You are on page 1of 8

Applied Surface Science 336 (2015) 226–233

Contents lists available at ScienceDirect

Applied Surface Science


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

Laser oxidative pyrolysis synthesis and annealing of TiO2


nanoparticles embedded in carbon–silica shells/matrix
C.T. Fleaca a,b , M. Scarisoreanu a,∗ , I. Morjan a , C. Luculescu a , A.-M. Niculescu a , A. Badoi a ,
E. Vasile c , G. Kovacs d
a
National Institute for Plasma, Laser and Radiation Physics (NILPRP), Atomistilor 409, P.O. Box MG 36, R-077125 Magurele, Bucharest, Romania
b
“Politehnica” University of Bucharest, Physics Department, Independentei 313, Bucharest, Romania
c
“Politehnica” University of Bucharest, Faculty of Applied Chemistry and Materials Science, Department of Oxide Materials and Nanomaterials,
Gh. Polizu 1-7, Bucharest, Romania
d
“Babes-Boyai” University, Faculty of Chemistry and Chemical Engineering, Arany Janos 11, Cluj-Napoca, Romania

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

Article history: Titania nanoparticles containing a mixture of anatase and rutile phases (with mean crystalline sizes up
Received 27 June 2014 to 24 nm) covered with/embedded in carbon/silica thin layers or matrix were obtained in a single step
Received in revised form 23 October 2014 using laser oxidative pyrolysis. Titanium tetrachloride and hexamethyldisiloxane (HMDSO) vapors were
Accepted 19 November 2014
separately introduced into the reaction zone – both together with the laser-absorbing agent (sensitizer)
Available online 3 December 2014
ethylene – which acts also as carbon source – and the oxidant (air) – through the inner and the concentric
nozzle, respectively. By increasing the air flow through the annular nozzle, while keeping constant the
PACS:
TiC4 , inner air, HMDSO and C2 H4 flows, the atomic carbon concentration as well as the rutile to anatase
81.16 Mk
82.33 Vx
ratio in the resulted nanopowders decrease. A much brighter and extended flame was observed for the
81.07 Wx experiment involving the greatest air flow. The Ti/Si atomic ratio in the resulted nanocomposites was
61.46 Df higher than that from the introduced precursors (1.8), indicating a partial siloxane conversion to silica.
62.23 Kn The annealed powders (at 450 ◦ C to further carbon content reducing) exhibit a lower bandgap energy
81.16 Hc than those of the reference sample without silica (and also lower than the commercial Degussa P25
nano-TiO2 ).
Keywords:
Laser pyrolysis
© 2014 Elsevier B.V. All rights reserved.
Core–shell nanoparticles
Anatase/rutile mixture

1. Introduction with very different morphologies such as aerogels [1] or nanopat-


terned nanodots/nanowires [2], having as targeted application the
Titanium dioxide-based nanomaterials have found a very wide performant photocatalytic oxidation of various organic molecules.
range of applications, from pigments to water remediation espe- Moreover, silica–titania material was found to able to photode-
cially due to the optical and catalytic properties of the two main compose the water (accelerated by RuS2 doping) with H2 and O2
TiO2 phases, anatase and rutile, enhanced by the high surface generation under UV light [3]. The same sol–gel approach was
area and reactivity provided by the nanosize dimension. More- used to obtain materials with various other light-related applica-
over, controlled combination of titania with other substances can tions: superhydrophilic [4] or antireflective self-cleaning films [5],
further enhance their performances through a synergistic interac- solar [6] and transparent UV-shielding poly(methyl methacrylate)
tion. Benefiting from the specific advantages of both components: (PMMA)-based composites [7]. The gas-phase methods starting
semiconductive behavior and crystallinity from TiO2 and porous, from various Ti and Si containing vaporized/sprayed precursors
protective characteristic associated with the hydrophilicity and allow the synthesis of TiO2 /SiO2 films and powders in a contin-
thermal resistance of SiO2 , the titania–silica (nano)composites are uous mode and much faster than in the sol–gel methods. Flame
a representative class of such materials. Sol–gel is the most used combustion is the most used gas-phase method and was employed
technique for their fabrication, allowing the synthesis of materials for the obtaining of TiO2 doped SiO2 films for waveguide applica-
tions (from titanium tetraisopropoxide (TTIP) and [(CH3 )2 SiO]4 in
a O2 /H2 flame) [8], TiO2 @SiO2 core–shell composite nanoparticles
∗ Corresponding author. Tel.: +40 21 4574489; fax: +40 21 4574243. (from TiCl4 and SiCl4 using N2 -shielded oxy-hydrogen flame and a
E-mail address: monica.scarisoreanu@inflpr.ro (M. Scarisoreanu). multitubular injector) [9], column-shaped TiO2 –SiO2 phocatalytic

http://dx.doi.org/10.1016/j.apsusc.2014.11.106
0169-4332/© 2014 Elsevier B.V. All rights reserved.
C.T. Fleaca et al. / Applied Surface Science 336 (2015) 226–233 227

