You are on page 1of 9

Colloids and Surfaces A: Physicochem. Eng.

Aspects 260 (2005) 199–207

EPR study of ceria–silica and ceria–alumina catalysts: Localization


of superoxide radical anions
Antoine Aboukais a , Elena A. Zhilinskaya a , Jean-François Lamonier a , Igor N. Filimonov b,∗
aLaboratoire de Catalyse et Environnement, E.A. 2598, Université du Littoral Côte d’Opale, MREID,
145 Avenue Maurice Schumann, 59140 Dunkerque, France
b Chemistry Department, Lomonosov Moscow State University, Vorob’evi Gori, Moscow 119992, Russia

Received 8 April 2004; received in revised form 18 January 2005; accepted 25 February 2005

Abstract

Oxygen adsorption experiments were performed on evacuated and prereduced CeO2 /SiO2 and CeO2 /Al2 O3 catalysts with and without
platinum. Considerable amounts of the superoxide radical ions were stabilized on all the samples. Signal parameters suggest Ce4+ –O2 −
positioning for all detectable superoxide species. Physisorbed oxygen broadens O2 − signal beyond detection for all the alumina-based
samples, while the same procedure for all the silica-based samples did not change signal shape of O2 − species. Detectable O2 − species are
localized in the bulk of ceria and the nature of support (silica or alumina) determines the number of oxygen vacancies and the rate of electron
transfer. XRD data suggest that for alumina-based samples small and/or thin islands of ceria dominate, while comparatively large ceria particles
are stabilized on the surface of silica-based samples with the same ceria content. Average size of ceria crystallites is still not determining
factor and cannot account for the observed differences. Higher concentrations of paramagnetic species may be stabilized on alumina-based
samples and thus, sensor-like behavior towards gaseous oxygen at room temperature was detected for them—oxygen admission reversibly
changes superoxide lineshape. For silica samples, only minor changes of O2 − lineshapes were typical upon the change of the partial pressure
of oxygen at ambient and low temperatures. Addition of platinum has little effect on parameters of the O2 − signal, except an enhancement of
the superoxide decay in the reducing media. Possible site for O2 − stabilization inside the lattice of CeO2 was proposed and relevance of the
observed effects to the redox catalysis discussed.
© 2005 Elsevier B.V. All rights reserved.

Keywords: Ceria–silica; Ceria–alumina; EPR; Superoxide radical anions; Oxygen vacancies

1. Introduction [3], maintaining oxygen balance in such a way that provides


good performance both in fuel-rich and fuel-poor conditions.
Ceria-containing catalysts are used in a number of indus- It is well established that the activity of catalysts based
trial applications [1], probably the most widespread being on oxides is related to the formation of oxygen vacancies.
an additive to automotive three-way catalysts (TWCs) [2]. The latter is true for ceria as was thoroughly discussed in the
In TWCs, ceria has a number of functions, one of the most recent review by Gellings and Bouwmeester [4]. Reduction
important being oxygen storage capacity (OSC). OSC arises increases the number of oxygen vacancies and thus creates
due to the involvement of ceria lattice oxygen in redox cycle additional sites for oxygen adsorption. Chemisorbed oxygen
tend to withdraw electrons from any available donor on a
basis of generally accepted transformations:
∗ Corresponding author at: LG Chem Ltd./Research Park, 104-1 Moonji-

dong, Yuseong-gu, Daejeon 305-380, Republic of Korea.


Tel.: +82 42 866 4729; fax: +82 42 863 7466. O2 (ads) → O2 − (ads) → O2 (ads) → 2O−
E-mail addresses: filimonov@lgchem.com,
INFilimonov@phys.chem.msu.ru (I.N. Filimonov). (ads) → 2O2− (lattice)

0927-7757/$ – see front matter © 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.colsurfa.2005.02.036
200 A. Aboukais et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 260 (2005) 199–207

