Vous êtes sur la page 1sur 6

Applied Catalysis B: Environmental 24 (2000) L7–L12

Letter
The comparative studies of titanium dioxide in gas-phase ethanol
photocatalytic oxidation by the FTIR in situ method
D.V. Kozlov ∗ , E.A. Paukshtis, E.N. Savinov
Novosibirsk State University, Boreskov Institute of Catalysis, Prospekt Akademika Lavrentieva 5, Novosibirsk 630090, Russia
Received 4 April 1999; received in revised form 20 July 1999; accepted 20 July 1999

Abstract
Gas-phase ethanol photocatalytic oxidation on the series of TiO2 samples has been studied by the FTIR in situ method. The
TiO2 samples were prepared by the usual method and they possessed similar physicochemical properties (porosity, specific
area), but were found to have different photocatalytic activities. The acidity and concentration of the Lewis centers have been
measured. The TiO2 photocatalytic activity has been found to correlate with the surface acidity, and the Ti3+ and carbonate
surface concentrations. ©2000 Elsevier Science B.V. All rights reserved.
Keywords: Photocatalysis; TiO2 ; FTIR; Surface acidity

1. Introduction this problem because its kinetics and mechanism are


well-studied [5–9].
Today, the gas-phase photocatalytic oxidation of or-
ganic admixtures in the air is the rapidly developing
field of investigations [1,2]. It is possible to oxidize 2. Experimental
any organic substances to CO2 and H2 O with hetero-
geneous photocatalysts [3,4]. Consequently, photocat- The highly purified ethanol ‘REACHIM’ was used
alytic processes have already been used in air and wa- in the experiment. We also used the series of TiO2
ter purification. Nevertheless, the reasons causing the samples H19 which were prepared by the hydrolysis
difference in photocatalytic activity of the TiO2 sam- of TiCl4 at the Institute of new chemical problems
ples of the same origin and morphology was not clear. under the direction of Professor Troitsky V.N.
The object of this work was to find out the cor- The H19 series, in which the samples differed in
relation between the TiO2 photocatalytic activity and the pH value of precipitation, was prepared by the
its physicochemical properties. The reaction of deep hydrolysis of TiCl4 under the pH 4.0, 5.8, 8.6, followed
ethanol photooxidation has been chosen to elucidate by annealing at 400◦ C for 1 h.
The characteristic of activity of TiO2 samples was
considered to be the quantum yield ϕ in the reaction
of gas-phase acetone photocatalytic oxidation [10]:
∗ Corresponding author. Fax: +7-383-235-5766

E-mail address: kdv@catalysis.nsk.su (D.V. Kozlov) CH3 COCH3 + 4O2 + 16 hν → 3CO2 + 3H2 O

0926-3373/00/$ – see front matter ©2000 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 6 - 3 3 7 3 ( 9 9 ) 0 0 0 9 5 - 8
L8 D.V. Kozlov et al. / Applied Catalysis B: Environmental 24 (2000) L7–L12

Table 1
Brief descriptions of the TiO2 samples used
Sample Crystalline modification ϕ (%) Specific surface (m2 /g) Preparation

H19-pH 4.0 anatase 61 81 hydrolysis of TiCl4 , precipitation at pH 4.0,


annealing at 400◦ C for 1 h
H19-pH 5.8 anatase 34 112 precipitation at pH 5.8
H19-pH 8.6 anatase 22 125 precipitation at pH 8.6

