You are on page 1of 7

Thin Solid Films 520 (2011) 1195–1201

Contents lists available at ScienceDirect

Thin Solid Films


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / t s f

Determination of photo-catalytic activity of un-doped and Mn-doped TiO2 anatase


powders on acetaldehyde under UV and visible light
Vassileios C. Papadimitriou a,⁎, Vassileios G. Stefanopoulos a, Manolis N. Romanias a,
Panos Papagiannakopoulos a,⁎, Kyriaki Sambani b, Valentin Tudose b, George Kiriakidis b
a
Laboratory of Photochemistry and Kinetics, Department of Chemistry, University of Crete, 71003 Heraklion, Crete, Greece
b
Institute of Electronic Structure and Laser, Foundation for Research & Technology-Hellas, PO Box 1527,Vasilika Vouton, 71110 Heraklion, Crete, Greece

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

Available online 5 August 2011 Titanium dioxide (TiO2) photocatalytic powder materials doped with various levels of manganese (Mn) were
synthesized to be used as additives to wall painting in combating indoor and outdoor air pollution. The
Keywords: heterogeneous photocatalytic degradation of gaseous acetaldehyde (CH3CHO) on Mn–TiO2 surfaces under
Photocatalysis ultraviolet and visible (UV/Vis) irradiation was investigated, by employing the Photochemical Static Reactor
Indoor pollution coupled with Fourier-Transformed Infrared spectroscopy (PSR/FTIR) technique. Experiments were performed
TiO2
by exposing acetaldehyde (~400 Pa) and synthetic air mixtures (~1.01 × 10 5 Pa total pressure) on un-doped
Mn doped TiO2
Acetaldehyde
TiO2 and doped with various levels of Mn (0.1–33% mole percentage) under UV and visible irradiation at room
temperature. Photoactivation was initiated using either UV or visible light sources with known emission
spectra. Initially, the photo-activity of CH3CHO under the above light sources, and the physical adsorption of
CH3CHO on Mn–TiO2 samples in the absence of light were determined prior to the photocatalytic
experiments. The photocatalytic loss of CH3CHO on un-doped TiO2 and Mn–TiO2 samples in the absence
and presence of UV or visible irradiation was measured over a long time period (≈ 60 min), to evaluate their
relative photocatalytic activity. The gaseous photocatalytic end products were also determined using
absorption FTIR spectroscopy. Carbon dioxide (CO2) was identified as the main photocatalysis product. It was
found that 0.1% Mn–TiO2 samples resulted in the highest photocatalytic loss of CH3CHO under visible
irradiation. This efficiency was drastically diminished at higher levels of Mn doping (1–33%). The CO2 yields
were the highest for 0.1% Mn–TiO2 samples under UV irradiation, in agreement with the observed highest
CH3CHO decomposition rates. It was demonstrated that low-level (0.1%) doping of TiO2 with Mn results in a
significant increase of their photocatalytic activity in the visible range, compared to un-doped TiO2. This
elevated activity is lost at high doping levels (1–33%). Finally, the photocatalytic degradation mechanism of
CH3CHO on 0.1% Mn–TiO2 surfaces under visible irradiation leading to low CO2 yields is different than that
under UV irradiation resulting to high CO2 yields.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction composition of the troposphere [12,13]. The photocatalytic activity of


TiO2 can be enhanced either by adding several transition metals (V, Cr,
The heterogeneous photocatalysis of organic and inorganic pol- Fe, Co, Ni, etc), nonmetals (B, N, C, S, etc), and amorphous carbon or as
lutants both in gas and liquid phase on TiO2 surfaces is a significant nanoparticles, providing stable and low cost materials that increase
physicochemical process with several environmental applications [1–4]. substantially the degradation yields of various chemical compounds
Recently, it has been used to remove from the atmosphere several types [14]. The increase in the photo-activity of TiO2 has been attributed to the
of air pollutants (NOx, SO2, volatile organic compounds etc) and toxic high surface area of the composite photocatalyst, the red-shift in their
contaminants upon UV–visible radiation at ambient temperature and absorption spectrum to the visible range [2], and thus the generation of a
pressure [5–11]. Furthermore, the heterogeneous chemistry of primary sufficient number of excited electrons in the conduction band. The
air pollutants on the surface of atmospheric dust particles (aerosols), above, combined with the recombination rate of the induced electron–
containing metal oxides, in the presence of sunlight is also an important hole pairs define the so called photocatalytic efficiency of the material.
and unknown atmospheric process that contributes to the chemical Acetaldehyde is a primary indoor and outdoor air pollutant, and
its removal from the atmosphere via heterogeneous photocatalytic
⁎ Corresponding authors. Fax: + 30 2810 545001.
activity on TiO2 surfaces presents considerable interest [15,16]. The
E-mail addresses: bpapadim@chemistry.uoc.gr (V.C. Papadimitriou), photocatalytic degradation of acetaldehyde on rutile TiO2(110) in the
panosp@chemistry.uoc.gr (P. Papagiannakopoulos). presence of both oxygen and UV light at ambient temperature has