nanoparticles/fibers deposed on glass (from TTIP, tetraethoxysilane photocatalyst by the synergetic effect of carbon and silver co-
(TEOS) and a liquefied petroleum gas (LPG) – air/O2 /N2 combustible doping [29]. Laser pyrolysis gas-phase technique was also used
mixture) [10] and also binary SiO2 /TiO2 mixed oxide and composite to obtain in a single step oxygen-deficient blue Ti/Si/C/O compos-
nanoparticles from N2 -diluted TTIP [11] or TiCl4 [12] and HMDSO ite nanopowders from TTIP aerosol mixed with SiH4 gas, without
(or also SiCl4 or SiBr4 [13]) in an O2 /N2 /CH4 flame at atmospheric external oxidative agent [30]. This approach (due to the fixed
pressure. atomic ratio between Ti and O atoms in the TTIP precursor) does not
Other material that can improve the properties of titania is the allow the independent variation of the introduced oxygen. More-
carbon, which was reported to stabilize the anatase phase [14] and, over, it is known that the silane reacts explosively when encounters
as a thin coating under porous form, to enhance the adsorption the molecular oxygen, and consequently, a considerable amount of
of the organic molecules, without reducing too much the amount inert gas it is necessary to control this reaction which dramatically
of UV radiation that reaches the TiO2 photocatalytic surface [15]. diminishes the resulted oxide powder productivity.
Due to the formation of titania–graphite heterojunction which In this paper, we report the using the laser pyrolysis method to
enhanced the separation of photogenerated electrons and holes, study the influence of the supplementary outer air oxidant intro-
the TiO2 anatase/carbon nanowalls (CNWs) material presented a duction together with the HMDSO vapors (carried by the same C2 H4
higher UV light-induced photodegradation of phenol compared flow after bubbling) when the TiCl4 inner flow and their accompa-
with TiO2 nanotubes [16]. Similar deposition of TiO2 coatings by nied air and C2 H4 sensitizer flows are also kept constant on the
metal organic chemical vapor deposition (MO CVD) on activated structure and optical properties of the resulted TiO2 @SiO2 /C nano-
carbon (using titanium tetrabutoxide (TBOT) precursor instead of composites.
TTIP) yielded a recyclable photocatalyst, successfully tested for
methylorange dye degradation in aqueous suspension under UV 2. Experimental
irradiation [17]. Moreover, the presence of carbon provides a strong
absorption in visible light (up to 750 nm) and also induces a visible- Titanium tetrachloride (>99%) and hexamethyldisiloxane (min.
light photodegradation activity [18] as compared with pure anatase 99%) volatile liquids were purchased from Merck, whereas ethyl-
which requires the more energetic UV photons to perform the pho- ene (99.95%) sensitizer, synthetic air (80% nitrogen, 20% oxygen)
tocatalytic role. Thus, by a very thin (few graphene layers) coating and auxiliary argon (99.998%) were purchased from Linde Gas.
deposed on the TiO2 surface resulted a hybrid material better than Our laser pyrolysis experimental set-up was described in [31,32],
the pure TiO2 Degussa P25 reference under UV light irradiation, when a similar three concentric nozzles injector was used for the
presenting also a high activity under visible light due to the cou- synthesis of titania-based nanoparticles. Briefly, the TiCl4 vapors
pling between the ␲ electrons of graphite-like carbon and those carried from a bubbler by an air–ethylene mixture were introduced
from the conduction band of TiO2 [19]. Visible-light activity was from the bottom of a vertical injector through the central noz-
reported too for sol–gel synthesized carbon-doped TiO2 where the zle, the ethylene-carried HMDSO vapors and various air flows
dopant was provided by the butyl groups from the TBOT precursor through the second annular nozzle, whereas an Ar confinement
[20]. Other authors used the one-pot hydrothermal method to form flow pass through the third annular nozzle. These flows orthog-
TiO2 @C (from TiCl4 and glucose) with greatly enhanced photocat- onally encounter a focused continuous-wave infrared laser beam
alytic activity under visible light compared with noncarbon-TiO2 ( = 10.6 ␮m) centered at 7 mm above the injector. Due to the pres-
toward Acid Orange dye or 2,4-dichlorophenol [21]. They also ence of the laser-absorbing ethylene molecules, the mixture is
report the presence of C OH, C O C, C O and COOH groups in rapidly heated, initiating the emergence of the oxidative pyrolysis
the carbonaceous layer and point out the existence of two differ- flames (presented in Fig. 1). The freshly resulted nanocompos-
ent sensitization processes: one induced by carbon and other by ite powders, carried by the gas flow, were then captured on a
the organic dye. The coating of pre-formed TiO2 nanoparticles with porous filter. The main experimental parameters are resumed in
disordered carbon was also successfully achieved using C2 H2 in a Table 1, including those for the synthesis of the reference sample
CVD process, resulting again an increasing visible-light photocat- without silica (TLOP 01). In order to decrease their surface-carbon
alytic activity against phenol [22]. One-step gas-phase synthesis content, the as-synthesized nanopowders were annealed during
of TiO2 @C nanostructured particles was reported using low oxi- 3 h in a furnace under synthetic air at 450 ◦ C. The raw powders
dant concentration from TTIP and C2 H2 in flame aerosol reactor morphology was examined with a Tecnai F30 G2 (300 kV) Trans-
[23] or from TiCl4 and C2 H4 with/without C2 H2 by laser pyrolysis mission Electron Microscope (TEM) and the elemental composition
[24]. Recently, carbon-incorporated TiO2 microspheres prepared by evaluation (for all (un)treated samples) was performed by energy-
flame-assisted hydrolysis of TBOT proved enhanced H2 production dispersive X-ray spectroscopy (EDS) using a liquid nitrogen-cooled
from water (when compared with TiO2 P25) during simulated solar SiLi detector from EDAX Inc. mounted inside a FEI Quanta Inspect
light irradiation [25]. S Scanning Electron Microscope (SEM) working up to 30 kV accel-
More complex than the TiO2 –SiO2 and TiO2 –C materials, the eration voltage. Their crystalline phase composition was revealed
ternary TiO2 /SiO2 /C composites can also be obtained, potentially using a PANalytical X’Pert MPD theta-theta X-ray diffraction (XRD)
benefiting from both specific properties of carbon and silica added apparatus using a Cu K␣ source (0.15418 nm). The weight fractions
to those of titania and possible enhanced by synergistic effect. (expressed in percents) of anatase (YA ) and rutile (YB = 100 − YA )
One such material was the carbonized mesoporous silica gel/CVD- phases in the raw and annealed nanopowders were calculated
titania composites with different C and TiO2 concentrations [26]. from the relative intensity of (1 0 1) anatase at 2 = 25.3◦ (IA )
The mechanochemical intercalation technique was used to obtain and (1 1 0) rutile at 2 = 27.5◦ (IB ) diffraction peaks according to
nanoporous graphite-derived carbon/TiO2 –SiO2 composites from Spurr and Myers formula YA = {1/[1 + 1.26 (IR /IA )]} × 100 [33]. A
C16 H33 N(CH3 )3 Br (CTAB) pre-intercalated graphite oxide, TEOS and Shimadzu 8400S Fourier Transform Infrared (FT-IR) Spectroscope
TTIP which was successfully tested for phenol adsorption and was used for the analysis of the different chemical bonds of the
degradation under UV sterilizing light [27]. The TiO2 supported nanocomposites. UV–vis diffuse reflectance spectroscopy (DRS)
by silica gel composite capacity to photodegrade the gray water spectra of calcined samples were recorded with a JASCO V-530
(rich in organic carbon) was increased through the addition of acti- UV–vis Spectrophotometer in the wavelength range 200–800 nm
vated carbon which enhances the composite’s absorption capacity for the determination of the absorption edge and bandgap energy
[28]. Another recent paper reported an enhancing of the visi- of the post-synthesis heat-treated powders using Kubelka–Munk
ble photocatalytic activity of TiO2 –SiO2 solvothermal-synthesized function or remission, F(R) and the Tauc plot.
228 C.T. Fleaca et al. / Applied Surface Science 336 (2015) 226–233

Fig. 1. Photographies of the reactive flames from TLOP 6 (left), TLOP 7 (center) and TLOP 8 (right) experiments.

