Vous êtes sur la page 1sur 10

Applied Catalysis A: General 177 (1999) 111120

Iron-doped titania semiconductor powders prepared by a solgel method. Part I: synthesis and characterization
A. Nav o1,a, Gerardo Colo na, Manuel Mac asa, Concepcio n Reala, Marta I. Litterb,* Jose
a

cas ``Isla de la Cartuja'', Av. Ame rico Vespucio, Instituto de Ciencias de Materiales, Centro de Investigaciones Cient s/n, Isla de la Cartuja, 41092-Sevilla, Spain b mica, Centro Ato mico Constituyentes, Comisio n Nacional de Energ a Ato mica, Unidad de Actividad Qu Av. del Libertador 8250, 1429 Buenos Aires, Argentina Received 2 July 1998; received in revised form 23 July 1998; accepted 23 July 1998

Abstract Samples of iron-doped titania containing different amounts of iron (0.55 wt%) were prepared from TiCl4 and Fe(III) acetylacetonate by the solgel method. The specimens were characterized by XRD, TPD, specic surface area (BET) measurements, SEM-EDX, XPS, atomic absorption, IR and diffuse reectance spectroscopies. From the structural point of view, these samples are different from those obtained by impregnation of TiO2 (P-25) with Fe(NO3)39H2O or Fe(acac)3, seeming to present anatase as the unique phase and a continuous distribution of iron not only on the particle surface but also within particles, together with an external surface titania rich layer. The doped oxides show suitable properties for photocatalytic purposes. # 1999 Elsevier Science B.V. All rights reserved. Keywords: Iron-doped titania; Solgel; Photocatalytic materials

1. Introduction Although TiO2 is the most useful photocatalyst, the search for new materials in heterogeneous photocatalysis continues to be a matter of interest at the present. Attention has been paid to iron-doped titania samples, testing their efciency to replace bare titania. Results have been controversial, depending on multiple factors such as iron-loading, preparation technique, temperature of calcination, iron content, etc. [13]. Generally, no direct correlation between UVvis spec*Corresponding author. Tel.: +54-1-704-1318; fax: +54-1-7041164; e-mail: litter@cnea.edu.ar 1 Corresponding author.

troscopic features and photochemical activity could be found, and the preparation technique seems to be one of the most important factors for the reactivity of the catalyst. In previous papers, we have processed and characterized iron-doped titania samples prepared by impregnation of TiO2 (Degussa P-25) with Fe(NO3)3 9H2O or iron acetylacetonate, Fe(acac)3, as iron precursors [46]. Also, we have tested these materials as photocatalysts for some oxidative and reductive reactions [59]. According to our results, these materials are useful in a wide eld of practical applications for environmental protection. Our previous results [6] suggest that, although the wet impregnation method does not seem to be a suitable procedure to obtain a

0926-860X/99/$ see front matter # 1999 Elsevier Science B.V. All rights reserved. PII: S0926-860X(98)00255-5

112

o et al. / Applied Catalysis A: General 177 (1999) 111120 J.A. Nav

uniform distribution of the dopant (Fe3) into the matrix, the use of Fe(acac)3 instead of Fe(NO3)39H2O yields a more homogeneous distribution of iron for each sample on the particle surfaces, keeping practically unchanged the specic surface area. However, their photocatalytic activity, at least for oligocarboxylic acids [6], is still lower than that of pure TiO2 (Degussa P-25), as in the case of Fe/Ti oxide samples prepared from iron nitrate [5]. The decreased activity was attributed to the fact that, in these particles, dopants act more as recombination centers than as trap sites for charge transfer [10]. However, samples with a high degree of hydroxylation present good activity in systems that involve hydroxyl radicals participating in rapid charge-transfer processes [8]. The extent of doping is also important, because the existence of separated hematite or pseudobrookite (Fe2TiO5) phases in samples containing more than 2 wt% iron can decrease the activity. Other factors that can inuence the photocatalytic efciency to some extent can be a larger particle size, a lower amount of surface hydroxyl groups and a lower anatase-to-rutile ratio compared with the TiO2 precursor sample. In the present work, a different technique was used to prepare new samples of iron-doped titania. An alkaline coprecipitation (solgel) method from TiCl4 and Fe(acac)3 was chosen because the lower temperature of the initial steps could lead to a higher photocatalytic activity as compared with samples prepared by impregnation. In addition, this method could provide, in principle, a better homogeneity of iron on the surface and inside the particles. The new materials were characterized by different techniques. In Part II of this work [11] we will present the results of photocatalytic tests, showing that the efciency is, however, affected by other structural and surface properties of the samples such as the amount of defects generated by the preparation technique. Some preliminary results have been recently published as part of a more general work related to the improvement of the photocatalytic efciency of TiO2 [12]. 2. Experimental TiO2 Degussa P-25 (Dg) was a commercial sample, kindly provided by the manufacturer (Degussa AG, Germany), and used as provided.