0040-6090/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.tsf.2011.07.073
1196 V.C. Papadimitriou et al. / Thin Solid Films 520 (2011) 1195–1201

been studied recently by using Photon Stimulated Desorption and sources. The first light source was a 300 W Osram lamp (Ultra Vitalux)
Thermal Programmed Desorption techniques [15]. It has been pro- with discrete emission lines and maximum intensity at ~ 350 nm,
posed that acetaldehyde undergoes photooxidation via two consec- further denoted as UV light source. This lamp also emits in the visible
utive processes; initially acetaldehyde reacts thermally with adsorbed region but at much lower power. The second light source was a 500 W
oxygen and forms a photoactive acetaldehyde–oxygen complex, Vito halogen projector lamp (VT 364) with a broad emission spectrum
which subsequently undergoes photodecomposition producing gas (350–1050 nm) and maximum intensity at ~ 720 nm, further denoted
phase methyl radical and formate bound on the surface, identical to as Vis light source. The emission spectrum of the visible light source
that observed previously for acetone [17]. simulates satisfactorily the indoor light conditions. It is worth noting
The primary aim of the present study was to investigate the photo- that the tail of the emission spectrum extends to the UV region, but
catalytic activity of TiO2 samples enriched with several percentages of composites a negligibly small portion of total emission spectrum. Both
Mn for the degradation of gaseous acetaldehyde in the presence of lamps were commercially available.
both O2 and UV–visible light at atmospheric pressures and room
temperature. Initially, the photocatalytic degradation of acetaldehyde 2.2. Materials
by un-doped TiO2 samples was evaluated in the presence of either UV
or visible light under identical pressure and temperature conditions, in All chemicals used were commercially available. CH3CHO stated
order to assess the effect of different doping levels with Mn. Con- purity was 99.5% (FERAK), while synthetic air was 20.5% O2 in N2
sequently, the final volatile infrared active products were identified in (Messer). CH3CHO was degassed through several freeze-pump-thaw
the above experimental conditions and the overall photocatalytic cycles at 77 K prior to use. Un-doped TiO2 and doped with various mole
degradation mechanism of acetaldehyde by TiO2 samples is discussed. percentages of Mn powder materials were synthesized (Patent GR-
20090100724 [18]) and were characterized prior use. Synthetic
2. Experimental details precursors, oxysulfate hydrate (TiOSO4 × H2O), manganese (II) acetate
tetrahydrate, ≥99% and titanium (IV) tetraisopropoxide [Ti(OCH
The photocatalytic activity of un-doped TiO2 and doped with Mn (CH3)2)4] were purchased from Aldrich and used without further
samples was determined by employing the Photochemical Static processing. In brief, Mn doped TiO2 samples were prepared by a mo-
Reactor equipped with a Fourier Transformed Infrared spectrometer dified sol gel method. The photocatalyst was obtained by precipitating
(PSR/FTIR). The experimental setup consisted of four main units: titanium dioxide on a sol of manganese dioxide, as shown in the
(a) the gas supply and preparation glass vacuum line with several flowchart of the synthesis (Fig. 2) employing TiOSO4. The hydrated
storage bulbs, (b) the properly designed optical cell/reactor assorted manganese dioxide sol was obtained by mixing the required volumes
with two retractable NaCl windows transparent to IR region, in order of manganese acetate, i.e. Mn(CH3COO)2, 0.1 M and potassium per-
to insert and remove the catalyst glass-bed (Fig. 1), (c) a Fourier manganate KMnO4, 0.1 M solutions and stirring the mixture for 24 h at
Transformed Infrared spectrometer (FTIR, Jasco 6300) to monitor room temperature. The obtained sol was mixed with a solution of
reactant and products concentrations, and (d) two light sources (UV TiOSO4. The concentration of TiOSO4 in the final solution was 0.1 M.
and Visible) positioned 80 cm away from the photocatalytic samples. The colloidal solution was stirred (Fig. 2, step1) at room temperature
for 48 h, in order to obtain the adsorption equilibrium. During this
2.1. Light sources phase, an exchange of Mn with Ti occurs and the final sol is a mixture of
both dioxides. After this step, the remaining Ti4+ ions were forced to
To evaluate the ability of the various TiO2 samples enriched with precipitate by adding NH3 solution, so that the final pH was 7 (Fig. 2,
Mn to reduce the photo-activation gap and shift the absorption of step 2). Gel formed was stirred continuously at room temperature
light into the visible region, the present study used two different light for aging (48 h). After aging, the sol separated using centrifuge or