Electrophilic species formed during oxygen incorporation XRD data for structural analysis were collected at room
into the lattice may be active centers in oxidative reactions. temperature with a Bruker D8 advance diffractometer using
Superoxide radical anions O2 − stabilized on the surface of Cu K␣ radiation (λ = 1.5418 Å) by placing samples in quartz
ceria and ceria-containing catalysts were thoroughly studied sample holders for 2θ = 10–80◦ with a step size of 0.02◦ and
the last decade [5–14] a step time of 32 s. The XRD patterns, after background cor-
Very interesting as well important feature of noble metal rection and K␣2 stripping, were assigned using the JCPDS
containing ceria catalysts is their improved activity in the ox- (Joint Committee on Powder Diffraction Standards) database.
idative reactions at low temperatures after reductive pretreat- TOPAS P version 2.0 software was used for the evaluation of
ment by any available reductants: CO, H2 and hydrocarbons the average size of ceria crystallites.
[15–19]. The effect was attributed to the creation of certain The electron paramagnetic resonance (EPR) measure-
platinum sites close to the ceria interface [19], or creation ments are performed at 25 and −196 ◦ C on an EMX Bruker
of active sites on ceria due to the SMSI-like effect caused spectrometer. A cavity operating with a frequency of 9.3 GHz
by partial decoration of platinum particles with ceria [17]. (X-band) was used with a modulation at 100 kHz. Precise g-
Holmgren et al. [20] invoked serious arguments against such values are determined from simultaneous precise frequency
interpretation since there are substantial doubts about the pos- and magnetic field values. For all the cases of rhombic or
sibility of metal particles decoration at low reduction tem- axial signals, g-values were determined using computer sim-
peratures as 370–470 K. In the latter work, isotopic labeling ulation.
was used to discriminate between the reaction pathways for The following standard treatment sequence was adopted
carbon monoxide oxidation. It appeared that oxygen for CO for all the samples (Fig. 1). Evacuation at 600 ◦ C was used
oxidation was delivered mainly from the catalyst even for the to achieve “clean” state of ceria surface, hydrogen treat-
prereduced sample. The latter seems rather strange since pre- ment at room temperature and 100 ◦ C was used to simulate
reduced sample must be stripped off any reduceable species. the reduced state of the surface. When oxygen admission
Therefore, the nature of an enhanced activity of Pt/ceria cat- was performed at room temperature, the EPR spectra were
alysts is somewhat controversial. The only common expla- taken both at room and liquid nitrogen temperature. The next
nation of the cause of such effect is a creation of certain yet step—evacuation at room temperature—was performed un-
unidentified sites in the close vicinity to platinum–ceria inter- til the background pressure of the vacuum system (<10−4 Pa)
face. Another related and as well strange effect reported for was achieved. At this step, the EPR spectra were also taken
ceria-based catalysts is their pronounced deactivation after at room and liquid nitrogen temperature. Before reduction in
short exposure to oxygen at room temperature [20]. hydrogen, the samples were evacuated for about 1 h. About
Understanding of these phenomena probably may be 90 kPa of hydrogen was admitted and the sample was exposed
achieved by the investigation of the formation and decay to hydrogen for 1 h. After cooling down to room temperature,
transient oxygen species during the oxidation and reduction the sample was evacuated and again the EPR spectra were
stages. Superoxide anion radicals are possible intermediates also taken at room and liquid nitrogen temperature. In order
during oxidation of ceria and at the same time are valuable to perform oxygen adsorption at 77 K, the samples after evac-
probe molecules for study of oxygen vacancies in a variety of uation at 600 ◦ C were cooled down in liquid nitrogen under
oxides [21]. In the present work, we followed the mode of for- evacuation for at least 40 min. Then, about 80 ␮mol of O2
mation of superoxide species on prereduced ceria-containing were dosed onto the sample and the sample was held at 77 K
catalysts. By variation of the extent of prereduction, we tried for 30 min. Then, an excess of oxygen was removed by evac-
to find out the differences of the superoxide localization in uation at 77 K until the background pressure was achieved.
alumina- and silica-based ceria-containing catalysts. The samples were held at 77 K for additional 24 h (during
this time, the EPR spectra were taken regularly) and then
warmed up to room temperature for 30 min. After this, the
2. Experimental spectra were taken at room and liquid nitrogen temperature.

Alumina (Rhone-Poulenc) and silica (Aerosil-300) from


Degussa were used as received. 3. Results
Ceria-containing samples were prepared by the incipient
wetness technique using aqueous Ce(NO3 )3 ·6H2 O solution Clean samples, i.e., evacuated at 600 ◦ C showed existence
of an appropriate concentration. After overnight drying at Fe3+ isolated ions (according to intense signals at g = 4.3) and
120 ◦ C, the samples were calcined under the flow of dry air isolated Mn2+ , typical for alumina-based samples (according
at 600 ◦ C. Platinum was supported on thus prepared ceria- to the six lines centered at g = 2, IMn = 5/2) clearly seen in
containing samples using chloroplatinic acid by the incipient Fig. 2. Alumina-based samples showed broad and asymmet-
wetness technique. Platinum-loaded samples were dried at ric signal at g ∼ 1.99. Considerable shift of g-value of the
120 ◦ C overnight, calcined under the flow of dry air at 500 ◦ C latter signal from ge suggests involvement of transition metal
and reduced under hydrogen flow for 3 h at 300 ◦ C. Data on ions. Assuming the presence of ceria in all studied samples the
chemical analysis and BET analysis are reflected in Table 1. most probable candidates are electronic defects or impurities
A. Aboukais et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 260 (2005) 199–207 201

