed irradiated
pressure precursors as follows: TiCl4 —titanium tetrachloride (as area mode were placed in the diffracted beam. HRTEM investi-
precursor for Ti), air (as oxidizing agent), C2 H4 —ethylene (fulfilling gations of the samples were performed using a TECNAI F30 G2
the triple role: sensitizer, carrier of the liquid vapors and C donor S-TWIN microscope operated at 300 kV with Energy Dispersive X
source) and S2 (CH3 )2 —dimethyldisulfide (as precursor for S). The Ray (EDX) Spectrometer with Si(Li) detector. For TEM investiga-
major advantage of using this method is that allows the obtain- tions, very small amount of powder samples were deposited on
ing of doped TiO2 nanoparticles in a single step, providing in the a TEM copper grid covered with a thin amorphous carbon film with
same time the possibility of tailoring nanopowders properties by holes. Surface analysis performed by X-Ray Photoelectron Spec-
controlling the synthesis parameters [13,26–28]. The principle of troscopy (XPS) was carried out on a Quantera SXM equipment,
laser pyrolysis method is based on the resonance of laser radiation with a base pressure in the analysis chamber of 10-9 Torr. The
(CW CO2 , wavelength 10.6 ␮m) with at least one of the reactants. X-ray source was Al K␣ radiation (1486.6 eV, monochromatized)
The incoming reactants and auxiliary gases were introduced in and the overall energy resolution is estimated at 0.65 eV by the full
the reaction chamber through a three concentric cylinders injec- width at half maximum (FWHM) of the Au4f7/2 line. In order to take
tor. Thus, the carried Ti precursor enters in the reaction chamber into account the charging effect on the measured Binding Energies
through the first (inner) nozzle, whereas the ethylene-bubbled S (BEs) the spectra were calibrated using the C1s line (BE = 284.8 eV,
precursor passes through the first annular nozzle (in order to avoid C–C (CH)n bondings) of the adsorbed hydrocarbon on the sample

Table 1
Experimental parameters and EDX measurements for the S-doped TiO2 samples.

Samples  Arf  Air →  C2 H4 →  C2 H4 → Yield EDAX (unheated EDAX (heated


[sccm] TiCl4 [sccm] TiCl4 [sccm] S2 (CH3 )2 [sccm] [g h−1 ] samples) (at%) samples) (at%)

S Ti O C S Ti O C

TS-0 1200 130 20 0 0.56 0 27.19 59.32 13.49 0 31.95 61.48 6.57
TS-1 1200 130 20 5 0.71 1.35 27.27 54.05 17.34 0.48 27.7 63.61 8.24
TS-2 1200 130 20 10 0.69 1.41 28.73 54.57 15.29 0.48 29.61 61.99 7.93
TS-3 1200 130 20 15 0.65 1.54 27.17 55.60 15.89 0.51 31.54 60.1 7.85
M. Scarisoreanu et al. / Applied Surface Science 302 (2014) 11–18 13

Table 2
Crystallographic parameters estimated from XRD measurements for the S-doped TiO2 samples.

Samples  C2 H4 → TiO2 as prepared TiO2 calcinated


S2 (CH3 )2 [sccm]
A (%) R (%) DA (nm) DR (nm) A (%) R (%) DA (nm) DR (nm)

TS-0 0 94 6.0 21.1 ± 0.8 16.4 ± 1.1 93 7 22.3 ± 0.1 16.8 ± 1.3
TS-1 5 72 28 20.7 ± 0.6 18.8 ± 0.7 71 29 18.9 ± 0.5 17.2 ± 1.0
TS-2 10 43 57 21.5 ± 1.0 18.0 ± 0.5 43 57 21.9 ± 0.7 18.9 ± 0.5
TS-3 15 46 54 22.3 ± 0.5 18.3 ± 0.9 46 54 22.4 ± 1.1 19.5 ± 0.6

surface. A dual beam neutralizing procedure (e− and Ar+ ion beams)
has been used to compensate the charging effect in insulating sam-
ples. Energy-dispersive X-ray spectroscopy (EDX) was performed
inside a Scanning Electron Microscope (SEM), FEI Co., model Quanta
Inspect S, 0–30 kV accelerating voltage, using an EDAX Co. SiLi
detector. We used a standardless method for the atomic percent-
ages estimation provided by EDAX Genesis software. The UV–Vis
absorption edge and band gap energies of the samples have been
determined using Kubelka–Munk function or remission, F(R) and
the Tauc plot.