Table 1
Main experimental parameters for the raw titania-based nanopowders laser pyrolysis syntheses.

Experiment Gas flows through nozzles [sccm] having each a specific cross-section area ˚

Inner (cylindrical) ˚1 = 4.9 mm2 Middle (annular) ˚2 = 46 mm2 Outer (annular) ˚3 = 44.5 mm2

Air > TiCl4 C2 H4 > TiCl4 Air C2 H4 > [(CH3 )3 Si]2 O Ar Ar

TLOP 01 200 60 – – 1700 –


TLOP 6 200 60 200 20 – 1700
TLOP 7 200 60 300 20 – 1700
TLOP 8 200 60 400 20 – 1700

The inner air flow (noted Air > TiCl4 ) was mixed with the inner ethylene flow (C2 H4 > TiCl4 ) and then bubbled through the liquid TiCl4 . The middle ethylene flow
(C2 H4 > [(CH3 )3 Si]2 O) was bubbled through HMDSO liquid and then mixed with supplementary air flow (Air). Flushing ZnSe windows flows: 2 × 850 sccm Ar; resulted TiCl4
vapor flow carried by air–ethylene mixture: 7.2 sccm; resulted [(CH3 )3 Si]2 O vapor flow carried by ethylene: 2 sccm; laser power density above the injector: 1780 W/cm2 ;
working pressure: 550 mbar.

3. Results and discussions their formation, one must take into account the composition of
the reaction mixture and the particularities of the laser pyrolysis
As can be seen in the TEM an HR-TEM images (Figs. 2–5), all technique. After absorbing the infrared laser photons, the vibra-
the resulted raw crystalline nanoparticles (including those from tionally excited ethylene molecules undergo collisions with the
the TLOP 01 reference sample without silica) are surrounded by surrounded species, transforming thus the laser energy into a
a more or less thick disordered shell/coating. In order to explain thermal one of the mixture molecules/radicals. In the same time,

Fig. 2. TEM and inserted HR-TEM images from TLOP 01 nanopowder. Fig. 3. TEM and inserted HR-TEM images from TLOP 6 nanopowder.
C.T. Fleaca et al. / Applied Surface Science 336 (2015) 226–233 229

air, respectively, for their complete oxidation (600 or 800 sccm for
incomplete oxidation to CO and H2 O). The employed air flows were
lower than those required for a stoichiometric ethylene oxidation
to CO2 and even to CO: 200 sccm (for TLOP 01), 400 sccm (for TLOP
6), 500 sccm (for TLOP 7) and 600 sccm (for TLOP 8) as can be
extracted from Table 1. It seems that the oxygen affinity of TiCl4
is higher than those of C2 H4 , as proved by the presence of TiO2
phases and the absence of metallic Ti, TiC or Ti suboxides in our raw
nanopowders – as can be observed in the X-ray diffractograms from
Fig. 6, where the phase identification was based on the reference
X-ray diffractograms edited by International Centre for Diffraction
Data (former Joint Committee on Powder Diffraction Standards):
JCPDS card 21-1272 for anatase and JCPDS card 21-1276 for rutile.
The elemental composition (extracted for the mediated EDX
analyses) presented in Table 2 proves the presence of carbon in
all samples and points out the ethylene as their main source. The
formation of carbonaceous material as a result of a partial ethylene
decomposition is an indicator for the oxygen deficiency in the
reactive mixtures used in our laser pyrolysis experiments and was
reported by us in [24,34] for nanotitania-based particle synthesis
using the same method, Ti precursor and sensitizer. Following all
these considerations, the titania-based core–shell nanostuctures
genesis can be explained starting from the hot freshly formed
titania nanoparticles in the flame central zone which act as cat-
alytic surfaces for the deposition of carbonaceous layer provided
Fig. 4. TEM, HR-TEM (right inset) and the electron diffraction pattern including the by ethylene decomposition, combined (for the TLOP 6,7 and 8
Miller indices and the measured interplanar distances (in nm) for anatase (A) and
experiments) with the amorphous silica from HMDSO (introduced
rutile (R) phases (left inset) images from TLOP 7 nanopowder.
through the annular nozzle) oxidation. As expected, with the
increasing amount of oxidant (O2 from air) the atomic carbon
due to the presence of oxygen, competing exothermal oxidative percent in the resulted silica-containing titania powders decreases
reactions occurs in the pyrolysis flame. Thus, TiCl4 , HMDSO and from ∼25% to ∼15% due to the conversion to CO2 (and/or CO) of
C2 H4 species compete for the available amount of oxidant, result- a larger percent of C2 H4 molecules. Moreover, when comparing
ing the corresponding solid (TiO2 , SiO2 ) and gaseous (CO2 ) oxides. the flame photographies from these three experiments (Fig. 1), a
In our experiments, the reactive mixtures were oxygen-deficient, clear change in their morphology for the experiment involving the
as can be calculated from the amount of precursors. The total highest air flow (TLOP 8) can be observed. The flame was brighter
oxidation of one C2 H4 mole (to form CO2 and H2 O) requires three and extended not only upwards but also downwards to the injector
moles of O2 , thus each sccm of C2 H4 needs 15 sccm of air. The total edge, whereas a light-green conical component appears in the
C2 H4 introduced flow was 60 sccm (for TLOP 01 experiment) or annular zone. Apparently, it seems that in this case the burning
80 sccm (for TLOP 6, 7, and 8), which requires 900 or 1200 sccm of process starts before the intersection with the infrared laser beam,

Fig. 6. Superposed X-ray diffractograms for the raw and post-synthesis annealed
Fig. 5. TEM and inserted HR-TEM images from TLOP 8 nanopowder. nanopowders.
230 C.T. Fleaca et al. / Applied Surface Science 336 (2015) 226–233

somewhat similar with the case presented in [35], where a N2

YR %rutile/[%anatase + %rutile]
shielded and air coflow assisted H2 /O2 diffusion flame was used
also for titania nanoparticles synthesis from TiCl4 . However, in
that experiment, the CO2 laser beam irradiated a pre-existing
flame and, because of the absence the laser-absorbing molecules,