Table 1
Textural properties of the samples
Sample (abbreviation used in the text) Composition BET surface area (m2 g−1 ) Mean pore Mean pore volume (cm3 g−1 )
diameter (Å)
CeO2 /SiO2 (Ce–Si) CeO2 : 3.9 wt.% 288 118 0.9
Pt/CeO2 /SiO2 (Pt–Ce–Si) Pt: 0.2 wt.%; Ce: 3.3 wt.% 261 125 0.8
CeO2 /Al2 O3 (Ce–Al) CeO2 : 2.9 wt.% 204 84 0.5
CeO2 /Al2 O3 (10Ce–Al) CeO2 : 10 wt.% 200 70 0.4
Pt/CeO2 /Al2 O3 (Pt–Ce–Al) Pt: 0.3 wt.%; CeO2 : 3.3 wt.% 189 85 0.5

related to ceria [22,23]. Silica-based samples produced nar- (Fig. 3b) are typical to superoxide radical ions, stabilized
row rhombic signals (g1 = 2.044; g2 = 2.011 and g3 = 2.0036) exclusively on ceria. These signals were stable for several
(Fig. 2c and d). Taking into account solely signal shape and hours and considerably diminished after, e.g., overnight pe-
g-values, the latter signals might be attributed to silica de- riod. Same signals were detected after oxygen admission at
fects, e.g., Si O O• [24]. Due to the stability of all these room temperature onto the reduced at 298 and 378 K sam-
signals, we could not obtain any additional information to ples. After the deeper prereduction of alumina-based samples
make more substantial assignments. All the signals of evac- at 400 ◦ C, the signal changed slightly into less anisotropic one
uated at 600 ◦ C samples will be termed further as Bkg-type with gz = 2.028, gy = 2.018 and gx = 2.012.
signals. The centers producing them were relatively inert and The lineshape of superoxide signals formed at room tem-
were not involved into the redox chemistry discussed further. perature on both silica- and alumina-based samples practi-
Admission of oxygen at room temperature onto evacu- cally was not dependent upon temperature contrary to the
ated at 600 ◦ C samples with a subsequent short evacuation typical behavior of superoxide species localized on the sur-
leads to the formation of intense signals both on alumina- face [25]. Therefore, superoxide ions detected here probably
and silica-based samples. Their characteristic parameters ac- do not possess rotational degrees of freedom and thus should
cording to computer simulation with gz = 2.031, gy = 2.0155 be localized in a rather confined space, i.e., in the bulk. For
and gx = 2.0125 for alumina-based samples (Fig. 3a) and pure ceria, both rotating and stable superoxide species were
gz = 2.031, gy = 2.018 and gx = 2.012 for silica-based samples detected [26].

Fig. 1. Standard sequence of treatments (flowchart in the middle); bar charts: intensity of superoxide signals for the corresponding step of the flowchart. Left
bar chart: Ce–Si sample; right bar chart: Ce–Al sample.
202 A. Aboukais et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 260 (2005) 199–207

Fig. 3. EPR signals taken at 77 K of alumina-based (a) and silica-based


(b) samples after oxygen adsorption and evacuation at room temperature.
Bkg-type signals were subtracted for the silica-based sample.

Superoxide-type signals on silica-based samples were


much less affected by the dipolar broadening. Admission of
80 or 320 ␮mol of oxygen onto Ce–Si and Pt–Ce–Si samples
did not changed signal shape. Only minor changes of the line-
shape (Fig. 4b) suggest weak interaction with physisorbed
oxygen. Therefore, in case of silica-based catalysts, super-
oxide radical ions are more in the “bulk state” compared to
alumina-based samples.
Moreover, the introduction of gaseous oxygen at room
temperature may change the shape of the superoxide-type sig-
nals for the alumina-based samples, while only slight changes
were found in case of silica-based samples (Fig. 5). O2 − sig-
nal on alumina-based samples broadens substantially, while
O2 − signal shape is preserved for the silica-based samples
during the same treatments. Therefore, sensor-like behavior
Fig. 2. EPR signals taken at 77 K of evacuated at 600 ◦ C samples. Upper
part: (a) Pt–Ce–Al; (b) Ce–Al; (c) Pt–Ce–Si; (d) Ce–Si. Lower part: exploded
view of silica defect signal.

Despite a close similarity between the signals produced


by superoxide radical ions on alumina- and silica-based cat-
alysts, their behavior towards dipolar broadening by the ph-
ysisorbed oxygen was quite different. Admission of excess
quantities (320 ␮mol) of oxygen at 77 K was enough for the
complete disappearance of superoxide signals on Ce–Al and
Pt–Ce–Al samples, while only slight change of the lineshape
was noticeable for all the silica-based samples.
Dipolar broadening experiments performed for Ce–Al
samples showed variations of the superoxide signal lineshape
upon admission of smaller (less than 320 ␮mol) quantities
of oxygen. Thus, for Ce–Al sample, admission of 80 ␮mol
of oxygen at 77 K led to the considerable broadening of
superoxide-type signal with its transformation into a sym-
metric lineshape (Fig. 4a) with giso = 2.0204. Integration of
this line showed considerable enhancement of the signal in- Fig. 4. EPR signals taken at 77 K after admission of: (a) 80 ␮mol of oxygen
tensity (Fig. 1 and Table 2). upon Ce–Al samples; (b) 320 ␮mol of oxygen upon Ce–Si samples.
A. Aboukais et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 260 (2005) 199–207 203