3. Results and discussion

The analyses of the obtained samples (see Tables 1 and 3) reveal


the carbon as ubiquitous element besides the majority titania Fig. 1. The X-ray superposed diffractograms of unheated (pure: TS-0 and S-doped:
phases and also the low concentration of the sulfur in TS-1, TS-2, TS-1, TS-2 and TS-3) and heated (pure: TS-0(H) and S-doped: TS-1(H), TS-2(H) and
TS-3(H)) TiO2 samples.
TS-3 nanoparticles. The scarce sensitivity of the S and C amounts
in the resulted nanopowders on the precursor flow rates can be
related with the complex chemical reactions which occur in the between the gas flows and sulfur content, and because the intro-
laser pyrolysis flame environment. Thus, the TiCl4 , C2 H4 and, for TS- duction of dimethyl disulfide vapors is made using ethylene, its
1, TS-2 and TS-3 experiments, (CH3 )2 S2 molecules will compete for growth is also correlated with increasing carbon content. The doped
the limited amount of available oxygen in the reaction zone. As will samples (both heated and unheated) shows a higher content of sul-
be discussed in the next paragraph regarding XRD diffractograms, fur and carbon as compared to the etalon sample (TS-0), which
the only solid product resulted from TiCl4 in these experiments decreases the after the calcination while keeping the same trend.
is TiO2 (and no metallic Ti, TiC or Ti suboxides) which means that These results are also in good connection with the results obtained
this precursor is more reactive towards oxygen than C2 H4 in our from XPS analysis (see below).
specific conditions. Also, the quantity of sulfur find in the TS-1–3 The authors from Ref. [25] observed an increasing carbon con-
raw powders is considerably lower than those injected as (CH3 )2 S2 , tent in the titania based nanopowders with increasing the fuel rate
which implies that the great majority was lost by oxidation as while maintaining constant the oxygen (as air) flow. In our case, the
gaseous sulfur oxides. Moreover, by increasing the amount of TS-1 samples have a higher carbon content than TS-0 as expected
sulfur precursor introduced, the amount of this element detected due to the increasing C2 H4 total flow from 20 to 25 sccm (see
in the collected nanopowders also increases but in a much lower Table 1). Further increasing the etylene amount to 30 and 35 sccm
proportion. The excited (after laser photons absorption) ethylene (sample TS-2 and TS-3, respectively) the carbon from nanopowders
molecules can follow various reaction pathways: to react with the diminishes with a tendency towards saturation. This behaviour
limited quantity of the available O2 with the formation of CO2 , CO, can be explained by considering the even increasing amount of
H2 O and/or H2 or to decompose to solid carbon (and H2 ) which (CH3 )2 S2 vapors carried by the supplementary etylene flow. The
can be found in all the raw powders. Also, part of C2 H4 molecules dimethyl disulfide is a well-known anti-coke agent in hydrocar-
can escape unharmed from the pyrolysis flame (by relaxation after bon cracking processes [31]; thus, we assume a similar role in the
collisions with inert species). On the other hand, the (CH3 )2 S2 diminishing of elemental carbon content in our S-TiO2 nanopow-
is known to readily decompose to CH3 S methylthiyl radical [29] ders.
which can react with C2 H4 or with O2 [30], depending of the
local microenvironment in flame. The chlorine resulted from TiCl4 3.2. XRD measurements
oxidation can also react with C2 H4 , resulting chlorinated derivates,
which are also involved in the oxidation/decomposition reac- The XRD patterns of the samples as synthesized and after the
tions. The presence of various titanium, chloride, sulfur, oxygen, thermal treatment (calcinated) are presented in Fig. 1. The diffrac-
carbon and hydrogen-containing species involved in the com- tograms show a mixture of two titania phases: the anatase phase
plex homogeneous/heterogeneous interconnected/competitive (JCPDS 21-1272) and the rutile phase (JCPDS 21-1276). The anatase
reactions occurring in the far-from-equilibrium local conditions and rutile phase percentages were evaluated according to the
encountered in the laser-induced flame makes very difficult to empiric Spurr and Myers formula using the (1 0 1) anatase peak
estimate the exact amount of carbon and sulfur in the resulted and the (1 1 0) of rutile phase, respectively. [32] The S-free TiO2
titania-powders based on the process parameters. sample exhibits a predominant anatase phase, 94%, (see Table 2).
The mean crystallites sizes were estimated via the Scherrer for-
3.1. EDX measurements mula using the instrumental corrected full width at half maximum
(FWHM) averaging over five reflections of each titania phases. The
The results of EDX semi quantitative elemental analysis are standard reference material used for the instrumental correction
shown in Table 1. Results of the analyses indicate a good correlation was alumina (corundum) 1976a from NIST The crystallites size
14 M. Scarisoreanu et al. / Applied Surface Science 302 (2014) 11–18