Fig. 1. Schematic of the pyrex glass Photochemical Static Reactor used in the photocatalytic experiments. Retractable windows, catalyst's glass-bed and pressure gage are shown.
V.C. Papadimitriou et al. / Thin Solid Films 520 (2011) 1195–1201 1197

2.3. Photocatalytic activity measurements

The photocatalytic activity of the un-doped and the various Mn–TiO2


samples (0.1–33% mole percentage of Mn) was determined by
employing the PSR/FTIR technique, by monitoring the steady state
concentration of both reactants and volatile photocatalytic products.
The pyrex-glass reactor had an optical length 16.5 cm and an internal
volume 40 cm3. Retractable NaCl windows, transparent in IR light, were
attached at both ends of the reactor in order to insert/remove the glass-
bed containing the photocatalyst. Acetaldehyde was admitted in the
reactor through a separate inlet and the pressure was measured by using
membrane-diaphragmatic transducers. Typical acetaldehyde concen-
trations were ~1.0 × 1017 molecule cm− 3. Infrared spectra were re-
corded over the range 500–4000 cm− 1 at a resolution of 1.0 cm− 1.
Fig. 2. Schematic flowchart of Mn doped TiO2 synthetic procedure.
Spectral subtraction on the infrared absorption bands of acetaldehyde
was performed using a least-squares fitting routine. Reaction products
alternatively it was filtered under vacuum, to obtain a powder (Fig. 2, were identified using known reference spectra.
steps 3–4). The powder was then washed (Fig. 2, step 5) with distilled The degradation of CH3CHO induced by the photocatalytic activity
water until to eliminate sulfate and ammonium ions. The powder was of Mn–TiO2 samples was evaluated by performing a series of separate
free of sulfate and ammonium ions when the test of sulfate and experiments. Initially, photolysis experiments were performed in gas
ammonium was negative. If the test was positive, the procedure (Fig. 2, mixtures of CH3CHO (~ 400 Pa) and synthetic air (20.5% O2 in N2) at
step 5) was repeated. The powders, after drying at 100 °C, were total pressure ~ 1.01 × 10 5 Pa, in the absence of Mn–TiO2 photocata-
calcined for 3 h at 700 °C. MnO2 loaded in TiO2 with molar ration in lyst, and the IR spectra were taken as reference spectra. The mixtures
different concentrations between 0 and 33 wt.% was estimated from were irradiated using the above two light sources (UV and Visible) for
the initial materials proportion. The presence of manganese was about 3 h and the IR spectra were recorded every 20 min, in order to
qualitatively verified by EDX spectroscopy. The amorphous powder confirm that no photodissociation or wall loses of CH3CHO that take
materials consisted of small irregular grains with crystalline structure place under those conditions. No detectable change in CH3CHO con-
mainly of anatase phase. Powder X-ray diffraction patterns were centration was observed during the course of these experiments.
obtained on a Rigaku D/MAX-2000H rotating anode diffractometer Consequently, two separate sets of measurements were performed in
(CuKα radiation) equipped with the secondary pyrolytic graphite the presence of the photocatalyst. In particular, the reactor was
monochromator operated at 40 kV and 80 mA over the 2θ collection initially covered with a black cloth, in order to inhibit natural lab light
range of 20–80°. The scan rate was 0.02°/s. The XRD pattern of to initiate any photocatalytic activity of the material and the glass-bed
composite nanostructure powder of 0.1% Mn:TiO2 calcined at 700 °C is with the photocatalyst was inserted into the reactor. Then the gas
shown in Fig. 3. The anatase phase was detected with corresponding mixture was introduced in the reactor following the above procedure,
values at 2θ = 25.3° for plane (101), 38° for (004), and 48° for (200). and the IR spectra (typically every ~ 10 min) were recorded con-
The grain size (D) was in the nanometer range of 26 nm and it tinuously in order to measure the uptake of CH3CHO on Mn–TiO2
was determined from the full width (w) at half maximum (FWHM) surface. The duration of these measurements was ca. 1 h. It is worth
of the (101) anatase peak according to the Scherrer's formula, noting that during the blank experiments, the reactor was not
D = kλ/wcosθ, where k is the shape factor (~ 0.9), λ the X-ray exposed to any UV or Vis-light other than the natural lab light (UV and
wavelength (0.15418 nm), and θ the diffraction angle.The photocata- Vis lamps were switched off) and reactor temperature was at room
lyst samples were also oven-heated at 343 K for ~ 30 min prior to use, temperature (T ~ 296 K). Then the gas mixture and the photocatalyst
in order to remove water and other volatile species. This sample were removed from the reactor, and the exact same mass of fresh
pretreatment was essential in all cases in order to compare blank catalyst was re-inserted into the reactor along with the particular gas-
(lamps-off) and photocatalytic experiments. mixture. Initially, the reference IR spectrum was taken, and sub-
sequently the IR spectra, in the presence of light were recorded at
identical time intervals as in the previous adsorption experiments
(blank experiments). In the UV and/or Vis irradiation experiments,
the reactor was fan-cooled to keep samples at room temperature, and
therefore degradations of CH3CHO with and without irradiation may
be compared. Temperature was continuously monitored via a
thermocouple that was attached to the reactor, and it was invariant,
±1 K (~296 K) within photocatalysis experiments. This procedure
was followed in all experiments with Mn–TiO2 and un-doped TiO2
samples. Experiments were performed by using photocatalyst sam-
ples of 0.100 ± 0.001 g. Samples surface area was A = 4.5 cm 2 for all
measurements carried out in present work, while catalyst's thickness,
d, was estimated by the expression: d = m/(ρ × A), to be ~50 μm, using
mass, m = 0.1 g and TiO2 density, ρ = 4.25 g cm − 3.
Finally, additional experiments were performed with different
but constant Mn–TiO2 mass samples, exhibiting maximum physical
adsorption of CH3CHO, in order to compare the photocatalytic efficiency
of the different Mn–TiO2 samples. Subsequently, following the above
filling procedure, the reactor was irradiated and IR spectra were
Fig. 3. Powder XRD pattern of 0.1% Mn doped TiO2 calcined at 700 °C for 3 h. Anatase
recorded. Thus, the degradation rate of CH3CHO on different Mn–TiO2
phase is shown with corresponding values at 2θ = 25.3° for plane (101), 38° for (004), samples under UV or visible irradiation provided with a qualitative
and 8° for (200). measure for evaluating their photocatalytic efficiency.
1198 V.C. Papadimitriou et al. / Thin Solid Films 520 (2011) 1195–1201