Mediated elemental composition extracted from EDX measurements, mean XRD anatase/rutile crystallite sizes and XRD derived anatase/rutile weight fractions from raw and annealed TiO2 -based nanopowders.
the only absorbing species were the freshly formed nanoparticles

Raw/calcined
which undergoes advanced coalescence, melting and even partial

46.45/45.74
17.70/16.99
60.64/60.35

22.00/21.80
evaporation. In our case, without the laser beam, the flame cannot
be ignited, requiring a certain laser power density threshold in
order to be enough to heat the ethylene-containing mixture and
to ignite the flame. In the TLOP 8 case, where the largest amount
%anatase/[%anatase + %rutile]

of available oxidant was used, it seems that the flame not only
propagates upstream, but also expanded downstream, thus part
of ethylene molecules from both inner and intermediate flow
(including also the TiCl4 and HMDSO vapors) undergo a similar
oxidation with those encountered in a normal combustion flame.
Raw/calcined

39.36/39.65
53.55/54.26
83.30/83.01

78.00/78.20

In the typical laser-induced suspended flame experiments, the


reaction zone could not propagate downstream due to the lower
amount of oxygen available and as consequence, even if the total
YA

gas velocity was lower than in the last experiment, the advancing
of the reaction front was possible only in the direction of the
crystalline size [nm]

reactive flows (one can observe from Fig. 1 that all the presented
Rutile XRD mean

flames expand upwards, inside the cylindrical collector, well above


Raw/calcined

the laser beam). In the last experiment, in spite of the higher


central flow velocity, the presence of supplementary oxygen
(even nitrogen-diluted) induces a more advanced oxidation of
17/15
15/14

24/23
19/20

combustible species with the releasing of more heat, which is


translated in the faster expansion of the reaction front in all direc-
tions, including also downstream (countercurrent propagation
crystalline size [nm]
Anatase XRD mean

flame). Here is possible that the system had passed a threshold


when the reaction front propagation velocity surpasses those of
Raw/calcined

the reactive mixture (which depends on the gas/vapors fixed mass


flows and on the nozzles cross-sections). The reaction front stops
22/21
17/17
18/19
18/18

near the end of the injector, where the gas confined in the tubular
and annular spaces suddenly expands in the reaction chamber, sig-
24.7 ± 1.7/12.4 ± 1.5
19.4 ± 1.7/10.2 ± 1.4

15.1 ± 0.7/12.3 ± 1.2

nificantly losing their laminarity (especially the annular flow which


19.9 ± 0.7/11 ± 0.4

clearly expands toward the inert argon external flow) and partially
Raw/calcined

mixing. The flames could not propagate further downstream due


to the higher gas velocities inside the confined concentric channels
C [at.%]

of the glass injector. Moreover, this change can be related not


only with the higher oxygen concentration but also with a higher
gas velocity through the intermediate annular nozzle from the
54.9 ± 0.3/63.4 ± 2.6
59.1 ± 2.6/9.6 ± 0.9

TLOP 8 experiment toward those from TLOP 6 and TLOP 7 runs.


54.1 ± 0.5/61.9 ± 2
55.5 ± 1.7/59 ± 1

Another observation come when comparing the 1.8 value of the


Raw/calcined

Ti/Si atomic ratio in the introduced volatile precursors (TiCl4 and


HMDSO) flows with those in the resulted powders extracted from
O [at.%]

the same Table 2 (∼2.4 for TLOP 6, ∼1.9 for TLOP 7 and ∼2 for TLOP
8). Thus, one can conclude that a part of siloxane precursor was not
converted to solid form and therefore the resulted powders have a
higher titanium (and a lower silicon) content than the precursors
9.8 ± 0.8/10.1 ± 0.9

mixture. This observation can be explained if we compare the rates


8.9 ± 0.8/9.3 ± 1
6 ± 0.4/7.7 ± 0.8
Raw/calcined

of our titania and silica precursors oxidative reactions at flame


temperatures around 1000 K (or less) presented in Fig. 9 from the
Si [at.%]

reference [13]. In these conditions the TiCl4 reacts much faster than
0/0

HMDSO, thus part of the Si-based precursor was not converted to


silica due to the insufficient residence time in the reactive flame.
Concerning the TLOP 01 reference sample (without silica) mor-
14.2 ± 1.7/16.5 ± 3.3
16.9 ± 1/17.7 ± 1.4
21 ± 4.2/30.1 ± 0.4

phology, the TEM image from Fig. 2 reveals the presence of round
19.2 ± 2/18.7 ± 1

particles, having sizes between 20 and 40 nm and surrounded with


Raw/calcined

a conformal few nm thick shell. The HR-TEM inset image from


Ti [at.%]

the same figure shows a group of crystalline nanoparticles sur-


rounded by turbostratic/disordered graphitic sheets. The 3.52 Å
interplanar distance measured on the top particle can be attributed
TLOP 01
Sample

to (1 0 1) anatase phase, found also in the corresponding X-ray


TLOP 6
TLOP 7
TLOP 8
Table 2

diffractgram (Fig. 6). In Fig. 3 TEM image from the raw TLOP 6
sample, the round (and few facetted) nanoparticles are embedded
C.T. Fleaca et al. / Applied Surface Science 336 (2015) 226–233 231