when calculated at different Bragg angles was found to remain


roughly unaltered for both titania phase, indicating the absence
of strain. Crystallites sizes lie between 16 and 22 nm. The pres-
ence of graphitic carbon or other sulfur compounds seems to be
not revealed by XRD, most likely due to a high dispersion of the
two elements. Thus, we can conclude that the main effect of the S
doping (S presence in the reaction step) is the significant increase of
rutile phase percentage in comparison with the S-free sample TS-0.
A possible explanation can be the partial transformation of anatase
to rutile phase at higher temperatures in the laser pyrolysis flame,
when larger amount of ethylene (carrying the S-precursor vapors)
were introduced. A complementary/alternative explanation of high
rutile amount in S-containing samples can be the presence of low
concentration of gaseous oxygen during particle formation who
leads to creation of anion vacancies which, in turn, enhances the
possibility of anatase-to-rutile phase transformation [24]. The ther-
mal treatment preserve almost the same proportion of the titania
phases and of their crystallite sizes.

3.3. TEM analysis and particle diameter distributions

TEM analyses for undoped and S-doped TiO2 nanoparticles


(Fig. 2a–c) show nanoparticles with irregular morphologies and
size, mostly of polyhedral shapes as compared with round nanopar-
ticles in the case of sample TS-0 and TS-3, respectively. The particles
seem to be embedded in the disordered/amorphous layer which
explain the slight differences in average particle size (26 nm)
compared with the XRD results (22 nm). On the other hand, the
introduction of different impurities or sulfur species which can-
not be detected by XRD analysis may be another explanation. The
SAED analysis (given as inset in Fig. 2(a) and (b)) shows the crys-
talline diffraction rings for the sample with the highest S doping
level (TS-3), simultaneously with the appearance of diffraction
rings corresponding to rutile phase. The major anatase presence
(d = 3.53 Å, (1 0 1)) is suggested in sample TS-0. Increasing sulfur
doping can be related to a slight increase in average particle size
from 18.6 to 26.8 nm (inset Fig. 2(a) and (c)) but not seem to induce
major changes to the morphology of nanoparticles. Morphologi-
cal and structural characterization was performed to obtain more
details about the geometrical size, morphology and nanostructure
of the samples by high resolution transmission electron microscopy
(HRTEM). Fig. 3 displays the HRTEM images for S-TiO2 nanopar-
ticles (TS-3). A mixture of anatase and rutile (d = 3.52 and 3.24
interplanar distances, respectively) surrounded by graphitic car-
bon (d = 3.8 Å) seems to be present in sample obtaining a good
correlation with SAED and XRD analysis.