Table 2
Results of double integration of the EPR signals of superoxide radical anions
during consecutive treatments
Treatment Signal intensity for Signal intensity for
Ce–Si sample (a.u.) Ce–Al sample (a.u.)
1 . Evacuation at 0 0
600 ◦ C

2 . O2 adsorption at 0 216.2
77 K for 20 h
3 . Warming to 25 ◦ C 1.0 3.4
1. Evacuation at 0 0
600 ◦ C
2. O2 adsorption at 4.9 26.9
25 ◦ C, 80 ␮mol
3. Evacuation at 25 ◦ C 5.1 2.2
4. O2 adsorption at 5.3 0
25 ◦ C, 320 ␮mol
5. H2 reduction at 4.6 0.8
25 ◦ C
6. O2 adsorption at 7.0 8.0
25 ◦ C, 80 ␮mol Fig. 6. EPR signals taken at 77 K after admission of 80 ␮mol of oxygen
7. Evacuation at 25 ◦ C 4.4 1.1 at 77 K upon evacuated Ce–Al sample. Excess of oxygen was removed by
8. O2 adsorption at 5.1 0 evacuation at 77 K after 1 h of exposure to oxygen and the sample was held at
25 ◦ C, 320 ␮mol 77 K for 24 h. (a) Solid line: experimental spectrum; dashed line: simulated
9. H2 reduction at 0 0 spectrum. (b and c) Two computer-simulated signals used for simulation of
100 ◦ C experimental spectrum.
10. O2 adsorption at 4.5 22.3
25 ◦ C, 80 ␮mol Experiments on adsorption of oxygen at 77 K permitted
7. Evacuation at 25 ◦ C 3.7 1.8 us to find out even more striking difference between silica-
11. O2 adsorption at 4.2 0 and alumina-based samples.1 Unlike the case of oxygen
25 ◦ C, 320 ␮mol
adsorption at room temperature onto evacuated samples,
All the signals were taken at 77 K.
admission of oxygen onto the cooled to 77 K evacuated
silica- and alumina-based samples did not produce any
towards gaseous oxygen of alumina-based samples relates
new signals immediately after oxygen admission, or after
to the formation of superoxide species on the surface or not
2 h after oxygen admission. Only after 20 h of exposure
far from the surface. Exposure to 80 ␮mol of oxygen at 77 K
at 77 K, very strong signals appeared on Ce–Al sample
leads to a 10-fold increase of O2 − intensity in case of alumina-
(Fig. 6) that will be designated hereafter as low-temperature
based samples (Fig. 1 and Table 2). Subsequent evacuation
signals. But still no signals were detected for Ce–Si sample.
decreases the intensity of O2 − signal.
Low-temperature signal in Fig. 6 is almost identical to the
signals of superoxide radical anions obtained under similar
conditions on alumina-supported ceria in the paper by
Soria et al. [10]. It was impossible to simulate experimental
spectrum of Fig. 6 using only one signal, thus same approach
of superpositioning two independent signals as was used
by Soria et al. [10] was employed. Two superoxide signals
with gz = 2.0258, gy = 2.0223, gx = 2.0107 and gz = 2.0258,
gy = 2.0169, gx = 2.011 (that corresponds to signals II and
III in Soria notation [10]) were used for satisfactory signal
simulation (Fig. 6). Intensity of the low-temperature signals
thus obtained was one to two orders of magnitude higher,
than intensity of superoxide signals obtained by oxygen
adsorption at room temperature (Fig. 1 and Table 2). Upon
heating of the sample exhibiting low-temperature superoxide
signals, the latter were transformed into signals typical to
the adsorption of oxygen at room temperature (Fig. 3a) with
corresponding decline of signal intensity (Fig. 1 and Table 2).

Fig. 5. EPR signals taken at 25 ◦ C of silica- and alumina-based samples after


admission of 80 ␮mol of oxygen upon: (a) Pt–Ce–Si sample; (b) Pt–Ce–Al 1 We are indebted to the reviewer who has drawn our attention to this

sample. The asterisk indicates a Mn2+ signal. possibility.