Iron-doped titania powders prepared by impregnation of TiO2 (Degussa P-25) with Fe(NO3)39H2O (named hereafter Fe/Ti(n)) or Fe(acac)3 (named hereafter Fe/Ti(a)) were the same as used in previous papers [48]. Solgel iron-doped samples (named hereafter Fe/Ti(sg)) were synthesized from Fe(acac)3, (Aldrich, 99%) and TiCl4 (Aldrich, 99.995%), according to the following procedure. Aqueous solutions of Fe(acac)3, containing different nominal concentrations of iron, were mixed at 273 K under nitrogen atmosphere with the corresponding amount of TiCl4 to give Fe/TiO2 specimens with nominal ratios ranging from 0.5 to 5 wt% Fe. Aqueous ammonia (Panreac, 28 wt%) was added dropwise to the mixture with continuous stirring at pH 9.09.5. After gelation (30 min for all the prepared samples), the solids were ltered and repeatedly washed until no chloride was detected by AgNO3 test, to ensure the absence of chloride ions in the samples, as conrmed by XPS analysis. The samples were dried overnight at 383 K for 24 h, and the resulting powders red in air at 773 K for 24 h. The ow sheet of the preparation method is reported in Scheme 1. A home prepared TiO2 solid (TiO2(hp)) was obtained by the same procedure above but in the absence of Fe(acac)3, and a home prepared hematite (a-Fe2O3(hp)) was obtained similarly but in the absence of TiCl4. The iron content present in the Fe/TiO2 samples was checked by atomic absorption using a Perkin-Elmer model 2380 spectrometer, and the iron content was very close to the nominal wt% for all the analyzed samples. X-ray diffraction (XRD) patterns of the samples were obtained at room temperature with a Philips PW 1060 diffractometer using Ni-ltered Cu K radiation. Scanning electron microscopy (SEM) was performed on gold-coated samples using a JEOL apparatus, model JSM-5400, equipped with a Link analyzer, model ISIS, for energy-dispersive spectroscopy (EDX). Specic surface areas (SBET) were obtained with an automatic system (Micromeritics 2200 A) with nitrogen gas as adsorbate at the liquid nitrogen temperature. Diffuse reectance spectra (DRS) were obtained on a Shimadzu 210A spectrophotometer, equipped with an integrating sphere, using MgO as the reference. Characterization of the surfaces was carried out using a combination of temperature-programmed desorption (TPD),

o et al. / Applied Catalysis A: General 177 (1999) 111120 J.A. Nav

113

Scheme 1. Flow sheet of the preparation procedure for specimens studied in the present work.

infrared (IR) and X-ray photoelectron (XPS) spectroscopies. The evolution of water during heating of the samples was performed under vacuum (103 Torr) using TPD linked to continuous monitoring of the gas phase with a quadrupole mass spectrometer (Leybold-Heraeus GMBM, quadruvac Q100) in the scanning mode of operation (TPD-MS method); a heating rate of 58 min1 was used. The IR spectra

were taken on a Perkin-Elmer model 883 spectrophotometer, using KBr pellets and working in the transmittance mode. The XPS study was carried out on a Leybold-Heraeus LHS-10 spectrometer, working with a constant pass energy of 50 eV; Mg K radiation was used for excitation (h 1253.6 eV). A nal pressure of 109 Torr was always attained before XPS recording.