3.4. XPS analysis

X-ray photoelectron spectroscopy (XPS) analysis technique was


used to determine the chemical states of the elements present on
the surface and, after quantitative analysis, to find the element and
the chemical state relative concentrations as well. After scanning
survey XPS spectra, the high resolution photoelectron spectra of
the most prominent XPS transitions (C1s, O1s, Ti2p and S2p) were
recorded for the TS-0, TS-1, TS-2 and TS-3 samples. It is appropriate
to note here that all the calculations were performed assuming
that the samples were homogeneous within the XPS detected
volume. We have to emphasize that the errors in our quantitative
analysis (relative concentrations) were estimated in the range of
±10%, while the accuracy for Binding Energies (BEs) assignments
was ±0.2 eV.
For the fitting process we used the capabilities of Mul-
tipak software and the recommendations of the ISO—TC201 Fig. 2. TEM images of the TiO2 nanoparticles: as synthesized (a: TS-0 and b: S-doped
sample TS-3) and a S-doped sample after heating at 450 ◦ C (c: TS-3(H)). The particle
(“Surface Chemical Analysis”—SCA Technical Committee), the
size distributions for TS-0 and TS-3(H) samples and the SAED images for TS-0 and
Sub-committee SC3 “Data management and treatment”. Thus, TS-3 samples are presented as insets in the corresponding TEM images.
M. Scarisoreanu et al. / Applied Surface Science 302 (2014) 11–18 15

Fig. 3. HRTEM images of the sample TS-3 (with the highest S doping) revealing both anatase and rutile nanoparticles (a) and carbon layers embedding a faceted TiO2
nanoparticles (b).

we fitted the photoelectron lines by using the Voigt functions Table 3


Element relative concentrations (at%) for S-doped TiO2 samples from XPS spectra.
(a mixture of Gaussian and Lorentzian functions) and imposed
the required constraints on the relative intensities of the doublet Samples C1s O1s S2p Ti2p
transitions as well as on the spin–orbit parameters. TS-0 29.1 52.1 0 18.8
The XPS spectra for Ti2p zone (Fig. 4a) shows the presence of Ti4+ TS-O(H) 22.8 55.4 0 21.8
signature with the two peaks at 464.5 and 458.7 eV, correspond- TS-1 38.9 41.8 2.3 17.0
ing to Ti2p1/2 and Ti2p3/2 binding energies, found also in the S-(un) TS-1(H) 16.3 60.6 1.9 21.2
TS-2 33.7 45.5 2.2 18.6
doped TiO2 samples obtained by acid catalysis hydrolysis method
TS-2(H) 14.6 59.6 2.7 23.1
in the presence of thiourea [33]. In the same figure, the XPS regions TS-3 28.5 49.4 1.9 20.2
of C1s from the TS0 sample before (a) and after (b) air calcination TS-3(H) 13.3 61.9 2.2 22.6
are presented. One can observe an increase of the intensity ratios
of the C–O (at 288.4 eV) and C=O (at 286.5 eV) corresponding peaks
versus those from C=C/C–H bonds (at 284.6 eV) after annealing at is witnessed not only by his signature in EDS and XPS spectra,
450 ◦ C in air. This trend can be explained by the partial oxidative but is also visible as turbostratic graphitic layers around the TiO2
conversion of the carbonaceous layer which surrounds the TiO2 nanoparticles in HR-TEM image from Fig. 3. As expected, the car-
nanoparticles to CO2 and H2 O, as can be verified also from the XPS bon content extracted from the EDS spectra (Table 1) is much lower
and EDX atomic composition estimations. that from XPS analyses (Table 3), due to the fact that XPS is sensi-
The superimposed S2p spectra show a quite different behaviour tive only up to few nm depth from the surface, whereas the EDS is
for the samples before and after calcination. The raw samples originated for photoelectrons emitted from deeper regions of the
(before calcination) display the features characteristic to prevalent sample (after electron bombardment). The few nm carbon coatings
zerovalent sulfur S0 (∼163.8 eV) mixed with S4+ (∼166.8 eV) and on the titania nanoparticles will be thus more “visible” for the XPS
S6+ (∼168.7 eV) as minor phases (Fig. 5a). The peak around 162 eV, measurements. Making the same comparison for the atomic sulfur
characteristic for the lowest sulfur oxidation state as S2− replac- content in S-doped TiO2 powders, even if the relative difference
ing the O2− ions in the TiO2 crystalline lattice and reported in Ref. between the values obtained from the two analysis techniques is
[34] was not found. Conversely, the S-doped calcinated samples smaller than that found for carbon, a similar conclusion can be
only exhibit the peak corresponding to the higher binding energies considered—the sulfur is more concentrated on/near the particles
at ∼168.7 eV (Fig. 5b), which is characteristic to the highest sul- surfaces. This tendency can be correlated with the separate intro-
fur oxidation state as SO4 2− sulfate type structure. This fact can duction of the dimethyldisulfide vapors on the annular intermediar
be explained by the post-synthesis oxidation of the S0 and S4+ nozzle, thus the freshly hot titania-based nanoparticles formed on
species during the air annealing treatment. Near the same high the center in the laser pyrolysis flame adsorb the sulfur species
binding energy was suggested to correspond to hexacoordinated resulted from decomposition/oxidation of this precursor.
S6+ ion existing as sulphur substitutionally replacing Ti4+ in the S-
doped TiO2 [33]. The presence of zerovalent and positive oxidation 3.5. Characterization by UV–Vis transmission spectroscopy
states of sulfur (and the absence of S2− ) in the doped samples can
be expected since the CH3 –S–S–CH3 precursor have sulfur atoms The optical properties of S-doped/undoped TiO2 samples coat-
in 1-oxidation state that change to higher oxidation states in the ing with C were investigated by measuring their band-gap energy
presence of the oxygen from air in the reactive mixture from the (Eg ). This was determined from the reflectance [F(R)] spectra using
laser-induced flame. Also, the existence of intermediate S0 and S4+ the Kubelka–Munk (K–M) formalism and the Tauc plot [35]. Fig. 6
states can be correlated with the limited amount of the available shows the plots of Kubelka–Munk functions F(R)1/2 vs. E (eV) of
oxygen due to the abundant presence of the ethylene species which the S-doped TiO2 powders with different sulfur contents (TS-1,TS-
compete (as well as the TiCl4 molecules) for the reaction with oxy- 2,TS-3) as compared with pure TiO2 sample (TS-0). Introduction
gen in the laser pyrolysis flame. Moreover, the oxigen deficiency of S, C anions narrows the band gap of TiO2 and this feature was
of the flame can be proved by the fact that some of the ethylene more evident as compared to undoped sample. For the P25 TiO2
undergo not oxidation (to CO2 and/or CO), but decomposition to nanoparticles, used in the catalytic light-induced reactions studies
carbon, which was found in all the samples. The carbon presence as a model TiO2 with high photocatalytic performance, the band
16 M. Scarisoreanu et al. / Applied Surface Science 302 (2014) 11–18