204 A. Aboukais et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 260 (2005) 199–207

of the crystallite size gave the following average size of ceria


crystallites: for Ce–Si sample 8.3 nm and for Ce–Al samples
4.8 nm. Thus, alumina support provides better dispersion of
ceria. Platinum addition did not alter the dispersion of ceria
since Pt-doped and Pt-free samples provided identical XRD
patterns.
In order to evaluate the effects of mean crystallite size,
Ce–Al sample with 10 wt.% loading of CeO2 was prepared
(designated as 10Ce–Al). Mean diameter for the CeO2 par-
ticles in the case of the latter sample was closer to the Ce–Si
sample—about 7.1 nm. Nevertheless, superoxide species sta-
bilized on 10Ce–Al sample had same labile behavior as those
stabilized on Ce–Al sample and no features resembling the
behavior of the superoxide species of Ce–Si sample. There-
fore, mean size of CeO2 crystallites is not decisive parameter
that regulates the behavior of O2 − species.

4. Discussion

Fig. 7. XRD patterns for Ce–Si (a), 10Ce–Al (b) and Ce–Al (c) samples. EPR data on superoxide anion radicals, stabilized on ceria
and ceria-containing catalysts are quite numerous at present
Quite different behavior towards low-temperature adsorption [5–14,22,27–29]. Thus, on the basis of chemical evidence and
of oxygen was found for silica-based samples where oxygen g-tensor values, the attribution of the signal related to oxygen
adsorption at 77 K did not produce any signals within 24 h. adsorption is easy. Soria et al. attributed comparable signals
At the same time, warming thus treated (i.e., containing on alumina-based samples to Ce4+ –O2 − species adjacent to
preadsorbed at 77 K oxygen) sample up to room temperature isolated oxygen vacancies [9]. Same attribution seems to be
also led to the formation of the superoxide signals typical to applicable to superoxide signals here assuming similarity of
the adsorption of oxygen at room temperature (Fig. 3b). signals and mode of pretreatments.
The nature of support (silica and alumina in our case) plays Despite an ease of the attribution of superoxide signals,
a key role in the susceptibility of superoxide radical ions to it is less clear why superoxide ions are quite differently lo-
the interaction with physisorbed oxygen according to the pat- calized on ceria–silica and ceria–alumina samples. In accor-
terns of O2 − intensity for the standard sequence of treatments dance with the dipolar broadening experiments, O2 − radicals
(Fig. 1). The most inert to the environmental changes are the should be localized in the bulk for silica-based samples and
superoxide species of silica-based samples. Complete dis- on the surface for alumina-based samples. At the same time,
appearance of the signal occurs only upon the reduction in both of them are stabilized on Ce4+ ion, and due to the fact
hydrogen at 373 K. The most labile are the O2 − species on of a comparable g-tensor, they should have comparable lo-
alumina-based samples. The intensity of the signals in the calization site. Taking into account that dipolar interaction
latter case is affected by the composition of the gas phase. is effective up to ca. 20 Å and therefore, all the signals of
The signals may be either dipolar broadened beyond detec- paramagnetic species within the layer of ca. 20 Å thickness
tion by the physisorbed oxygen in the case of high oxygen should be broadened upon oxygen physisorption. Then, if ce-
pressure, or additional pool of radical anions may be detected ria patches are small (or thin enough), i.e., with a <20 Å thick-
for the low oxygen pressure according to the increase of sig- ness, then EPR signals of all paramagnetic species stabilized
nal intensity. Addition of platinum changes appreciably the within these small patches may be broadened upon oxygen
pattern of superoxide signal intensity: platinum-doped sam- physisorption. Therefore, the observed effects are possible
ples had less loosely bound oxygen species according to only only in the case when ceria particles on alumina are much
ca. 40% of intensity increase upon adsorption of 80 ␮mol of smaller (or much thinner) than on silica. According to XRD
oxygen. At the same time, platinum addition enhances super- results, ceria particles are indeed smaller on ceria–alumina
oxide decay in hydrogen at room temperature for both silica- than on ceria–silica with the same content of ceria and thus
and alumina-based samples. all the superoxide radical ions stabilized on alumina-based
XRD patterns of the samples are presented in Fig. 7: no- samples are more susceptible to broadening than on silica
ticeable peaks related to ceria (JCPDS no. 340394) may be samples. This is as well supported by UV–vis and XRD
found in case of Ce–Si samples, while for the Ce–Al samples comparative data of ceria–silica and ceria–alumina samples
these peaks are almost completely masked by the reflexes of [30]. Martı́nez-Arias et al. proposed existence of bidimen-
␥-alumina phase (JCPDS no. 742206). Nevertheless, using sional and tridimensional patches of ceria on alumina [14].
the strongest diffraction (1 1 1) line of CeO2 , the refinement In our case, due to the low ceria loading, mostly bidimen-
A. Aboukais et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 260 (2005) 199–207 205