in/surrounded by a thicker layer formed due the supplementary


introduction of HMDSO silica precursor in the reaction zone. In
this image, a bimodal particle distribution can be observed with
small nanoparticles having sizes around 10–12 nm, whereas for
the bigger ones diameters between 15 and 45 nm can be mea-
sured. The inset from the same image shows a well-crystallized
particle (25 nm in size) coated with rather disordered carbona-
ceous/amorphous silica layer, where the 2.18 Å interplanar distance
specific to (1 1 1) planes of the rutile phase can be identified. This
phase is also the majority one in this sample (60.64%) according to
Table 2 data extracted from XRD. For the next TLOP 7 sample, the
TEM image from Fig. 4 reveals the presence of aggregates composed
from round, ellipsoidal or partially facetted nanoparticles (with
sizes from 8 up to 54 nm, typical around 25–30 nm) also coated
with the more diffuse layer. A more detailed structure can be visu-
alized in the right down corner HR-TEM inset from the same image,
where the (1 0 1) anatase planes can be identified, delimited by tur-
bostratic graphite units composed from short packed graphenes
(spaced at ∼3.6–4 Å) mixed with more disordered zones (possibly
containing amorphous silica). The presence of turbostratic carbon
on the surface of laser-assisted synthesis of titania nanocrystals was Fig. 7. Superposed FT-IR spectra of the raw titania-based nanopowders.
also reported by us in [24,32]. In the same Fig. 4, the left SAED image
on the inset (presenting rings and spots due to the diffraction of the
electrons on multiple nanocrystals) attests the simultaneous pres- supplementary silica formation. There are also many peaks that
ence of anatase and rutile titania phases in this sample, confirming are common to all samples. The large peak at 3390–3420 cm−1 and
thus the X-rays diffraction analysis (see Fig. 6 and Table 2). The the small peak at 1632 cm−1 can be also found in [4,38,39]. They
nanoparticles resulted from the last experiment (TLOP 8) – where were attributed to the stretching modes of O H bonds from the sur-
the greatest amount of air (and oxygen) was used – have sizes face adsorbed water and to the bending vibrations of physisorbed
between 8 and 25 nm and a more facetted, polyhedral morphol- and/or chemosorbed water, respectively. The weak band around
ogy, being covered with a very thin disordered layer and as can be 1410–1420 cm−1 found also in all samples was identified at
seen in the TEM and HR-TEM images from Fig. 5. In this case, due to 1400 cm−1 in a TiO2 hydrosol and attributed to Ti O Ti vibra-
the different flame morphology (discussed above), associated with tions [5]. The intense peak centered at 1095 cm−1 with his shoulder
the separate introducing of the two precursors we suspect that a at 1227 cm−1 , found only for silica-containing powders was also
part of siloxane precursor was oxidated independently from the reported at 1065 and 1200 cm−1 in [1] and attributed to Si O Si
titania particles surfaces, and a mixture of titania-rich and silica- vibrations. The small peak visible in the same powders spectra
rich particles was obtained. Thus, the particles from Fig. 5 seem centered around 940 cm−1 was considered an indication of the
to below to the titania-rich category. In the same time, having the chemical connection between titania and silica, being reported and
lowest carbon loading from all experiments, it is expected that the ascribed to Ti O Si vibrations [2,4,38]. Also, at 955 cm−1 the pres-
TLOP 8 nanoparticles to have the thinnest carbonaceous shell. The ence of Si OH stretching vibrations corresponding band was also
facetted nanocrystal form the HR-TEM inserted image (see Fig. 5) reported in [1], whereas the same Si OH attributed peak was found
can be attributed to the rutile phase due to the 3.21 Å identified at 921 cm−1 from SiOx films prepared by plasma CVD [40]. The
interplanar distance value of their (1 1 0) planes (a value close to (Si C) stretching vibration infrared peak observed at 846 cm−1 for
3.247 Å from XRD standard), which in this case is the minority phase Si/C/H materials obtained by CO2 laser decomposition of Si(C2 H3 )4
(22%) as resulted from Table 2. Generally, the mean crystallite size [41] or the peak at 802 cm−1 attributed to (Si C) + (CH3 Si)
values of each phase extracted with the aid of Scherrer equation vibrations (seen in a film obtained from [(CH3 )2 HSi]2 O laser decom-
from the anatase (1 0 1) and rutile (1 1 0) XRD peaks (Table 2) can position) [42] are missing from our silicon-containing samples
be correlated with those from the presented TEM images, where spectra, indicating the oxidation of HMDSO molecules in the pres-
much smaller nanoparticle populations were considered. A clear ence of oxygen from the injected air in our experiments. The large
tendency of the rutile to anatase ratio (as well as the carbon con- peak around 690–700 cm−1 was fond also in flame-made titania
tent) to decrease with the increasing amount of injected air through FTIR spectrum and attributed to Ti O Ti bonding [9]. Another
the second annular nozzle can be observed and was also reported intense peak, found also for all samples around 460–510 cm−1 and
by us in [24]. This effect can be related to the role of carbon in may be a signature from Ti O bonds, reported also for titania thin
the enhancing the transformation of anatase to rutile through the films [43] at 430 cm−1 or nanotubes [44] at 496 cm−1 . In the same
formation of oxygen vacancies [36], correlated with the high tem- article [44], for titania–carbon nanocomposite, a strong IR line at
peratures in the flame environment. The mean crystallite size and 488 cm−1 was attributed to Ti C bond. Finally, for TLOP 8 sample,
anatase to rutile ratios of the 450 ◦ C annealed powders (with lower another peak at 460 cm−1 can be identified, close to the Si O Si
carbon content) are close to those of the raw powders as resulted bending mode peak at 470 cm−1 observed in the laser-synthesized
from Fig. 6 and summarized in Table 2. These results are expected oxygen-deficient titania–silica powders [30].
since for the pure titania nanopowders (synthesized by laser pyrol- The UV–vis absorbance superposed spectra of the annealed
ysis from alkoxides) the anatase to rutile transformation becomes powders (with lower carbon content) obtained from diffuse
significant only over 550 ◦ C [37]. It is worth nothing that no other reflectance spectroscopy (DRS) are presented in Fig. 8a. Their ultra-
peaks than anatase and rutile were found in our raw or annealed violet absorption peaks are positioned around 330 nm, a value close
powders X-ray diffractograms, indicating thus the disordered or to those measured for TiO2 @C laser-synthesized nanocomposites
amorphous state of the carbon and/silica components. [24]. A clear enhanced visible light absorption can be observed for
The superposed infrared spectra of the raw powders (Fig. 7) the annealed silica-containing nanopowders (especially for TLOP
show significant changes after the HMDSO introduction due to the 6(H) and TLOP 7(H) particles), when compared with the calcined
232 C.T. Fleaca et al. / Applied Surface Science 336 (2015) 226–233

Fig. 8. UV–vis absorbance spectra of annealed samples (a) and the corresponding Tauc plot data (b) for indirect electronic transition [F(R)h]1/2 vs. E (eV) for the same
annealed powders.