114

o et al. / Applied Catalysis A: General 177 (1999) 111120 J.A. Nav

Table 1 Some properties of selected oxide samples prepared by different techniques Type of oxide TiO2(Dg) TiO2(hp) 0.5% Fe/Ti(sg) 1% Fe/Ti(sg) 2% Fe/Ti(sg) 3% Fe/Ti(sg) 5% Fe/Ti(sg) 5% Fe/Ti(a) 5% Fe/Ti(n) a-Fe2O3(hp)
a b

Specific surface area (m2 g1) 49.0c 57.8 57.2 56.8 48.1 44.0 45.0 41.8c 29.2c 39.0

XAa 0.773 1 1 1 1 1 1 0.472 0.452

Particle size (mm)b 0.03d 1060 10100 10100 10100 10100 100200 110 5e180f

Molar fraction of anatase. Measured by SEM. c Taken from [6]. d Provided by the manufacturers. e Pure TiO2. f Aggregates containing Fe.

3. Results and discussion In Table 1 the main characteristics of the oxide samples are listed. Bulk and surface characterization of the mixed Fe/Ti(n) and Fe/Ti(a) samples can be

found elsewhere [46]. Some data from these papers have been taken for comparative purposes. Coprecipitated samples show BET surface areas similar to TiO2(Dg). The presence of different amounts of iron in the TiO2(sg) matrix does not lead to any considerable differences in the SBET values, similarly to that observed previously for the Fe/Ti(a) samples [6]. Only a slight decrease is observed with the increase in iron concentration. Fig. 1 depicts the XRD spectrum of the commercial TiO2(Dg) sample showing the presence of both anatase (A) and rutile (R) phases. The molar fraction of anatase to rutile (XA), calculated according to the method described elsewhere [4,13], was estimated as 0.773. In Fig. 2, the XRD patterns of selected samples, 5 wt% Fe/Ti(sg) and of the parent oxides, TiO2(hp) and a-Fe2O3(hp), are shown. The comparison of the XRD spectra of TiO2(Dg) and TiO2(hp) indicates that, within the detection limits of this technique, this last sample consists of anatase as the practically unique phase, although obviously, amorphous phases cannot be detected. Fig. 2 suggests a decrease of crystallinity in the 5 wt% Fe/Ti(sg) in comparison with the undoped (sg) sample, indicated by the increase in the width-at-half-maximum of the peaks. It also suggests that in the doped sample,

Fig. 1. XRD pattern of the TiO2 (Degussa P-25) specimen. Aanatase; Rrutile.

o et al. / Applied Catalysis A: General 177 (1999) 111120 J.A. Nav

115

Fig. 2. XRD patterns of selected oxide samples: Fe/TiO2(sg) (5 wt%) and the parent oxides TiO2(hp) and a-Fe2O3(hp) specimens. Aanatase.

neither rutile nor a-Fe2O3 phases are present. Regarding previously reported structural features of Fe/TiO2 powders prepared by wet impregnation [46], both types of oxides (Fe/Ti(n) and Fe/Ti(a)) present practically similar A/R ratios (see Table 1), suggesting that the TiO2 P-25 structure of the precursor is preserved. In addition, the 5 wt% Fe sample showed X-ray peaks assigned to hematite and/or pseudobrookite (Fe2TiO5) phases. As all Fe/Ti(sg) samples have been calcined at 773 K for 24 h, the presence of amorphous oxo/ hydroxo iron phases can be excluded. This was reinforced for the fact that the thermal decomposition in air of Fe(acac)3, at the same conditions, gives the wellknown hematite phase (shown by XRD results). As known, the solubility of Fe3 ions in rutile is smaller than in anatase (see [1] and references therein). Thus, as observed previously by us [46], impregnated Fe/ Ti(n) and Fe/Ti(a) samples, containing rutile, present iron oxides as segregated phases. However, in the case of Fe/Ti(sg) samples, in which only anatase is present,