sional patches are stabilized on Ce–Al sample. At the same sample, a decay of the signal during the warming may be at-
time, an attempt to increase mean size of CeO2 crystallites tributed to the second electron transfer and formation of EPR
on alumina-based sample by increasing the ceria content to silent species:
10 wt.% did not “shift” the behavior of superoxide species
O2 − (ads) + e → O2 (ads)
to the one typical for silica-based samples. Therefore, either
average diameter of particles found by XRD peak broad- Not only electron transfer rate is dissimilar for the ceria–silica
ening does not account for a real shape of the crystallites, and ceria–alumina samples, the highest concentration of su-
i.e., plate-like versus cube-like, or superoxide radical anions peroxide species observed for ceria–alumina is at least one
are preferentially localized on bidimensional ceria. Possible order of magnitude higher than the corresponding value for
explanation of the observed effects may be related to the ceria–silica samples. This reflects obviously the difference in
possibility of epitaxy formation and preferential exposure of the amount of oxygen vacancies. Since more oxygen vacan-
certain crystallographic planes of ceria on alumina-supported cies are available for ceria–alumina, then more electrons (or
samples already mentioned by Martı́nez-Arias et al. [14]. higher density of Ce3+ species) for dioxygen reduction are
There are another two striking dissimilarities of the super- available as well. The latter may account for higher transfer
oxide formation–decay for alumina- and silica-based sam- rate, i.e., in case of electron tunneling.
ples: (i) superoxide species on alumina-based samples are If superoxide species are indeed localized in the bulk and
susceptible to partial pressure of oxygen both at low and am- subsurface layers of ceria on supported catalysts then evi-
bient temperature; and (ii) slow but detectable rate of O2 − dent questions should follow: (i) how do dioxygen species
formation at 77 K found only in case of alumina-based sam- penetrate into the bulk at room temperature?; (ii) what is the
ples, while O2 − formation was suppressed for silica-based nature of possible sites for the localization of O2 − inside the
samples under the same conditions. ceria lattice?; and finally, (iii) what are the consequences of
Effects of the changes of lineshape and intensity of super- all this for the catalysis?
oxide signals during changes of partial pressure of oxygen In general, it is difficult to find realistic pathway for the
may be explained by two mechanisms. The simple mech- dioxygen transfer in neutral or ionized form through the ionic
anism is fast adsorption–desorption of loosely bound O2 − lattice at ambient or subambient conditions. Comparative
species: ease of oxygen vacancy formation and its efficient transport
within the lattice of ceria suggests that it must be monooxy-
O2 (gas) + e  O2 − (ads)
gen entities that may produce O2 − under favorable condi-
If the local density of O2 − species increases then the possi- tions inside the lattice. According to the XPS studies [31],
bility of spin–spin interaction increases as well leading to the reoxidation of the prereduced ceria is even more effective
corresponding changes of lineshape and signal intensity. in the bulk than on the surface, suggesting bulk enrichment
Another possible mechanism is an oxidation of EPR-silent with oxygen compared to surface. Another XPS study [32]
peroxide species during admission of gaseous oxygen with revealed a formation of specific oxygen species attributed to
formation of EPR-active superoxide species: O− , when CO + O2 mixture was added to the Pt/ceria cata-
lyst. Thus, provided the space permits, a recombination of
O2 + O2 (gas)  2O2 −
two monooxygen species (e.g., O− + O2− → O2 − + 2e) with
Same considerations for the above case predict signal broad- a corresponding electron transport to the nearest electron-
ening with increasing concentration of paramagnetic species. accepting site might provide fast O2 − stabilization inside the
Reverse reaction occurs during evacuation. Complete dis- lattice. At present level of knowledge, we cannot provide any
crimination among the mechanisms is probably impossible evidence for this pathway, yet there are seemingly no better
without monitoring peroxy species on the surface. alternatives for the explanation of experimental results.
Formation of superoxide radical anion requires one oxy- The next point is the site of the superoxide stabilization
gen vacancy and one electron to reduce adsorbed neutral in the ceria lattice and therefore, certain structural exami-
dioxygen. Both silica- and alumina-based samples are ca- nation should be performed. The unit cell of CeO2−x may
pable of stabilizing of superoxide species and therefore, both be viewed as eight edge-shared tetrahedra with Ce4+ cations
have sufficient number of oxygen vacancies. Thus, differ- in the corners and O2− anions in the center of each tetrahe-
ence in the results of superoxide formation at 77 K reflects dron if stoichiometric oxide is formed (Fig. 8a). Here, we
the difference in electron transport properties of the samples. would like to stress an existence of an empty octahedron in
Evidently, the activation energy of electron transfer is much the center of the unit cell of stoichiometric CeO2 (Fig. 8b).
lower in case of alumina-based samples. Thus, the rates of Oxygen vacancies, amount of the latter may reach up to 25%
formation (and obviously decay) of superoxide species are without a collapse of the structure, tend to form arrayed sub-
much faster on alumina-based samples. Since warming of structures, e.g., along
1 1 1 direction [33]. Neither isolated
the Ce–Al sample with high content of superoxide species oxygen vacancies, nor pairs of vacancies can afford stabi-
formed during adsorption at 77 K leads to the decrease (by lization of O2 − species for the geometric limitations. Central
two orders of magnitude) of the signal intensity, and tak- empty octahedron combined with any of oxygen vacancies
ing into consideration easier electron transfer for the latter provides more than enough space for O2 − stabilization (O O
206 A. Aboukais et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 260 (2005) 199–207