Fig. 4. Ti2p for TS-0(H) sample (a) and C1s deconvoluted core level spectra for TS-0 Fig. 5. Deconvoluted S2p XPS spectra for heated (a) and unheated (b) TS-2 samples;
(b) and TS-0(H) (c). superposed S2p XPS spectra for all (heated and unheated) doped samples (c).

gap energy was observed at 3.2 eV corresponding to the absorption


edge 388 nm from anatase TiO2 . For undoped as compare with the whereas anatase exhibits only an indirect bandgap at 3.23 eV [37].
S-doped samples prepared by laser pyrolysis the Eg values are 3.2 The bandgap of all S-doped samples was found to be around 2.7 eV,
and 2.7 eV, respectively. The energy gaps of S-doped TiO2 photo- which is lower than both those of anatase and rutile. This consid-
catalysts are similar in value, yet narrower than the reference TS-0 erable shifting band gap energy creates the prerequisites for an
without sulfur. The corresponding absorption edges are 388 and improved photocatalytic efficiency even if these sulfur contain-
459 nm, respectively, which indicate that the S-doped TiO2 samples ing titania powders have a considerable content of the rutile—the
coated with C have photoactivity in the visible region, which can so-called less reactive phase. However, it is well know that pure
be translated to an enhancing photocatalytic activity in this region. TiO2 commercial samples Degussa P25 used as high photoactive
The greatest reduction in the band gap is due to the presence of sul- standard material in the study of photocatalytic reaction contains
fur in TiO2 nanopowders considering that in the undoped sample around 25% rutile [38]. Other authors [39] presented a TiO2 sample
(TS-0) carbon is already present. There is currently great concern to containing 52.9% rutile with a high efficiency in the photodegrada-
elucidate the inserting mechanism of the sulfur and carbon in TiO2 tion of 4-chlorophenol process. Moreover, there are studies which
nanoparticles and their influence to their photocatalytic proper- show that anatase and rutile have a similar photocatalytic activity
ties since both S and C could be added either as cation (replacing for organic degradation or water oxidation (with the same amount
Ti ions) or a anion (by replacing the lattice oxygen) in the TiO2 of electron scavenger on the catalyst surfaces) [40]. In conclusion,
matrix [14,15,36]. From literature it is well known that rutile has both band gap value and the amount of titania phases, corroborated
a direct band gap at 3.06 eV and an indirect bandgap at 3.10 eV, with other factors contribute to an optimal photocatalytic effect.
M. Scarisoreanu et al. / Applied Surface Science 302 (2014) 11–18 17