3. Results and discussion

Typical experimental results showing the photocatalytic activity of


various Mn–TiO2 samples are shown in Fig. 4a–d. It is worth noting
that all experiments performed using identical fresh samples with
certain Mn-doping level of TiO2 lead to identical trends regarding the
UV and Vis light photocatalytic activity. Furthermore, the gaseous
photocatalytic end products were specified at the above experimental
conditions. The build-up of CO2 under UV and visible light is given in
Fig. 5. UV–Vis absorption spectra were taken utilizing a UV/Vis
spectrophotometer (Perkin-Elmer, Lambda 950) in reflection mode
and are presented in Fig. 6. Spectra were recorded at resolution of
2 nm between 250 and 2500 nm with a scan rate of 200 nm/min.

3.1. The role of Mn in the photocatalytic activity of TiO2

3.1.1. Photocatalytic activity of un-doped TiO2 samples


Fig. 4a presents the loss of CH3CHO on un-doped TiO2 samples as a
function of time upon irradiation with UV and Vis light sources, and in
the absence of light. The decay of CH3CHO in the absence of light is due
to the heterogeneous physical adsorption of acetaldehyde on TiO2
surface. It is clearly shown that in the presence of UV light the decay of
CH3CHO becomes more pronounced due to the photocatalytic activity
of the TiO2 surface. On the contrary, in the presence of visible light the
decay is similar to that in the absence of light within experimental
precision. This clearly indicates that the UV irradiation resulted in
CH3CHO decompositions, in agreement with well-known photocata-
lytic behavior of un-doped TiO2 surfaces [19–21].