decay in reducing media in line with the earlier findings


[13]. This may occur via hydrogen spillover due to the well-
established properties of supported platinum to facilitate the
latter process [36]. On the other hand, platinum doping low-
ered the amount of loosely bound superoxide species prob-
ably due to considerable contamination of the surface with
chloride ions introduced during the impregnation step. In gen-
eral, we could not find any effect of platinum doping on signal
parameters of O2 − radicals, thus detectable superoxide radi-
Fig. 8. CeO2 fluorite unit cell. Larger spheres: Ce4+ cations; smaller spheres: cal anions are stabilized relatively far from noble metal.
O2− anions. (a) Face centered cubic array of Ce4+ cations tetrahedrally co-
ordinates O2− anions; lower empty tetrahedron depicts oxygen vacancy. (b)
Empty octahedron formed by Ce4+ ions positioned in the center of cube
faces.
5. Conclusions

distance in O2 is about 1.33–1.34 Å, unit cell parameter for
CeO2 is 5.41 Å, thus the distance between apical Ce4+ ions Ceria-containing alumina- and silica-based samples are
able to stabilize considerable amount of superoxide radical
in the empty octahedron is as well 5.41 Å). Adjacent oxygen
ions inside the lattice of ceria. The susceptibility of the shape
vacancies provide local distortion (due to the displacement
of superoxide signal on alumina-based samples on one hand
of Ce4+ ions) sufficient to diminish the symmetry from octa-
and relative inertness of superoxide anions of silica-based
hedral to rhombohedral. This distortion may account for the
samples to gaseous environment on another reflects the dif-
rhombic symmetry of the O2 − lineshapes.
ferent rate of electron transfer and the quantity of oxygen
Finally, it is necessary to comment on the relevance of
vacancies (both are higher for alumina-based samples). This
discovered effects to the catalysis. The first and evident
is the result of the difference in the interaction of a support
point—the ease of O2 − formation inside CeO2−x lattice
with the active phase, which is manifested by the smaller
at ambient conditions—reflects a very high rate of oxy-
ceria particles on alumina-samples at comparable ceria con-
gen/oxygen vacancy transport within the lattice. This is of
tent. At the same time, we could not detect a correlation of
importance from the point of view of Mars–van Krevelen type
the average size of ceria particles with the behavior of O2 − .
of redox catalysis. On the other hand, formation of compar-
atively inert O2 − species may account for the low activity of Ease of the formation of superoxide radical ions inside
oxidized ceria catalysts. Indeed, in order to supply the surface the lattice reflects considerable oxygen mobility at ambient
with the oxygen it is necessary at first to disrupt O O bonds conditions for the ceria. Fast formation of O2 − is followed by
in superoxide ions in the bulk and supply resulting oxygen its relatively slow decay at room temperature. The observed
species with electrons in order to avoid the recombination. effects may be important for the catalytic oxidation and re-
Thus, O2 − radicals might be somewhat dead-end species in lated to high activity of prereduced ceria catalysts and low
activity of oxidized ones.
redox processes on ceria and explain the results of lower ac-
Genetic octahedral voids of CeO2 fluorite structure with
tivity of completely oxidized ceria catalysts compared to the
prereduced ones [20]. adjacent oxygen vacancies are probable sites for O2 − stabi-
The effect of fast transport of oxygen into the lattice in lization inside the lattice.
conjunction with the data of Holgado et al. [31] on more effec- Platinum addition does not change appreciably properties
tive bulk oxidation of ceria (compared to surface oxidation) of superoxide ion radicals, except higher rate of their decay
may explain the results of isotope tracing by Holmgren et al. in reducing media.
[20] of CO + O2 reaction on Pt/CeO2 catalysts. According to
the latter study, even for the prereduced at 300 ◦ C catalysts
oxygen for the oxidation was supplied from the lattice and References
not from the gas phase. Therefore, oxidation reaction goes
through the stage of ceria reoxidation with a quite possible [1] A. Trovarelli, C. de Leitenburg, M. Boaro, G. Dolcetti, Catal. Today
formation of the certain more active forms of oxygen coming 50 (1999) 353.
from the bulk. Still, we are admitting that relatively fast trans- [2] J. Kaspar, P. Fornasiero, M. Graziani, Catal. Today 50 (1999) 285.
[3] T. Jin, T. Okuhara, G.J. Mains, J.M. White, J. Phys. Chem. 91 (1987)
port of oxygen vacancies at ambient temperature is a sort of 3310.
a mystery just when analyzing the recent results of the works [4] P.J. Gellings, H.J.M. Bouwmeester, Catal. Today 58 (2000) 1.
by Hosono and co-workers on formation and transport of oxy- [5] X. Zhang, K.J. Klabunde, Inorg. Chem. 31 (1992) 1706.
gen species into the bulk of nanoporous oxide 12CaO·7Al2 O3 [6] E. Abi-Aad, R. Bechara, J. Grimblot, A. Aboukais, Chem. Mater. 5
(1993) 793.
[34,35]. In the latter case, superoxide species were formed in
[7] M. Haneda, T. Mizushima, N. Kakuta, J. Chem. Soc., Faraday Trans.
the bulk only at elevated temperature. 91 (1995) 4459.
Final remarks are necessary on the effects of platinum [8] J. Soria, A. Martı́nez-Arias, J.C. Conesa, J. Chem. Soc., Faraday
doping. Quite expectedly, platinum increased the rate of O2 − Trans. 91 (1995) 1669.
A. Aboukais et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 260 (2005) 199–207 207