reference sample TLOP 01(H), even if their atomic carbon concen- the fixed introduced O2 flow together with the fuel and the oxidic
tration are quite similar (between 10.2% and 12.3% as can be seen precursor. Moreover, the laser power density in the flame can be
in Table 2). Moreover, unlike the carbon-modified TiO2 (n-TiO2 ) varied with the possibility to control the size, morphology, agglom-
prepared from Ti sheet in furnace or in flame whose absorbance eration degree and crystalline phase as was presented for the case
goes to zero at 430 and 570 nm, respectively [45], our annealed of pure TiO2 or SiO2 nanoparticles in [35]. A better degree of the
titania–carbon–silica samples do not show an absorbance satura- control of carbon percent and their morphology in these nanocom-
tion even at 800 nm. The optical bandgap values of the annealed posites – without the need of the final air annealing step – can
powders were extracted from the Tauc plot of the Kubelka–Munk also be achieved with our method by using higher oxygen to nitro-
functions [F(R)h]1/2 versus energy (h) measured in electron- gen (or argon) ratio in the reactive flow, independent introduction
volts, according to Tandom and Gupta’s method [46] and presented of TiCl4 vapors and ethylene sensitizer/carbon donor flows and/or
in Fig. 8b. Due to the fact that pure SiO2 phases do not exhibit by tuning the gas velocities through the injector nozzles. Besides
absorption in the 200–500 nm domain [38], the absorption prop- the fast continuous production of nanopowders, another advantage
erties observed for our annealed powders are only due to titania of the laser oxidative pyrolysis method is the avoiding the large
phases and carbon. Yet, the amorphous silica on/near the titania amount of solvents usually needed by the sol–gel methods. The
surface interacting with carbonaceous layer can influence the pho- scalability of the laser pyrolysis method to pilot plant production
tocatalytic properties for example by enhancing the hydrophylicity of TiO2 was demonstrated [52], being thus an important advantage
and the organic species adsorption. The pure (non-doped) anatase of this technique.
and rutile phases exhibit an indirect bandgap of 3.2–3.3 eV and
3.0–3.1 eV, respectively [47]. Our annealed samples with the high- 4. Conclusions
est anatase content: TLOP 01(H) – 83.3% and TLOP 8(H) – 78%
show a lower bandgap (3.04 eV and 3.03 eV, respectively) than Carbon and silica disordered/amorphous layers coated TiO2
pure anatase. Lower bandgap values were obtained for the sam- nanopowders containing a mixture of anatase and rutile crystalline
ples with lower anatase content TLOP 6(H) – 39.65% and TLOP 7(H) phases were successfully synthesized using a continuous oxidative
– 54.26% (2.94 eV and 2.79 eV, respectively). For these samples, the laser pyrolysis method. We employed TiCl4 and HMDSO vapors as
corresponding absorption edge is 408, 409, 422 and 444 nm, respec- precursors and C2 H4 as sensitizer/carbon precursor in the presence
tively, which is an indication of possible photocatalytic activity in of substoechiometric (relative to the ethylene oxidation) amount
the visible region. These results confirm our precedent studies on of air. Infrared spectra of nanopowders confirm the presence of
laser-synthesized TiO2 @C doped [32] or undoped [24] with sul- Ti O, Si O and Ti O Si bonds, whereas EDX estimations show a
fur, when also lower bangap values were reported. There are even Ti to Si atomic ratio between 1.9 and 2.4. Their carbon content was
lower bandgap values reported for some interstitial and superficial controlled using different air flows and subsequent furnace calcina-
carbon doped/covered anatase obtained by sol–gel from TBOT [48] tion down to ∼11–12 at.%. For the raw powders, a carbon content
which showed visible-light activity in methylorange degradation. decreasing accompanied by the anatase to rutile ratio increasing
Recently, using a solvothermal method, other authors also obtained with the increasing air flow was observed. Using the highest air
carbon-doped TiO2 with low bandgap from TBOT (2.84 eV) – con- flow, the reaction zone change their morphology by extending
taining only anatase – or from TiCl3 (2.39 eV) – where anatase, rutile both upward and downward and the apparition of a second conical
and brookite phases coexist –, both presenting in the same time flame, possibly due to exceeding of a threshold when the burning
intense photoactivity under visible light [49]. It must be noted that velocity becomes higher than those of the central reactive flow.
the reference commercial photocatalyst Degussa P25 pure nano- UV–vis spectra of the annealing samples reflect the influence of
metric titania contains a significant rutile percent (∼25%) [50] and the anatase/rutile mixture and carbon content, whereas their lower
have a higher bandgap energy (3.2 eV) [32]. Moreover, another bandgap (when compared with those of Degussa P25 commercial
nanometric titania photocatalyst having rutile as majority phase nanopowder) values suggest their possible photocatalytic activity
(∼53%) was able to efficiently photodegrade chlorinated phenol under visible light against organic substances, a direction that will
[51]. be explored in the future experiments.
In contrast to our oxidative laser pyrolysis method, the carbon
incorporation control is more difficult to be achieved in the alter- Acknowledgements
native classic atmospheric pressure open flame synthesis of such
materials due to the uncontrolled involvement of the surrounding The financial support is gratefully acknowledged to Roma-
oxygen from air which acts as a supplementary oxidant besides nian Ministry of Education and Research under the Project
C.T. Fleaca et al. / Applied Surface Science 336 (2015) 226–233 233