3.1.2. Photocatalytic activity of ≥5% Mn–TiO2 samples


Then, the photocatalytic activity of several highly doped Mn–TiO2
samples (5, 10 and 33%) was studied in the presence of UV or visible
light irradiation. Fig. 4b presents the loss of CH3CHO as function of
time on 5% Mn–TiO2 upon UV and visible irradiation, and in the
absence of light. The decay of CH3CHO under UV irradiation is slightly
more pronounced than in the absence of light, within experimental
precision. On the other hand, no photocatalytic decomposition of
CH3CHO was observed under visible irradiation. Similarly, the TiO2
samples doped with higher levels of Mn (10 and 33%) showed minor
photocatalytic activity in the UV, and none in the visible range. This
trend was identical to that observed in un-doped TiO2 sample. Fig. 4. Typical CH3CHO photocatalytic decay profiles for (a) un-doped TiO2 samples,
(b) 5%, (c) 1% and (d) 0.1% Mn–TiO2 samples in the absence or presence of UV and
3.1.3. Photocatalytic activity of 1% Mn–TiO2 samples visible light, at 296 K using ~ 400 Pa of CH3CHO, balanced with ~ 1.01 × 105 Pa of
Furthermore, the photocatalytic activity of 1% Mn–TiO2 samples was synthetic air (80.5% N2 and 19.5% O2). Red-shift of the photocatalytic activity and best
visible light performance observed for 0.1% Mn–TiO2 samples, at which UV light
studied, and Fig. 4c presents the loss of CH3CHO on 1% Mn–TiO2 samples irradiation had a minor effect.
as function of time upon UV and visible irradiation, and in the absence of
light. Typical time profiles of CH3CHO due only to uptake (blank
experiments: in absence to any light exposure of the photocatalyst), adsorption process of CH3CHO on 0.1% Mn–TiO2 is more pronounced
before UV and Vis light irradiation, are also presented. Note that these than on un-doped TiO2, probably due to drastic changes on TiO2
CH3CHO uptake profiles were identical within the experimental surface structure upon low level doping with Mn. Most important, it is
precision for all samples. Therefore, Fig. 4a, b and c show only one observed that CH3CHO loss becomes substantially higher upon visible
blank experimental result for clarity and simplicity purposes. It appears light irradiation (≈10%) than in the absence of light, while remains
that the photocatalytic loss of CH3CHO under visible light is less unaffected upon UV irradiation. This behavior of CH3CHO loss is
pronounced compared to the one observed in the presence of UV light. completely opposite to the one observed either on un-doped TiO2 or
This trend is similar but more pronounced than that observed in un- on higher level doping with Mn (≥1%). It is worth noting that the
doped TiO2 samples, indicating an increased photocatalytic activity of decays of CH3CHO in the presence and absence of UV light converge at
the sample, most likely due to substantial changes in the surface longer experimental times. It was expected that the presence of Mn in
structure and induced electron–hole recombination rate discussed later. the TiO2 matrix will reduce the energy gap between valence and
conduction bands of TiO2 and result in a red-shift of the absorption
3.1.4. Photocatalytic activity of 0.1% Mn–TiO2 samples spectrum of the photocatalyst facilitating the creation of carriers
Finally, the photocatalytic activity of 0.1% Mn–TiO2 samples was under visible light irradiation [22]. Such an absorption spectrum was
studied. Fig. 4d presents the loss of CH3CHO on 0.1% Mn–TiO2 samples obtained as a function of Mn presence and was compared to a
as a function of time in the presence of UV or visible light irradiation as commercial “un-doped” TiO2 powder from Degussa (Fig. 6). It is
well as in the absence of light. At first it appears that in the absence of clearly shown that the optical properties of our un-doped material are
light, the CH3CHO loss on 0.1% Mn–TiO2 is higher than on un-doped significantly different compared to Degussa, for which no information
TiO2 surfaces by ca. 10%. This clearly shows that the physical on its composition or starting materials were available. Therefore, the
V.C. Papadimitriou et al. / Thin Solid Films 520 (2011) 1195–1201 1199

effect of Mn doping on TiO2 photocatalytic activity was assessed based


on the comparison of the un-doped TiO2 sample with the Mn–TiO2
doped powders, both synthesized in our labs. First, for the lowest level of
Mn-doping on TiO2 powders (0.1%), a small decrease in absorption
intensity observed on the value of the TiO2 band-gap in the range of
3.15 eV. Moreover in the case of 0.1% Mn doped TiO2, the sub-band-gap
response was present in the whole range from 3.15 eV to almost 2 eV
demonstrating that the introduction of Mn in the TiO2 lattice does
increase significantly the material sub-band-gap absorption (Fig. 6).
This relative increase was estimated to be of the order of ~50% compared
to the un-doped powder at E = 3.0 eV. It is worth noting that such a
relative increase in absorption was even more pronounced for the 1%
Mn doped powder.
The above observation provides an explanation for the fact that the
0.1% Mn-doped sample shows a photocatalytic activity for visible light,
but does not explain why there is an optimum in the photocatalytic
efficiency at this (0.1%) doping concentration. Doping of TiO2 samples
with higher levels than 1% of Mn, although it results in higher
absorbance in the visible, it leads to substantial lower photocatalytic
decomposition yields of CH3CHO under both UV and visible irradiation.
Such an effect may be due to drastic changes on the surface structure of
the photocatalyst that increases the amount of defects and induces high
Fig. 5. Typical CO2 formation profiles during photocatalytic decomposition of CH3CHO
electron–hole recombination rates. Table 1 summarizes the compara- on un-doped TiO2 and doped with Mn samples under UV and visible light irradiation.
tive results for all the photocatalytic samples used. It is clearly shown CO2 yield was much higher under UV light than under visible light irradiation, which
that 0.1% Mn doped TiO2 sample is the most effective photocatalyst on indicates that the degradation mechanism of CH3CHO on Mn–TiO2 is different than on
decomposing CH3CHO when exposed in visible irradiation. TiO2.