Fe3 should be in solid solution even for concentrations as high as 5 wt% (Fig. 2). Fig. 3 shows the micrographs obtained for TiO2(hp). As it can be seen, this sample consists of particles with a very large distribution of shape and dimensions (1060 mm); it can be even observed of particles smaller than 1 mm. A detailed SEM examination of the particle surfaces (Fig. 3(b)) shows that the grains are constituted by agglomerates of very small and round particles (1 mm), leading to a high degree of porosity. In contrast, TiO2(Dg) consists of very homogeneous and regular particles of polyhedral shape. The micrographs of Fe/Ti(sg) oxides (not shown) indicate that these samples are formed by particles of very irregular shape and dimensions, independently of the iron content. Most of the particles have straight edges and sharp corners, with a dimension within 10100 mm; the more concentrated sample (5 wt% Fe) exhibits an average size around 100200 mm. A detailed SEM examination of the

116

o et al. / Applied Catalysis A: General 177 (1999) 111120 J.A. Nav

Fig. 3. SEM micrographs of the TiO2(hp) sample: (a) general field showing the particles (shape and dimensions); (b) detailed examination of the particle surface.

surface of Fe/Ti(sg) samples (Fig. 4) shows smooth surfaces with deposits. The more diluted samples (less than 3 wt% Fe) present deposits irregular in shape and dimensions that seem to be more or less free. In contrast, the more concentrated ones (e.g. 5 wt% Fe) show deposits very homogeneous in shape and dimensions that appear to be incrustated at the surface. EDX analysis of the 5 wt% Fe/Ti(sg) sample showed iron uniformly distributed between particles and within one particle. This result represents another difference between the new samples and the impregnated ones, in which loaded samples (>2 wt%) showed a-Fe2O3 as a separate phase [46]. The homogeneous distribution of iron into the particles agrees with the XRD results, suggesting again the absence of a-Fe2O3 phases. The TPD proles of water desorbing from TiO2(hp) and Fe/Ti(sg) oxides are shown in Fig. 5. Below

Fig. 4. Detailed SEM micrographs of the surface particle for the Fe/TiO2(sg) specimens containing different amounts of iron: (a) 0.5; (b) 3; (c) 5 wt%.

773 K, the TPD proles show peaks associated to molecular water desorption at about 383, 523 and 623673 K, which reect different types of adsorbed molecular water. The peak at 383 K was ascribed to the desorption of physically adsorbed molecular water, whereas those at 523 and at 623673 K were associated with chemically dissociated water leading to Brnsted basic (OH) and acidic (O2. . .H) hydroxyl groups, respectively [1416]. The results

o et al. / Applied Catalysis A: General 177 (1999) 111120 J.A. Nav

117

Fig. 5. TPD profiles of TiO2(hp) and Fe/TiO2(sg) samples containing the indicated amount of iron.

indicate that these samples are somewhat hydrated hydroxylated. From the shapes of TPD proles of the Fe/Ti(sg) samples, it can be concluded that all of them are rather similar and independent of the amount of iron. By comparing the TPD proles of these samples with those reported for the impregnated ones [4], it can be noted that the water desorption peak found at ca. 523 K in the Fe/Ti(sg) samples was practically absent in the former (with the exception of the 0.5 wt% Fe/ Ti(n) sample). From this, it can be concluded that in

solgel samples a larger amount of basic hydroxyl groups are present (see later). The infrared spectra of the solgel specimens show a broad band in the region 36002600 cm1 (OH stretching) with a maximum centered about 3440 cm1 and a shoulder at 3260 cm1 (Fig. 6(A)), together with a band at 1630 cm1 (HOH bending, not shown). The band centered at 3440 cm1 can be ascribed to basic hydroxyl groups [14], whereas the band at 1630 cm1 corresponds to adsorbed molecular water. These qualitative results, similar for all the sol gel samples, indicate again that they are hydrated/ hydroxylated. It is important to remember that IR spectra of Fe/Ti(n) samples show no characteristic peaks of basic hydroxyl groups [4], and that Fe/Ti(a) samples show surfaces almost dehydrated/dehydroxylated [6]. Fig. 6(B) shows the 1000300 cm1 IR absorption region for TiO2(Dg) and Fe/Ti(sg) samples. All the spectra show peaks at 810, 680, 620, 525, 460, 425 and 380 cm1. Articial anatase and rutile have a large broad band in the 850530 cm1 region that masks their characteristic bands at 810 (anatase) and 603 cm1 (rutile) [17]; additional bands at 425 and 525 cm1 are also present in both TiO2 forms, and therefore cannot be unequivocally assigned. The assignment of the band at 460 cm1 requires greater attention because it can be present in anatase, rutile and brookite phases [17]; in addition, bands at 450 460 cm1 can be assigned to a-Fe2O3 [4]. As the characteristic bands of brookite (1165, 950, 850 and 585 cm1 [17]) are not present, and rutile and a-Fe2O3 phases are not observed in the XRD patterns, it can be concluded that the band at 460 cm1 could be assigned only to anatase. In summary, Fe/Ti (sg) samples present bands associated to hydroxyl groups and to the anatase phase, and do not show bands corresponding to iron phases.