All data concerning samples are summarized in


Table 1.
Unfortunately, we do not have the data on the con-
tent of catalyst impurities (Cl− , Na+ . . . ), but we sup-
pose that the catalyst surface may contain a small
amount of chloride ion and carbon. This is evident
from the TiO2 preparation method in which hydrochlo-
ric acid is formed during TiCl4 hydrolysis and car-
bon is formed during TiO2 calcination at 400◦ C out
of doors. No doubt that these impurities have an in-
fluence on the TiO2 photocatalytic activity, but this
influence is equal for all three samples which are re-
ported in Table 1. So it is not important in our compar-
ative study. Samples were prepared by two methods. Fig. 1. Cell for the FTIR in situ photocatalytic oxidation of organic
The samples for the FTIR in situ photocatalytic oxi- compounds.
dation were prepared as follows. The aliquot of TiO2
water suspension was evenly supported on CaF2 glass 500 ppm vapor concentration was adjusted. Then, we
(25 mm × 30 mm). Then, the samples were dried at turned on the high pressure Hg lamp (1000 W) and
room temperature during 12 h. periodically registered the surface IR spectra. The
The samples for the investigation of low-temperature light intensity was 20–25 mW/cm2 . CO adsorption
CO adsorption were prepared by TiO2 powder press- was carried out at −110◦ C and CO was added to the
ing into the tablets under a pressure of 100 bar. cell by portions.
The typical density of the tablets obtained was
7–10 mg/cm2 . All IR spectra measurements were
carried out on the Fourier transform spectrometer 3. Results and discussion
‘Bruker IFS 113v’.
Then the samples were placed in the cell (Fig. 1), 3.1. FTIR in situ ethanol photooxidation
which was installed in the spectrometer and linked
with the vacuum pump. The sample treatment included Drastic changes in the IR spectra of ethanol ad-
1 h of heating at 120◦ C in vacuum (10−5 bar) fol- sorbed at room temperature on titanium dioxide dur-
lowed by 1 h of heating at 120◦ C in oxygen atmo- ing its irradiation were observed. These changes are
sphere (65 mbar). Then, we pumped out the residual well seen for the TiO2 sample H19-pH 8.6 and are
oxygen. This method enables to get rid of the physi- shown in Fig. 2.
cally adsorbed water, leaving the activity of the TiO2 Fig. 3 presents the IR spectrum of the same sam-
samples unaffected. ple after 1 h and 42 min of irradiation (surface steady
For the investigation of gas-phase ethanol pho- state setting-up time). The assignment of the adsorp-
tocatalytic oxidation, we injected ethanol into the tion bands to surface structures is based on the data
cell, where the treated sample was placed, while the presented in [11–14].
D.V. Kozlov et al. / Applied Catalysis B: Environmental 24 (2000) L7–L12 L9

In this way, during ethanol photooxidation on


the catalyst surface, we can detect acetic acid
(ν asCO = 1750 cm−1 ), acetaldehyde (ν asCO = 1725
cm−1 ), adsorbed carbonyl species (ν asCO = 1667 cm−1 ),
water (δ HOH = 1629 cm−1 ), two types of adsorbed
carboxyl species (ν asCOO = 1568 and 1517 cm−1 ) and
surface carbonate (ν as = 1437 cm−1 ). It should be
noted that appropriate symmetric adsorption bands lie
in the area below 1400 cm−1 and we cannot correlate
them unambiguously with the corresponding antisym-
metric adsorption bands. This is the reason why we
do not report them. The registration of these bands
is adjusted with the following scheme of ethanol
photooxidation:

From Fig. 2, we can see that, in the beginning of


ethanol photooxidation, bands 1 and 2 (which are at-
Fig. 2. The changes in IR spectra of the TiO2 sample H19-pH tributed to acetic acid and acetaldehyde) have com-
8.6 (ϕ = 22%) during ethanol photocatalytic oxidation. (a) Ethanol parable areas (spectra b, c) and then they both rise
adsorption on the TiO2 surface (vapor concentration 200 ppm); (b) (spectra d, e, f), but the rate of rising of the acetic acid
after 10 min of irradiation; (c) 34 min; (d) 1 h; (e) 1 h and 20 min; adsorption band is greater than that of the acetalde-
(f) 1 h and 42 min.
hyde adsorption band. This is so because acetic acid
is a more stable compound than acetaldehyde. We do
not observe the following decrease in the acetaldehyde
adsorption band because of the small rate of ethanol
photooxidation. In our case, only about 10% of initial
ethanol was oxidized after 1 h and 42 min of irradia-
tion. So the spectra f (Fig. 2) presents the steady state
of TiO2 surface during ethanol photooxidation.
From Fig. 2, we can also see that adsorption band
5 which is attributed to the surface carboxyl species is
already present in the spectra after ethanol adsorption
(spectra a). That is why we can not relate it to pho-
tooxidation and we think that this species forms on
the TiO2 surface due to dark oxidation during ethanol
adsorption.
The IR spectra during ethanol photooxidation for
the TiO2 samples H19-pH 4.0 and H19-pH 5.8 are
reported in Fig. 4. These samples had higher activity
than the H19-pH 8.6 sample.
The comparative analysis of the spectra reported in
Figs. 3 and 4 is rather complicated. But it is evident
Fig. 3. IR spectrum of the TiO2 sample H19-pH 8.6 after 1 h and that the less active the sample is, simpler is its spec-
42 min of irradiation by the full light of the high pressure Hg tra, i.e. a lesser number of intermediate species can
lamp. And the assignment of the adsorption bands. be seen on its surface. For example, chemisorbed car-
L10 D.V. Kozlov et al. / Applied Catalysis B: Environmental 24 (2000) L7–L12