PNCD2 IDEI no. 80/2011 and to Romanian Ministry of European [26] R. Leboda, M. Marciniak, V.M. Gun’ko, W. Grzegoroczyk, A.A. Malygin, A.A.
Funds and Sectoral Operational Program for Human Resources Malkov, Structure of carbonized mesoporous silica gel/CVD-titania, Colloids
Surf. A 167 (2000) 275–285.
Development 2007–2013 through the Financial Agreement POS- [27] Y.-H. Chu, M. Yamagishi, Z.-M. Wang, H. Kanoh, T. Hirotsu, Synthe-
DRU/159/1.5/S/132395 (INNO Research). sis of nanoporous graphite-derived carbon/TiO2 –SiO2 composites by a
mechanochemical intercalation method, Microporous Mesoporous Mater. 118
(2009) 496–502.
References [28] C. Ludwig, H. Byrne, J. Stokke, P. Chadik, D. Mazyck, Performance of silica–titania
carbon composites for photocatalytic degradation of gray water, J. Environ. Eng.
[1] F.J. Heiligtag, N. Kranzlin, M.J. Suess, M. Niederberger, Anatase–silica compos- 137 (2011) 38–45.
ite aerogel: a nanoparticle-based approach, J. Sol–Gel. Sci. Technol. 70 (2014) [29] Q. Chen, H. Shi, W. Shi, Y. Xu, D. Wu, Enhanced visible photocatalytic activity
300–306. of titania–silica photocatalysts: effect of carbon and silver doping, Catal. Sci.
[2] D. Kannaiyan, S.T. Kochuveedu, Y.H. Jang, J.Y. Lee, J. Lee, J. Kim, D.H. Technol. 2 (2012) 1213–1220.
Kim, Enhanced photophysical properties of nanopatterned titania nan- [30] H. Maskrot, N. Herlin-Boime, Y. Leconte, K. Jursikova, C. Reynaud, J. Vicens, Blue
odots/nanowires upon hybridization with silica via block copolymer templated TiO2−x /SiO2 nanoparticles by laser pyrolysis, J. Nanopart. Res. 8 (2006) 351–360.
sol–gel process, Polymers 2 (2010) 490–504. [31] C.T. Fleaca, M. Scarisoreanu, I. Morjan, R. Alexandrescu, F. Dumitrache, C.
[3] T.-V. Nguyen, S. Kim, O.-B. Yang, Water decomposition on TiO2 –SiO2 and Luculescu, I.P. Morjan, R. Birjega, A.-M. Niculescu, G. Filoti, V. Kuncser, E. Vasile,
RuS2 /TiO2 –SiO2 photocatalysts: the effect of electronic characteristics, Catal. V. Danciu, M. Popa, Recent progress in the synthesis of magnetic titania/iron-
Commun. 5 (2004) 59–62. based composite, nanoparticles manufactured by laser pyrolysis, Appl. Surf. Sci.
[4] A. Eshaghi, A. Eshaghi, Investigation of superhydrophilic mechanism of titania 302 (2014) 198–204.
nanolayer thin film – silica and indium oxide dopant effect, Bull. Mater. Sci. 35 [32] M. Scarisoreanu, I. Morjan, R. Alexandrescu, C.T. Fleaca, A. Badoi, E. Dutu, A.-
(2012) 137–142. M. Niculescu, C. Luculescu, E. Vasile, J. Wang, S. Bouhadoun, N. Herlin-Boime,
[5] C. Huang, H. Bai, Y. Huang, S. Liu, S. Yen, Y. Tseng, Synthesis of neutral TiO2 –SiO2 Enhancing the visible light absorption of titania nanoparticles by S and C doping
hydrosol and its applications as antireflective self-cleaning thin films, Int. J. in a single-step process, Appl. Surf. Sci. 302 (2014) 11–18.
Photoenergy 2012 (2012) 8, http://dx.doi.org/10.1155/2012/620764. [33] R.A. Spurt, H. Myers, Quantitative analysis of anatase–rutile mixtures with an
[6] D.A. Kumar, J.M. Shyla, F.P. Xavier, Synthesis and characterization of X-ray diffractometer, Anal. Chem. 29 (1957) 760–762.
TiO2 /SiO2 nano composites for solar cell applications, Appl. Nanosci. 2 (2012) [34] R. Alexandrescu, F. Dumitrache, I. Morjan, I. Sandu, M. Savoiu, I. Voicu, C. Fleaca,
429–436. R. Piticescu, TiO2 nanosized powders by TiCl4 pyrolysis, Nanotechnology 15
[7] H. Wang, P. Xu, S. Meng, W. Zhong, W. Du, Q. Du, Poly(methylmethacrylate)/ (2004) 537–545.
silica/titania ternary nanocomposites with greatly improved thermal [35] D. Lee, M. Choi, Coalescence enhanced synthesis of nanoparticles to control
and ultraviolet-shielding properties, Polym. Degrad. Stabil. 91 (2006) size, morphology and crystalline phase at high concentrations, J. Aerosol Sci.
1455–1461. 33 (2002) 1–16.
[8] J.P. Lange, L.S. Patil, D.K. Gautam, Influence of titanium-tetra-isopropoxide flow [36] D.A.H. Hanaor, C.C. Sorell, A review of anatase to rutile phase transformation,
on TiO2 doped SiO2 films for waveguide applications, J. Optoelectron. Biomed. J. Mater. Sci. 46 (2011) 855–874.
Appl. 1 (2009) 319–324. [37] L.E. Depero, P. Bonzi, M. Zocchi, C. Casale, G. De Michele, Study of the
[9] S. Sheen, S. Yang, K. Jun, M. Choi, One-step method for the synthesis of coated anatase–rutile transformation in TiO2 powders obtained from laser-induced
composite nanoparticles, J. Nanopart. Res. 11 (2009) 1767–1775. synthesis, J. Mater. Res. 8 (1993) 2709–2715.
[10] S.H. Jeon, T.G. Lee, Preparation of column-shaped TiO2 –SiO2 nanoparticles by [38] B. Llano, M.C. Hidalgo, L.A. Rios, J.A. Navio, Effect of the type of acid used in
a diffusion flame reactor and their photocatalytic properties, Appl. Chem. 10 the synthesis of titania–silica mixed oxides on their photocatalytic properties,
(2006) 364–368. Appl. Catal. B 151–152 (2014) 389–395.
[11] K.A. Akurati, R. Ditman, A. Vital, U. Klotz, P. Hug, T. Graule, M. Winterer, Silica- [39] X.F. Hou, H. Ding, Y.X. Zheng, M.M. Wang, Preparation and characterization
based composite and mixed-oxide nanoparticles from atmospheric pressure of amorphous silica/anatase composite through mechanochemical method,
flame synthesis, J. Nanopart. Res. 8 (2006) 379–393. Mater. Res. Innov. 17 (2013), S1-234–S1-238.