Table 2 Elemental surface composition and XPS binding energies of the elements present at the surface of TiO2(hp) and Fe/Ti(sg) specimens Sample Surface concentration (at %); in parenthesis, energy binding (eV) Ti(2p) TiO2(hp) 0.5% Fe/Ti 3% Fe/Ti 5% Fe/Ti
a

O(1s) (458.5) (459.2) (458.1) (459.1) 68.23 68.09 63.80 65.42 (530.0) (530.6) (529.5) (530.6)

Fe(2p) Not detected 1.07/2.59a (711.1) 1.96/4.54a (712.4)

C(1s) 4.30 1.86 6.95 6.39 (284.6) (284.6) (284.6) (284.6)

27.46 30.02 28.18 26.21

Calculated surface concentration (wt%).

118

o et al. / Applied Catalysis A: General 177 (1999) 111120 J.A. Nav

Fig. 6. IR spectra in the region (A) 36002600 and (B) 1000400 cm1 of the indicated samples.

Table 2 summarizes the information about elemental surface composition and XPS binding energies (in parenthesis) of the elements present at the surface of TiO2(hp) and Fe/Ti(sg) specimens. As it can be seen, all the samples show typical binding energies at about 458.5 and 530 eV, characteristic of the Ti(2p) and O(1s) peaks, respectively [18]. In addition, Fe/Ti(sg) samples show peaks around 711 eV, typical of Fe(2p) [18]. Additional peaks at around 284.6 eV are found in all these specimens, corresponding to carbon impurities, arising probably from the thermal decomposition of the precursors during the ring process. It is noteworthy that in all these samples, the O(1s)/Ti(2p) ratios are greater than 2.0, indicating that the surfaces of the samples are hydrated/hydroxylated, conrming the ndings of TPD studies. The amounts of iron calculated from these XPS data are lower than the nominal ones. In contrast, XPS results of Fe/Ti samples prepared by impregnation are consistent with the formation of iron-rich surface phases [19]. These results clearly indicate that a well-dispersed dilute solid solution of Fe3 in TiO2 can be obtained by the solgel procedure described above, even for the more concentrated specimens. In the case of the

0.5 wt% Fe/Ti(sg) sample, iron is not detected by XPS or by EDX. Although this amount of iron cannot be seen by EDX, because it is under the limit of detection of this method, it should be detected by XPS. The only reasonable tentative explanation is that, in this sample, an external surface titania rich layer is formed. In fact, as observed by XPS, even the more loaded samples show an iron content at the surface lower than the nominal one. These particles would present, therefore, special features, being a mixed Fe/ TiO2 system with a surface enrichment in TiO2. DRS of Fe/Ti(sg) samples in comparison with TiO2(hp) and a-Fe2O3(hp) are depicted in Fig. 7. Only qualitative data in absorption are shown because the spectra have been shifted along the ordinate axis for the sake of comparison. The spectra of the 0.5, 1 and 2 wt% (sg) samples are very similar, whereas the remaining ones are increasingly similar to that of hematite. The spectral characteristics and assignation of bands in Fe/Ti(n) and Fe/Ti(a) samples have been reported in our previous papers [1,5,6]. In the visible range, the new doped samples show higher absorptions than the undoped TiO2(hp), and this absorption increases with the iron content, together with changes

o et al. / Applied Catalysis A: General 177 (1999) 111120 J.A. Nav

119

Fig. 7. DRS of TiO2(hp); a-Fe2O3(hp) and Fe/TiO2(sg) samples containing the indicated amount of iron.