Fig. 5 shows the IR spectra of TiO2 during


low-temperature CO adsorption. The absorption band
2180–2190 cm−1 corresponds to CO interaction with
the surface Ti4+ and 2120 cm−1 corresponds to that
with the surface Ti3+ ions. Table 2 summarizes the
data obtained.
It is clear that as the activity of samples rises, the
absorption band corresponding to Ti4+ interacts with
CO molecule shifts from 2183 to 2190 cm−1 . It means
that the samples activity rises with the increase in the
surface acidity. A similar correlation was observed by
J. Papp et al. [18] for 1,4-dichlorobenzene degradation
in aqueous solution.
Fig. 4. IR spectra of the TiO2 samples I H19-pH 5.8 (ϕ = 34%) An additional point to emphasize is that the band
and II H19-pH 4.0 (ϕ = 61%) during ethanol photooxidation. Let- 2120 cm−1 , corresponding to the CO molecule bonded
ters correspond to: (A) Ethanol adsorption (vapor concentration with the surface Ti3+ ions, is essentially missing from
200 ppm); (B) after 1 h and 30 min of irradiation. the spectrum of the most active sample (ϕ = 64%), is
appreciable for the middle (ϕ = 34%) sample and is
pronounced for the (ϕ = 22%) sample with the lowest
boxyl species RCOO− (bands 1517 and 1568 cm−1 ) activity.
are well-defined in the spectrum of H19-pH 8.6 sam- It must be noted that the amount of Ti4+ ions on
ple and are almost absent in the spectrum of H19-pH the catalyst surface is much more then of Ti3+ ions
4.0 sample. Also, acetic acid (band 1750 cm−1 ), ac- though the Ti3+ adsorption band is greater than the
etaldehyde (band 1725 cm−1 ) and the surface carbon- Ti4+ one (Fig. 5, III). This is due to the fact that the
ate concentrations (band 1437 cm−1 ) decrease as the CO absorbance coefficient (A0 ) for low-charged ions
activity of the TiO2 samples increases. is greater than that for high-charged ions. For example
[15], the A0 for CO adsorption on Cu1+ ions is 10
3.2. Low-temperature CO adsorption times greater than the A0 for CO adsorption on Cu2+
ions and the CO adsorption band is close to that on
For the TiO2 surface to be studied more completely, the Ti3+ ions and lies in the 2120–2140 cm−1 area.
we investigated low-temperature CO adsorption on our Thus, the TiO2 activity in the VOC’s photooxidation
catalysts. processes decreases as the Ti3+ surface concentration
Low-temperature CO adsorption is a widely used rises. The same correlation was found previously [19].
method for measuring surface acidity and its ion In those experiments, surface Ti3+ was formed during
make-up [12,15,16]. It is well known that, in a CO TiO2 reduction by hydrogen.
molecule, one electron pair is on the 3␴ antibonding Based on the data above, the samples of H19 series
MO [17]. So the higher the acidity of the surface which differ in the pH of precipitation can be ordered
Lewis center, stronger is its intertaction with carbon in the increasing order of surface acidity: H19-pH 8.6,
monoxide. The bond in the CO molecule becomes H19-pH 5.8, H19-pH 4.0. This coincided with the
less antibonding and its adsorption band shifts to order of the increase in TiO2 photocatalytic activity.
higher wavenumbers. Moreover, this method allows The lower the acidity of the TiO2 surface, stronger
to estimate the surface ion concentrations accord- is the interaction of carbon acids with it to form the
ing to the formula A = A0 × d × C, where A is the carboxylate structures (RCOO− ). These carboxylate
adsorption band area (cm−1 ), A0 is the absorbance structures, in turn, occupy the surface sites preventing
coefficient for the particular adsorbate, adsorbent and further interaction of the gaseous reagents (ethanol,
absorption wavenumbers (cm/mkmol), d is the tablet water, oxygen) with the catalyst surface. So the higher
density (mg/cm2 ), C is the ion surface concentration the carboxylate concentration on the surface, lower is
(mmol/g). the photooxidation rate.
D.V. Kozlov et al. / Applied Catalysis B: Environmental 24 (2000) L7–L12 L11