3.2. Photocatalytic end-products analysis used as painting additives to remove indoor pollution, where visible
light dominates. Our experimental results show that Mn doping
The end-product analysis of the volatile species in the photocatalytic changes substantially the physical–chemical properties of TiO2. At
degradation of CH3CHO on un-doped TiO2 and Mn–TiO2 has shown that first, Mn–TiO2 samples exhibit systematic and substantially higher
carbon dioxide was the major product in all cases. For un-doped TiO2 uptakes than un-doped TiO2 samples under dark conditions, which
and high levels of Mn–TiO2 (N5%) samples, the CO2 yields were indicates that Mn doping causes significant changes in the surface
negligible under both UV and visible light irradiation. Fig. 5 presents the structure of TiO2 due to higher or more active sites (O, Ti, H) [23,24].
yield of CO2 versus time due to photocatalytic decomposition of CH3CHO Moreover, although the photocatalytic activity of TiO2 is mainly
on un-doped TiO2 and on 0.1 and 1% Mn–TiO2 samples, under both UV effective under UV light irradiation [1,2,14,25,26], doping with Mn in
and visible light irradiation. The CO2 yields were higher in UV irradiation the range 0.1 and 1% enhances further this efficiency, probably by
experiments and were increasing with the lower level of Mn doping in inducing higher UV absorption strength. On the other hand, at high
TiO2. A small amount of the toxic carbon monoxide (CO) was also levels of Mn doping it appears that this efficiency is reduced con-
detected. No substantial CO2 yields were observed in the visible siderably, indicating that the level of Mn doping in TiO2 is very crucial
irradiation experiments in agreement with the low CH3CHO losses to their photocatalytic activity. Similar dependence of the photo-
observed previously. However, it is surprising that the high CH3CHO catalytic activity on the level of Mn doping in TiO2 was observed under
losses observed with 0.1% Mn–TiO2 samples under visible light visible light irradiation. This behavior may be understood in the
irradiation correspond to low CO2 yields, orders of magnitude lower context of the following discussion:
than employing UV light, as it is depicted in Fig. 5. This observation is In general, the introduction of some amount of metal dopants in
elaborated in the discussion below and probably indicates that visible TiO2 matrix is known to improve light absorption in the visible region
photons initiate different photocatalytic decomposition mechanism of by inserting electronic states and thereby narrowing the band gap.
CH3CHO than UV photons, which is consistent with the observations in Such a relative increase in the sub-band-gap absorption spectra as a
the photocatalytic experiments using the 1% Mn–TiO2 sample, but the function of Mn doped TiO2 was measured (Fig. 6). However the
above behavior needs further investigation. observed existence of an optimum in dopant concentration with a
subsequent enhanced catalytic behavior of the 0.1% Mn doped TiO2
4. Remarks and conclusions powder, in the present study, is mainly attributed to the role of Mn in
trapping of the photogenerated charge carriers under both UV and
The role of Mn doping in the photocatalytic degradation (UV and visible light illumination [27,28] and the electron–hole recombination
visible irradiation sources) of gaseous acetaldehyde on TiO2 samples mechanism [29].
was investigated in order to design materials that can potentially be Indeed, Mn 2+ has valence electronic configuration of 3d 5. When it
traps electrons the electronic configuration changes to d 6 and if it
traps holes its electronic configuration changes to d 4, both highly
Table 1 unstable. To restore its stable electronic configuration, the trapped
CH3CHO decomposition (%) on pure TiO2 and various Mn–TiO2 surfaces in the absence
electron will be transferred to oxygen molecule and trapped hole to
and presence of UV or visible light after 1 h. The highest values are denoted in bold.
surface adsorbed water molecules or to dye molecule to generate
TiO2 substrates Dark UV Visible superoxide (O2−) and hydroxyl (OH•) radicals as shown by Devi et al.
Un-doped 8.5 13 11.5 [27] in line with the following Eqs. (1)–(4).
0.1% Mn 17 17 26
1% Mn 14 23 20
5% Mn 2.7 5.6 4.5 2þ − þ
Mn þ e →Mn ð1Þ
1200 V.C. Papadimitriou et al. / Thin Solid Films 520 (2011) 1195–1201