[6] R. Asahi, T. Morikawa, Nitrogen complex species ant its chemical nature in TiO2
for visible-light sensitized photocatalysis, Chem. Phys. 339 (2007) 57–63.
[7] X.T. Hong, Z.P. Wang, W.M. Cai, F. LU, J. Zhang, Y.Z. Yang, N. Ma, Y.J. Liu, Visible-
light-activated nanoparticle photocatalyst of iodine-doped titanium dioxide,
Chem. Mater. 17 (2005) 1548–1552.
[8] J.C. Yu, J. Yu, W. Ho, Z. Jiang, L. Zhang, Effects of F-doping on the photocatalytic
activity and microstructures of nanocrystalline TiO2 powders, Chem. Mater. 14
(2002) 3808–3816.
[9] T.K. Ghorai, S.K. Biswas, P. Pramanik, Photooxidation of different organic dyes
(RB, MO, TB, and BG) using Fe(III)-doped TiO2 nanophotocatalyst prepared by
novel chemical method, Appl. Surf. Sci. 254 (2008) 7498–7504.
[10] Y. Cong, J. Zhang, F. Chen, M. Anpo, D. He, Preparation, photocatalytic activ-
ity, and mechanism of nano-TiO2 Codoped with nitrogen and iron(III), J. Phys.
Chem. C 111 (2007) 10618–10623.
[11] P. Periyat, K.V. Baiju, P. Mukundan, P.K. Pillai, K.G.K. Warrier, Aqueous colloidal
sol–gel route to synthesize nanosized ceria-doped titania having high surface
area and increased anatase phase stability, J. Sol-Gel Sci. Technol. 43 (2007)
299–304.
[12] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Visible-light photocatalysis
in nitrogen-doped titanium oxides, Science 293 (2001) 269–271.
[13] M. Scarisoreanu, R. Alexandrescu, I. Morjan, R. Birjega, C. Luculescu, E. Popovici,
E. Dutu, E. Vasile, V. Danciu, N. Herlin-Boime, Structural evolution and optical
properties of C-coated TiO2 nanoparticles prepared by laser pyrolysis, Appl.
Surf. Sci. 278 (2013) 295–300.
Fig. 6. Absorbance spectra for the indirect electronic transition F(R)1/2 vs. E (eV) for [14] T. Umebayashi, T. Yamaki, S. Tanala, K. Asai, Visible light-induced degradation
the S-doped as compared with undoped raw TiO2 samples. of methylene blue on S-doped TiO2 , Chem. Lett. 32 (2003) 330.
[15] T. Umebayashi, T. Yamaki, S. Yamamoto, A. Miyashita, S. Tanala, T. Sumita,
K. Asai, Sulfur-doping of rutile-titanium dioxide by ion implantation: photo-
4. Conclusions current spectroscopy and first-principles band calculation studies, J. Appl. Phys.
93 (2003) 5156.
[16] T. Ohno, M. Akiyoshi, T. Umebayashi, K. Asai, T. Mitsui, M. Matsumura, Prepa-
In summary, we have succeeded in preparing S-doped TiO2 ration of S-doped TiO2 photocatalysts and their photocatalytic activities under
photocatalysts by using a simple one-step method. This approach visible light, Appl. Catal., A 265 (2004) 115.
can efficiently control the sulfur dopant atoms. TiO2 nanoparticles [17] T. Ohno, Preparation of visible light active S-doped TiO2 photocatalysts and
their photocatalytic activities, Water Sci. Technol. 49 (2004) 159–163.
modified with sulfur and carbon were synthesized by oxidative [18] K. Lv, H. Zuo, J. Sun, et al., (Bi, C and N) codoped TiO2 nanoparticles, J. Hazard.
laser pyrolysis of the titanium tetrachloride, dimethyldisulfide and Mater. 161 (2009) 396–401.
ethylene precursors. Sulfur-doped TiO2 nanoparticles coated with [19] H. Xu, L. Zhang, Controllable one-pot synthesis and enhanced visible light
photocatalytic activity of tunable C–Cl-codoped TiO2 nanocrystals with high