on color from pale yellow to reddish-brown. Same changes in color have been reported in the case of Fe/ Ti(n) and Fe/Ti(a) specimens compared with their precursor TiO2(Dg) [5,6]. However, impregnated samples are more colored than those prepared by the sol gel technique, due to the existence of segregated iron phases in the former. On the other hand, the broad band centered at ca. 500 nm that appeared in Fe/Ti(a) samples [6] (assigned in part to surface states [1]) is less remarkable in the solgel samples. This is another
Table 3 Summary of the main properties found for the oxide particles Property Sample TiO2(Dg) Surface area (m2 g1) Particle size (mm) Phases Iron distribution homogeneity Iron-rich surface phases Surface hydroxylation
a b

evidence of the enhanced homogeneity of iron into the TiO2 bulk. A recent paper [20] reports features about the characterization of irontitania mixed oxides prepared by a solgel method starting from Ti(OnBut)4 and Fe(acac)3, with a coprecipitation technique similar to that followed in this work. From this report, it was concluded that the undoped sample possesses the anatase structure and that this phase is preserved in Fe-containing samples, similarly to what occurs in the present case. Therefore, it can be concluded that the formation or not of a unique phase seems to be inherent to the preparation technique, and does not depend on the nature of the precursor. A previous EXAFS analysis [21] of irontitania catalysts (prepared similarly by a solgel processing from Ti(OiBut)4 and Fe(acac)3) showed that for concentrations up to 5.49 wt%, Fe-ions isomorphously substitute Ti-ions of anatase in the surface plane of the lattice (1 0 0), leading to a certain deformation. Only at a higher loading, amorphous and multiphase products can be observed. These observations conrm our previous assumption that amorphous phases seem to be absent. Table 3 summarizes the main properties found for the new samples (Fe/Ti(sg)) compared with TiO2(Dg), Fe/Ti(n) and Fe/Ti(a). From the comparison, and according to the conclusions reported in [1], it can be said that Fe/Ti(sg) samples, although having a larger particle size than that of TiO2(Dg), seem to exhibit some suitable photocatalytic features: (a) only the anatase phase is present; (b) they present a homogenous distribution of iron, with no separated phases of iron oxides or pseudobrookite; (c) their surface area

Fe/Ti(n)a 29 Large (180) ARc No High Low

Fe/Ti(a)b 4147 Large (110) ARc No High Low

Fe/Ti(sg) 4557 Large (10200) A Yes Absent Intermediate

50 Small (0.03) A>R High

Taken from [4]. Taken from [6]. c For iron content>1 wt% separated phases of a-Fe2O3 and/or Fe2TiO5 are observed. Aanatase; Rrutile.

120

o et al. / Applied Catalysis A: General 177 (1999) 111120 J.A. Nav

is similar to that of TiO2(Dg); and (d) they possess a relatively higher amount of basic hydroxyl groups (which are crucial for photocatalysis [1416]) compared with the samples prepared by wet impregnation. In Part II of this work [11] we examine the activity of these samples in oxidative and reductive photocatalytic reactions, in comparison with doped samples prepared by the impregnation method. In spite of the promissory suitable properties, the efciency as photocatalysts was found lower than those of Fe/ Ti(n) and Fe/Ti(a) samples; the lack of the P-25 structure and the presence of defects arising from the preparation method could account for the reduced activity. 4. Conclusions Fe/Ti oxide samples have been prepared by a sol gel method using TiCl4 and Fe(III) acetylacetonate as precursors, providing a monophase material (only anatase), with no separated phases of iron oxides in the more loaded samples. A continuous distribution of iron into the TiO2 particle and between different particles was also obtained, together with an external surface titania rich layer. Iron-doped titania samples prepared by the solgel technique have a higher surface hydroxyl density in comparison with similar samples prepared by impregnation, and a specic surface area similar to that of TiO2 (Degussa P-25). In principle, all these properties make the new specimens to be expected to possess a good photocatalytic activity. In Part II [11] we report the results of our photocatalytic studies. Acknowledgements n General de InvesJAN wishes to thank ``Direccio n Cient ca y Te cnica (DGICYT-Spain), PB93tigacio 0917 and PB96-1346'' for supporting part of this work. Work performed as part of CNEA-CAC-UAQ PROJECT no. 95-Q-03-05. We also gratefully acknowledge CSIC (Spain) and CONICET (Argenn tina) for an interchange grant. MIL thanks ``Fundacio