[12] S.H. Ehrman, S.K. Friedlander, M.C. Zachariach, Phase segregation in binary [40] Y. Song, T. Sakurai, K. Kishimoto, K. Maruto, S. Matsumoto, K. Kikuchi, Syntheses
SiO2 /TiO2 and SiO2 /Fe2 O3 nanoparticle aerosol formed in premixed flame, J. and optical properties of low-temperature SiOx and TiOx thin films prepared
Mater. Res. 14 (1999) 4551–4561. by plasma enhanced CVD, Vacuum 51 (1998) 525–530.
[13] S.H. Ehrman, S.K. Friedlander, M.C. Zachariach, Characteristics of SiO2 /TiO2 [41] V. Drinek, Z. Bastl, J. Subrt, R. Taylor, J. Pola, IR laser-induced decomposition of
nano-ascomposite particles formed in a premixed flat flame, J. Aerosol Sci. 29 tetravinylsilane for chemical vapor deposition of Si/C/H materials, J. Anal. Appl.
(1998) 687–706. Pyrol. 35 (1995) 199–206.
[14] M. Inagaki, Y. Hirose, T. Matsunaga, T. Tsumura, M. Toyoda, Carbon coating of [42] J. Pola, D. Pokorna, Z. Bastl, J. Subrt, IR laser-induced chemical vapour deposition
anatase-type TiO2 through precipitation in PVA aqueosu solution, Carbon 41 of silicon oxicarbide phases from 1,1,3,3-tetramethyldisiloxane, J. Anal. Appl.
(2003) 2619–2624. Pyrol. 38 (1996) 153–159.
[15] M. Inagaki, F. Kojin, B. Tryba, M. Toyoda, Carbon-coated anatase: the [43] L.-L. Yang, Y.-S. Li, J.S. Chen, P.H. Tsai, C.L. Chen, C.J. Chang, Compositional
role of the carbon layer for photocatalytic performance, Carbon 43 (2005) tailored sol–gel SiO2 –TiO2 thin films: crystallization, chemical bonding con-
1652–1659. figuration, and optical properties, J. Mater. Res. 20 (2005) 3141–3149.
[16] H. Wang, X. Quan, H. Yu, S. Chen, Fabrication of TiO2 /carbon nanowall hetero- [44] A. Anson-Casaos, I. Tacchini, A. Unzue, M.T. Martinez, Combined modification of
junction and its photocatalytic ability, Carbon 46 (2008) 1126–1132. a TiO2 photocatalyst with two different carbon form, Appl. Surf. Sci. 270 (2013)
[17] X. Zhang, M. Zhou, L. Lei, Preparation of photocatalytic TiO2 coatings of 675–684.
nanosized particles on activated carbon by AP-MOCVD, Carbon 43 (2005) [45] S.U.M. Khan, M. Al-Shahry, W.B. Ingler Jr., Efficient photochemical water split-
1700–1708. ting by a chemically modified n-TiO2 , Science 297 (2002) 2243–2244.
[18] C.X. Dong, A.P. Xian, E.H. Han, J.K. Shang, C-doped TiO2 with visible light pho- [46] S.P. Tandom, J.P. Gupta, Measurement of forbidden energy gap of semi-
tocatalytic activity, Solid State Phenom. 121–123 (2007) 939–942. conductors by Diffuse Reflectance Technique, Phys. Status Solidi B 38 (1970)
[19] L.-W. Zhang, H.-B. Fu, Y.-F. Zhu, Efficient TiO2 photocatalyst from surface 360–367.
hybridization of TiO2 particles with graphite-like carbon, Adv. Funct. Mater. [47] A.L. Linsebigler, G. Lu, J.T. Yates, Photocatalysis on TiO2 surfaces: principles,
18 (2008) 2180–2189. mechanisms, and selected results, Chem. Rev. 95 (1995) 735–758.
[20] Y. Park, W. Kim, H. Park, T. Tachikawa, T. Majima, W. Choi, Carbon-doped TiO2 [48] C. Chen, M. Long, H. Zheng, W. Cai, B. Zhou, J. Zhang, Y. Wu, D. Ding, D. Wu,
photocatalyst synthesized without using an external carbon precursor and the Preparation, characterization and visible-light activity of carbon modified TiO2
visivle licht activity, Appl. Catal. B 91 (2009) 355–361. with two kinds of carbonaceous species, J. Mol. Catal. A 314 (2009) 35–41.
[21] J. Zhong, F. Chen, J. Zhang, Carbon-deposited TiO2 : synthesis, characterization, [49] X. Wu, S. Yin, Q. Dong, C. Guo, H. Lui, T. Kimura, T. Sato, Synthesis of high visible
and visible photocatalytic performance, J. Phys. Chem. C 114 (2010) 933–939. light active carbon doped TiO2 photocatalyst by a facile calcination assisted
[22] M. Wojtoniszak, D. Dolat, A. Morawski, E. Mijowska, Carbon-modified TiO2 for solvothermal method, Appl. Catal. B 142–143 (2013) 450–457.
photocatalysis, Nanoscale Res. Lett. 7 (2012) 235–240. [50] T. Ohno, K. Sarukawa, K. Tokeida, M. Matsamura, Morphology of a TiO2 pho-
[23] K.H. Kammler, S.E. Pratsinis, Carbon-coated titania nanostructured particles: tocatalyst (Degussa P25) consisting of anatase and rutile phases, J. Catal. 203
continuous, one-step flame-synthesis, J. Mater. Res. 18 (2003) 2670–2676. (2001) 82–86.
[24] M. Scarisoreanu, R. Alexandrescu, I. Morjan, R. Birjega, C. Luculescu, E. Popovici, [51] S. Bakardjieva, J. Subrt, V. Stengl, M.J. Dianez, M.J. Sataguez, Photoactivity of
E. Dutu, E. Vasile, V. Danciu, N. Herlin-Boime, Structural evolution and optical anatase–rutile mixtures obtained by heat treatment of homogeneously pre-
properties of C-coated nanoparticles prepared by laser pyrolysis, Appl. Surf. Sci. cipitated anatase, Appl. Catal. B 58 (2005) 193–202.
278 (2013) 295–300. [52] B. Pignon, H. Maskrot, V. Guyot Ferreol, Y. Leconte, S. Coste, M. Gervais,
[25] X. Zhang, Y. Sun, X. Cui, Z. Jiang, Carbon-incorporated TiO2 microspheres: facile T. Pouget, C. Reynaud, J.-F. Tranchant, N. Herlin-Boime, Versatility of laser
flame assisted hydrolysis of tetrabutyl orthotitanate and photocatalytic hydro- pyrolysis applied to the synthesis of TiO2 nanoparticles – application to UV
gen production, Int. J. Hydrogen Energy 37 (2013) 1356–1365. attenuation, Eur. J. Inorg. Chem. 6 (2008) 883–889.

You might also like