carbon were shown to exhibit different structural, morphologi- surface area, J. Phys. Chem. C 114 (2010) 940–946.
cal and optical properties when compared to undoped TiO2 . The [20] P. Wang, P.-S. Yap, T.-T. Lim, C–N–S tridoped TiO2 for photocatalytic degrada-
effects of the introduction of C and S in TiO2 nanoparticles were tion of tetracycline under visible-light irradiation, Appl. Catal., A 399 (2011)
252–261.
the decreased proportion of anatase phase (from 94 to 43%) simul- [21] L. Zhao, X. Chen, X. Wang, Y. Zhang, W. Wei, Y. Sun, M. Antonietti, M.-M.
taneously with the slightly growth of average particle size (in the Titirici, One-step solvothermal synthesis of a carbon@TiO2 dyade structure
18–26 nm range). TEM analysis for the samples reveals a mor- effectively promoting visible-light photocatalysis, Adv. Mater. 22 (2010) 317–
3321.
phology consisting of titania-based core– turbostratic carbon shell
[22] Y. Yan, J. Lu, C. Deng, X. Zhang, Facile synthesis of titania nanoparticles coated
nanoparticles. The successful incorporation of the dopant into the carbon nanotubes for selective enrichment of phosphopeptides for mass spec-
host was confirmed by X-ray photoelectron spectroscopy and the trometry analysis, Talanta 107 (2013) 30–35.
elemental analysis quantified this addition (with the features S6+ at [23] D. Wang, D. Choi, J. Li, Z. Yang, Z. Nie, R. Kou, D. Hu, C. Wang, L.V. Saraf, J.
Zhang, I.A. Aksay, J. Liu, Self-assembled TiO2 –graphene hybrid manostructures
relative concentrations about 2 at%). All the doped samples showed for enhanced Li-ion insertion, ACS Nano 3 (2009) 907–914.
significantly a lower band gap energy (Eg = 2.7 eV) than the com- [24] G.P. Fotou, S. Vemury, S.E. Pratsinis, Synthesis and evaluation of titania powders
mercial P25 Degussa TiO2 sample (Eg = 3.2 eV). for photodestruction of phenol, Chem. Eng. Sci. 49 (1994) 4939–4948.
[25] A.K. Bhanwala, A. Kumar, D.P. Mishra, J. Kumar, Flame aerosol synthesis and
The introduction by one step method of other or more ions characterization of pure and carbon coated titania nanopowder, J. Aerosol Sci.
codoped TiO2 nanoparticles, correlated with photocatalytic effi- 40 (2009) 720–730.
ciency tests for different pollutants are the concerns that we will [26] R. Alexandrescu, I. Morjan, M. Scarisoreanu, R. Birjega, E. Popovici, I. Soare,
L. Gavrila-Florescu, I. Voicu, I. Sandu, F. Dumitrache, G. Prodan, E. Vasile, E.
exploit in future studies. Figgemeier, Structural investigations on TiO2 and Fe-doped TiO2 nanoparticles
synthesized by laser pyrolysis, Thin Solid Films 515 (2007) 8438–8445.
Acknowledgments [27] R. Alexandrescu, I. Morjan, M. Scarisoreanu, R. Birjega, C. Fleaca, I. Soare, L.
Gavrila, V. Ciupina, W. Kylberg, E. Figgemeier, Development of the IR laser
pyrolysis for the synthesis of iron-doped TiO2 nanoparticles: structural prop-
The first author acknowledges financial support in the frame of erties and photoactivity, Infrared Phys. Technol. 53 (2010) 94–102.
the Projects: PD 106 and C1-07 CEA. [28] R. Alexandrescu, M. Scarisoreanu, I. Morjan, R. Birjega, C. Fleaca, C. Luculescu, I.
Soare, O. Cretu, C.C. Negrila, N. Lazarescu, V. Ciupina, Preparation and char-