Antorchas'' (Argentina) for a travel grant to Sevilla. MIL is a member from CONICET. References
o, J. Photochem. Photobiol. A 98 (1996) [1] M.I. Litter, J.A. Nav 171. [2] N. Serpone, D. Lawless, Langmuir 10 (1994) 643. [3] W. Choi, A. Termin, M.R. Hoffmann, J. Phys. Chem. 98 (1994) 13669. o, M. Mac as, M. Gonza lez-Catala n, A.J. Justo, [4] J.A. Nav Mater. Sci. 27 (1992) 3036. o, J. Photochem. Photobiol. A 84 (1994) [5] M.I. Litter, J.A. Nav 183. o, G. Colo n, M.I. Litter, G.N. Bianco, J. Mol. Catal. [6] J.A. Nav 106 (1996) 267. nech, J.A. Nav o, J. Mol. Catal. 87 [7] A. Milis, J. Peral, X. Dome (1994) 67. o, G. Colo n, M. Trillas, J. Peral, X. Dome nech, J.J. [8] J.A. Nav n, D. Rodr guez, M.I. Litter, Appl. Catal. B 16 Testa, J. Padro (1988) 187. o, in: G. Grassi, D.O. Hall (Eds.), [9] R.I. Bickley, J.A. Nav Photocatalytic Production of Energy-Rich Compounds, Elsevier, London, 1985, p. 105. [10] L. Palmisano, V. Augugliaro, A. Sclafani, M. Schiavello, J. Phys. Chem. 95 (1988) 6710. o, J.J. Testa, P. Djedjeian, J.R. Padro n, D. [11] J.A. Nav guez, M.I. Litter, Appl. Catal. A, in press. Rodr n, J.A. Nav o, M.I. Litter, J. Adv. Oxid. [12] E.A. San Roma Technol., in press. [13] J.M. Criado, C. Real, J. Chem. Soc., Faraday Trans. 1 79 (1983) 2765. [14] G. Munuera, V. Rives-Arnau, A. Saucedo, J. Chem. Soc., Faraday Trans. 1 75 (1979) 736. lez-Elipe, G. Munuera, J. Soria, J. Chem. Soc., [15] A.R. Gonza Faraday Trans. 1 75 (1979) 748. lez-Elipe, J. Soria, J. Sanz, J. Chem. [16] G. Munuera, A.R. Gonza Soc., Faraday Trans. 1 76 (1980) 1535. [17] H.W. Van Der Marell, H. Beutelspacher, Atlas of Infrared Spectroscopy of Minerals and their Admixtures, Elsevier, Amsterdam, 1976, p. 257. [18] Handbook of X-Ray Photoelectron Spectroscopy, PerkinElmer Corp., Published by Physical Electronics Division, USA, 1979. lez Carren lez-Elipe, G. o, A.R. Gonza [19] R.I. Bickley, T. Gonza Munuera, L. Palmisano, J. Chem. Soc., Faraday Trans. 90 (1994) 2257. [20] Yu.V. Maksimov, I.P. Suzdalev, M.V. Tsodikov, V.Ya. Kugel, o, J. Mol. Catal. A O.V. Buthtenko, E.V. Slivinsky, J.A. Nav 105 (1996) 167. [21] M.V. Tsodikov, O.V. Buthtenko, O.G. Ellert, V.M. Shcherbakov, D.I. Kochubey, J. Mater. Sci. 30 (1995) 1087.

Vous aimerez peut-être aussi