[9] J. Soria, A. Martı́nez-Arias, J.M. Coronado, J.C. Conesa, Colloids [22] M. Gideoni, M. Steinberg, J. Solid State Chem. 4 (1972) 370.
Surf. A 115 (1996) 215. [23] M. Figaj, K.D. Becker, Solid State Ionics 141–142 (2001)
[10] J. Soria, J.M. Coronado, J.C. Conesa, J. Chem. Soc., Faraday Trans. 507.
92 (1996) 1619. [24] V.A. Radtsig, Kinet. Katal. 37 (1996) 302.
[11] A. Martı́nez-Arias, J.M. Coronado, R. Cataluna, J.C. Conesa, J. So- [25] W. Chamulitrat, L. Kevan, J. Phys. Chem. 89 (1985) 4989.
ria, J. Phys. Chem. B 102 (1998) 4357. [26] X. Zhang, K.J. Klabunde, Inorg. Chem. 31 (1992) 1706.
[12] J.J. Lecomte, P. Granger, L. Leclercq, J.F. Lamonier, A. Aboukais, [27] M. Setaka, T. Kwan, Bull. Chem. Soc. Jpn. 43 (1970) 2227.
G. Leclercq, Colloids Surf. A 158 (1999) 241. [28] M. Che, J.F.J. Kibblewhite, A.J. Tench, M. Dufaux, C. Naccache, J.
[13] J. Soria, J.C. Conesa, A. Martı́nez-Arias, Colloids Surf. A 158 (1999) Chem. Soc., Faraday Trans. I 69 (1973) 857.
67. [29] A.L. Tarasov, L.K. Przheval’skaya, V.A. Shvets, V.B. Kazanskii,
[14] A. Martı́nez-Arias, M. Fernández-Garcı́a, L.N. Salamanca, R.X. Kinet. Katal. 29 (1988) 1181.
Valenzuela, J.C. Conesa, J. Soria, J. Phys. Chem. B 104 (2000) [30] M.I. Zaki, G.A.M. Hussein, S.A.A. Mansour, H.M. Ismail, G.A.H.
4038. Mekhemer, Colloids Surf. A 127 (1997) 47.
[15] Y.-F.Y. Yao, J. Catal. 87 (1984) 152. [31] J.P. Holgado, G. Munuera, J.P. Espinos, A.R. Gonzalez-Elipe, Appl.
[16] J.G. Numan, H.J. Robota, M.J. Cohn, S.A. Bradley, J. Catal. 133 Surf. Sci. 158 (2000) 164.
(1992) 309. [32] C. Hardacre, R.M. Ormerod, R.M. Lambert, J. Phys. Chem. 98
[17] S.E. Golunski, H.A. Hatcher, R.R. Rajaram, T.J. Truex, Appl. Catal. (1994) 10901.
B 5 (1995) 367. [33] Z.C. Kang, L. Eyring, J. Alloys Compd. 275–277 (1998) 30.
[18] A.F. Diwell, R.R. Rajaram, H.A. Shaw, T.J. Truex, Stud. Surf. Sci. [34] K. Hayashi, S. Matsuishi, M. Hirano, H. Hosono, J. Phys. Chem. B
Catal. 71 (1991) 139. 108 (2004) 8920.
[19] C. Serre, F. Garin, G. Belot, G. Maire, J. Catal. 141 (1993) 9. [35] S. Yang, J.N. Kondo, K. Hayashi, M. Hirano, K. Domen, H. Hosono,
[20] A. Holmgren, F. Azarnoush, E. Fridell, Appl. Catal. B 22 (1999) 49. Chem. Mater. 16 (2004) 104.
[21] M. Che, A.J. Tench, Adv. Catal. 32 (1983) 2. [36] W.C. Conner, J.L. Falconer, Chem. Rev. 95 (1995) 759.

You might also like