acterization of nitrogen-doped TiO2 nanoparticles by the laser pyrolysis of
References N2 O-containing gas mixtures, Appl. Surf. Sci. 255 (2009) 5373–5377.
[29] A.G. Vandeputte, M.F. Reyniers, G.B. Marin, Theoretical study of the thermal
[1] J. Santhanalakshmi, R. Komalavalli, P. Venkatesan, Photo catalytic degradation decomposition of dimethyl disulfide, J. Phys. Chem. A 114 (2010) 10531–10549.
of chloropyrifos, endosulphon, imidocloprid and quinolphos by nano crys- [30] L. Zhu, J.W. Bozzelli, Kinetics of the multichannel reaction of methanethiyl
talline TiO2 —a kinetic study with pH and mass effects, Nanosci. Nanotechnol. radical (CH3S*) with 3 O2 , J. Phys. Chem. A 110 (2006) 6923–6937.
2 (2012) 8–12. [31] J. Wang, M.-F. Reyniers, G.B. Marin, Influence of dimethyl disulfide on coke for-
[2] A. Hossain, J.R. Jennings, C. Shen, J.H.Z. Pan, Y. Koh, N. Mathews, Q. Wang, CdSe- mation during steam cracking of hydrocarbons, Ind. Eng. Chem. Res. 46 (2007)
sensitized mesoscopic TiO2 solar cells exhibiting >5% efficiency: redundancy of 4134–4148.
CdS buffer layer, J. Mater. Chem. 22 (2012) 16235–16242. [32] R.A. Spurr, H. Myers, Quantitative analysis of anatase-rutile mixtures with an
[3] W. Zhang, L. Zou, L. Wang, Photocatalytic TiO2 /adsorbent nanocomposites pre- X-ray diffractometer, Anal. Chem. 9 (1957) 760–762.
pared via wet chemical impregnation for wastewater treatment: a review, Appl. [33] X. Liu, X. Chen, A visible light response TiO2 photocatalyst realized by cationic S-
Catal., A 371 (2009) 1–9. doping and its application for phenol degradation, J. Hazard. Mater. 152 (2008)
[4] H. Yaghoubi, N. Taghavinia, E.K. Alamdari, Self cleaning TiO2 coating on poly- 48–55.
carbonate: surface treatment, photocatalytic and nanomechanical properties, [34] T. Umebayashi, T. Tamaki, H. Itoh, K. Asai, Band gap narrowing of titanium
Surf. Coat. Technol. 204 (2010) 1562–1568. dioxide by sulfur doping, Appl. Phys. Lett. 81 (2002) 454–456.
[5] W.J. Moon, J.H. Yu, G.M. Choi, Selective gas detection of SnO2 -TiO2 gas sensors, [35] F. Fabregat-Santiago, I. Mora-Sero, G. Garcia-Belmonte, J. Bisquert, Cyclic
J. Electroceram. 13 (2004) 707–713. voltametry studies of nanoporous semiconductors. Capacitive and reactive
18 M. Scarisoreanu et al. / Applied Surface Science 302 (2014) 11–18

properties of nanocrystalline TiO2 electrodes in aqueous electrolyte, J. Phys. [39] S. Bakardjieva, J. Subrt, V. Stengl, M.J. Dianez, M.J. Sayagues, Photoactivity
Chem. B 107 (2003) 758. of anatase–rutile TiO2 nanocrystalline mixtures obtained by heat treat-
[36] C. Di Valentini, G. Pacchioni, A. Selloni, Theory of carbon doping of titanium ment of homogeneously precipitated anatase, Appl. Catal., B (2005) 193–
dioxide, Chem. Mater. 17 (2005) 6656–6665. 202.
[37] K.-H. Hellwege, O. Madelung, Landolt-Börnstein, Band 17: Halbleiter, Springer, [40] Q. Sun, Y. Xu, Evaluating intrinsic photocatalytic activities of anatase and
Berlin, 1984. rutile TiO2 for organic degradation in water, J. Phys. Chem. C 114 (2010)
[38] T. Ohno, K. Sarukawa, K. Tokieda, M. Matsumura, Morphology of a TiO2 pho- 18911–18918.
tocatalyst (Degussa, P-25) consisting of anatase and rutile crystalline phases, J.
Catal. 203 (2001) 82–86.

You might also like