Fig. 5. Low-temperature IR spectra of CO adsorption on the TiO2 H19 series (at temperature −110◦ C). I H19-pH 4.0 (ϕ = 61%), II H19-pH
5.8 (ϕ = 34%), III H19-pH 8.6 (ϕ = 22%). (1) 4 mbar (CO pressure); (2) 14 mbar; (3) 28 mbar; (4) 39 mbar; (5) 54 mbar; (6) 64 mbar; (7)
79 mbar.

Table 2
The data obtained in low-temperature CO adsorption experiments on the TiO2 series H19
Sample ν CO (Ti4+ ) (cm−1 ) A(Ti4+ ) (A0 =1.5) [15] C(Ti4+ ) (␮kmol/m2 ) Ti3+ quantity

H19-pH 4.0 2190 3.68 3 very small


H19-pH 5.8 2185 4.97 3 small
H19-pH 8.6 2183 0.46 0.4 appreciable

4. Conclusion [4] M.L. Sauer, D.F. Ollis, J. Catal. 149 (1994) 81.
[5] M.R. Nimlos, E.J. Wolfrum, M.L. Brewer, J.A. Fennell, G.
Bintner, Environ. Sci. Technol. 30 (1996) 3102.
Since the stability of the carboxylate species in-
[6] D.S. Mugli, J.T. McCue, J.L. Falconer, J. Catal. 173 (1998)
creases as the surface acidity decreases, it is the rea- 470.
son for the difference in the photocatalytic activity of [7] M.L. Sauer, D.F. Ollis, J. Catal. 158 (1996) 570.
the TiO2 samples having the same origin and mor- [8] D.S. Muggli, S.A. Larson, J.L. Falconer, J. Phys. Chem. 100
phology. That is, the carbonic acids which form dur- (1996) 15886.
ing the VOC’s photooxidation can be stabilized on the [9] J.C. Kennedy III, A.K. Datye, J. Catal. 179 (1998) 375.
[10] A.V. Vorontsov, E.N. Savinov, G.B. Barannik, V.N. Troitsky,
catalyst surface preventing further interaction of the
V.N. Parmon, Catal. Today 39 (1997) 207.
gaseous reagents with it. [11] G. Busca, J. Lamotte, J.-C. Lavalley, V. Lorenzelli, J. Am.
Chem. Soc. 109 (1987) 17.
[12] A.A. Davydov, IK spectroskopia v himii poverhnosti okislov,
References Nauka, Novosibirsk, 1984 (in Russian).
[13] G. Busca, P. Forzatti, J.C. Lavalley, E. Troncony, in: B.
[1] J. Peral, D.F. Ollis, J. Catal. 136 (1992) 554. Emelik et al. (Eds.), Catalysis by Acids and Bases, Elsevier,
[2] D.F. Ollis, H. Al-Ekabi (Eds.), Photocatalytic purification and Amsterdam, 1985, p. 15.
treatment of water and air, Elsevier, Amsterdam, 1993. [14] G.A.M. Hussein, N. Sheppard, M.I. Zaki, R.B. Fahim, J.
[3] R.M. Alberici, W.F. Jardim, Appl. Catal. B 14 (1997) 55. Chem. Soc., Faraday Trans. 1 85(7) (1989) 1723.
L12 D.V. Kozlov et al. / Applied Catalysis B: Environmental 24 (2000) L7–L12

[15] E.A. Paukshtis, Infrakrasnaya Spektroskopia v geterogehhom [18] J. Papp, S. Soled, K. Dwight, A. Wold, Chem. Mater. 6
kislotno osnovnom katalize, Nauka, Novosibirsk, 1992 (in (1994) 496.
Russian). [19] A.V. Vorontsov, Photokataliticheskoe okislenie gazoobraznyh
[16] K. Hadjiivanov, J. Lamotte, J.-C. Lavalley, Langmuir 13 organicheskih veshestv na poluprovodnikovyh oksidah, Ph.D.
(1997) 3374. Dissertation, Institute of Catalysis SO RAN, 1998.
[17] L.H. Littl, Infrared Spectra of Adsorbed Species, Academic
Press, London, 1966.

Vous aimerez peut-être aussi