electron–hole pairs in the semiconductor becomes easier. At a high


dopant concentration with induced large number of structural defects
and trap centers, the charge carriers may also be trapped more than
once on their way to the surface so that their mobility becomes
extremely low and undergoes recombination before they can reach
the surface. In addition at higher concentrations, these trap centers
are close to one another and the trapped charge carriers can recom-
bine, the probability of which increases with increase in the dopant
concentration [30].
Hence it is highly probable that an optimum concentration of
doping ions may exist for which the space charge thickness layer is
comparable to the light penetration depth [31]. It is encouraging that
he reported values of 20–30 nm by L.G. Devi et al. [27,28] are in the
same range (10–40 nm) as in our material. This explanation proposed
also for Ni and Zn dopants is in line with our results showing an
optimal dopant concentration for 0.1% Mn in the TiO2 matrix.
The end-products analysis has shown that CO2 is the major
photocatalytic product in CH3CHO degradation under either UV or
visible irradiation. In addition, UV light activation of the photocatalyst
Fig. 6. UV absorbtion spectra for Degussa, un-doped and 0.1, 1, and 5% Mn–TiO2
samples. Mn-doping of titania, induces sub-band-gap absorbance contributing to the resulted in much higher CO2 yields than visible light activation. This
photocatalytic activity of Mn-doped TiO2. finding provides strong evidence that the two degradation processes
occur via different reaction mechanisms. It is probable that the visible
þ 2þ photons energy is not sufficient to dissociate the C\CH3 bond in the
Mn þ O2 ads→Mn þ O2 − ð2Þ acetaldehyde diolate intermediate, considering the decomposition
reaction mechanism for acetone degradation on TiO2 proposed by
Henderson [7]. Therefore, the visible light irradiation initiates an
2þ þ 3þ alternative photocatalytic degradation mechanism of CH3CHO on 0.1%
Mn þ h →Mn ð3Þ
Mn–TiO2 samples. However, more experiments are necessary to
determine the above photocatalytic decomposition of CH3CHO on
Mn–TiO2 samples under visible light irradiation.
3þ − 2þ In conclusion, it was found that 0.1% Mn doped TiO2 powder resulted
Mn þ OH →Mn þ OH• ð4Þ
in the highest photocatalytic loss of CH3CHO under visible irradiation.
This optimized efficiency, despite an observed enhancement in the
The enhanced activity of Mn 2+ may also be attributed to its half relative absorption as a function of Mn concentration, it was found to
filled electronic structure, which serves as shallow trap for the charge drastically diminish at higher levels of Mn doping (1–33%) due to
carriers to accelerate charge transfer processes. These processes may induced high electron–hole recombination rates. In addition it was
enhance the generation of highly reactive oxidative species like shown that CO2 is the major degradation product of CH3CHO. The CO2
superoxide and hydroxyl radicals. yields were the highest for 0.1% Mn–TiO2 samples under UV irradiation,
It is widely accepted that the mechanisms involved in photo- in agreement with the observed highest CH3CHO decomposition rates.
catalysis are complex and multiple, so it is difficult to predict the exact Since the above photocatalytic degradation process, despite of
role of dopants. Indeed, the structural changes of the TiO2 matrix, harmful intermediaries, such as CO and CH3COOH, results in, mainly,
upon doping, are strongly dependent on the charge and the ionic CO2 – an environmental friendly gas with minor impact on global
radius of the dopant. The dopant with a lower charge than Ti 4+ can warming at these levels of production – one may expect a broad
alter the concentration of oxygen vacancies depending on their posi- application up-take. Thus 0.1% Mn–TiO2 powders may be utilized as a
tion in TiO2 matrix; they can replace Ti in TiO2 lattice or can occupy photocatalytic additive in various cleaning agents to combat not only
interstitial position which in turn depends on their size and outdoor, but also indoor air pollution as soon as levels of potential
concentration. According to L.G. Devi et al. [27,28], due to the larger harmful compounds (CO and CH3COOH) have been clarified by
ionic radius of Mn 2+ (0.80 Å) compared to Ti 4+ (0.68 Å), it is im- independent studies at the JRC center in ISPRA/Italy utilizing their
possible for Mn 2+ ions to act as interstitial ions in the TiO2 matrix. certified INDORTRON facility in the framework of the FP7 IP project
Hence Mn 2+ ions may only replace Ti 4+ in the lattice sites. The “Clear-up” [32].
substitution of metal ion with valence less than +4 and higher ionic
radius would induce oxygen vacancies at the surface of anatase grains
Acknowledgments
which favors bond rupture and solid state ionic rearrangement. At
some dopant concentration and temperature in the range of 700 °C
Funding through the E.C. FP7 “Clear-Up” No 211948 is acknowledged.
and 800 °C, the dopant ions stabilize the anatase phase of TiO2.
A plausible explanation to the existence of an optimum in dopant
concentration has been reported by Pleskov [29], who has shown that References
for the efficient separation of electron–hole pairs the value of the [1] K. Hashimoto, H. Irie, A. Fujishima, Jnp. J. Appl. Phys. 44 (2005) 8269.
space charge region potential has a lower limit. He has shown that as [2] T.L. Thompson, J.T. Yates, Chem. Rev. 106 (2008) 4428.
the concentration of doping ions increases, the surface barrier [3] C.R. Usher, A.E. Michel, V.H. Grassian, Chem. Rev. 103 (2008) 4883.
[4] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Chem. Rev. 95 (1995) 69.
becomes higher, and the space charge region becomes narrower,
[5] W.J. Liang, J. Li, Y.Q. Jin, J. Environ. Sci. Health A45 (2010) 1384.
the electron–hole pairs within the region are efficiently separated [6] M. Ndour, B. D'Anna, C. George, O. Ka, Y. Balkanski, J. Kleffmann, K. Stemmler, M.
before recombination. However, as the concentration of doping ions is Ammann, Geophys. Res. Lett. 35 (2008).
excessively high, the space charge region becomes too narrow and the [7] M.A. Henderson, J. Phys. Chem. C 112 (2008) 11433.
[8] P. Pichat, Appl. Catal. B: Environ 99 (2010) 428.
light penetration depth into TiO2 considerably exceeds the space [9] C.M. Schmidt, A.M. Buchbinder, E. Weitz, F.M. Geiger, J. Phys. Chem. A 111 (2007)
charge layer; therefore the recombination of the photo-induced 13023.
V.C. Papadimitriou et al. / Thin Solid Films 520 (2011) 1195–1201 1201

[10] S. Yamazaki, K. Ichikawa, A. Saeki, T. Tanimura, K. Adachir, J. Phys. Chem. A 114 [23] N.A. Deskins, R. Rousseau, M. Dupuis, J. Phys. Chem. C 114 (2010) 5891.
(2010) 5092. [24] A.C. Papageorgiou, N.S. Beglitis, C.L. Pang, G. Teobaldi, G. Cabailha, Q. Chen, A.J.
[11] T. Maggos, J.G. Bartzis, P. Leva, D. Kotzias, Appl. Phys. A 89 (2007) 81. Fisher, W.A. Hofer, G. Thornton, Proc. Natl. Acad. Sci. U.S.A. 107 (2010) 2391.
[12] A.R. Ravishankara, Science 276 (1997) 1058. [25] J.M. Herrmann, Appl. Catal. B: Environ. 99 (2010) 461.
[13] G. Rubasinghege, S. Elzey, J. Baltrusaitis, P.M. Jayaweera, V.H. Grassian, J. Phys. [26] M. Kaneko, I. Okura, Photocatalysis: Science and Technology, Kodansha, Tokyo,
Chem. Lett. 1 (2010) 1729. 2003.
[14] U. Diebold, Surf. Sci. Rep. 48 (2003). [27] L.G. Devi, N. Kottama, B.N. Murthya, S.G. Kumara, J. Mol. Catal. A: Chem. 328
[15] R.T. Zehr, M.A. Henderson, Surf. Sci. 602 (2008) 2238. (2010) 44.
[16] J. Raskó, J. Kiss, Appl. Catal. A 287 (2005). [28] L.G. Devi, S.G. Kumara, B.N. Murthya, N. Kottama, Catal. Commun. 10 (2009) 794.
[17] M.A. Henderson, J. Phys. Chem. B 109 (2005) 12062. [29] Y.V. Pleskov, Soviet Electrochem. 17 (1981) 1.
[18] G. Kiriakidis, V. Tudose, V. Binas, OBI, Greece, Patent No. GR-20090100724, 2009. [30] W. Zhou, Q. Liu, Z. Zhu, J. Zhang, J. Phys. D: Appl. Phys. 43 (2010) 035301 6 pp.
[19] A. Fujishima, K. Honda, Nature 238 (1972) 37. [31] L. Palmisano, V. Augugliaro, A. Sclafani, M. Schiavello, J. Phys. Chem. A 92 (1988)
[20] A. Fujishima, T.N. Rao, D.A. Tryk, J. Photochem. Photobiol. C 1 (2000) 1. 6710.
[21] H. Tada, M. Tanaka, Langmuir 13 (1997) 360. [32] C. Cacho, O. Geiss, J. Barrero-Moreno, V.D. Binas, G. Kiriakidis, L. Botalico, D.
[22] A. Zaleska, Rec. Pat. Eng. (2008) 157. Kotzias, to be published to: J. Photochem. Photobiol. A (2011).

You might also like