You are on page 1of 10

Topics in Catalysis 8 (1999) 189–198 189

Generation of superoxide ions at oxide surfaces


Masakazu Anpo a , Michel Che b , Bice Fubini c , Edoardo Garrone c , Elio Giamello c and Maria Cristina Paganini c
a Department of Applied Chemistry, College of Engineering, Osaka Prefecture University, Gakuen-cho 1-1, Sakai, Osaka 599-8531, Japan
b Laboratoire de Réactivité de Surfaces, Université P. et M. Curie (Paris VI), 4 Place Jussieu, 75252 Paris, France
c Dipartimento di Chimica IFM, Università di Torino, Via P. Giuria 9, 10125, Torino, Italy

The superoxide radical anion O− 2 is both an important intermediate in heterogeneous catalytic oxidation and a useful probe for
positive charges in ionic solids, such as metal oxides and zeolites. The paper illustrates the main circumstances under which stable
superoxide anions are formed at surfaces: (i) direct surface–oxygen electron transfer; (ii) photoinduced electron transfer; (iii) surface
intermolecular electron transfer; (iv) decomposition of hydrogen peroxide.
Keywords: superoxide ions, radical ion, oxide surfaces, EPR study, hydrogen peroxide

1. Introduction cationic fields, the principal routes to generate superoxide


on metal oxides or in zeolites, not previously compared,
Reactive oxygen species adsorbed at surfaces are impor- will be illustrated and collectively reviewed.
tant intermediates in both total and selective catalytic oxida-
tion [1–3], as well as in various phenomena taking place at
gas–solid and liquid–solid interfaces, such as those involv- 2. The adsorbed superoxide radical anion as a probe
ing the contamination of highly pure materials [4] or in the of the surface crystal field
pathogenic activity of some solid materials [5]. Among the
various species reported in the literature [2,3], the superox- The superoxide radical anion O− 2 is characterized by the
ide radical anion O− 2 has been widely investigated because, presence of three electrons in the two 2π ∗ antibonding or-
being paramagnetic, the EPR (electron paramagnetic reso- bitals, thus rendering it a paramagnetic species. However,
nance) technique provides a great deal of information on an EPR spectrum is only observed when the degeneration
its structure, location, stability, surface mobility and on-site of the two orbitals is removed by an external perturbation,
dynamics. such as the electric field of surrounding ions. Particular
The adsorbed superoxide ion, besides having an intrinsic cases are when the superoxide radical is trapped in the bulk
interest in the above mentioned fields, costitutes a particu- of an ionic solid, or when it is adsorbed at the surface of
larly useful probe of the electric field at the surface of ionic an oxide. The latter case is the one which is treated in this
solids, particularly of oxides [6]. This is because its EPR paper.
features depend on the local positive field felt by the neg- As the unpaired electron of the superoxide ion is con-
ative anion, usually stabilized on positive ions at the solid fined in a π type orbital, its EPR spectrum is intrinsically
surface. In this way, in some instances, the features of the orthorhombic, in that the three axes of the molecule are
superoxide species formed during a catalytic process yield magnetically unequivalent. The expected principal values
information on the nature of the active sites. In other cases of the g tensor were first derived by Känzig and Cohen [7]
the superoxide is generated at the surface of ionic solids or for the case of superoxide impurities in the bulk of alkali
in the framework of zeolites in order to obtain information halides. Assuming that the z direction corresponds to the
on the cationic fields present in the system. internuclear axis and the y direction to the orbital hosting
The present article presents a brief survey of the condi- the unpaired electron, these are as follows:
tions in which O− 2 species can be formed at the surface of
metal oxides (by far the ionic solids most relevant to cataly- gxx = ge ;
sis) and to illustrate the possible methods for the generation gyy = ge + 2λ/E;
of the species and for its use as a surface probe. Compari- gzz = ge + 2λ/∆,
son with similar, but more general, review work [2,3] will
show that in the present case the approach is more lim- where ge is the free spin value (2.0023), λ is the spin–orbit
ited: for instance, it does not take into consideration the coupling constant for the oxygen atom, and E and ∆ are
structural features of the various oxygen species nor does the two energy intervals reported in the scheme of figure 1.
it make a systematic analysis of all the possible types of The original expressions given by Känzig and Cohen are
oxide. Nevertheless the merit of the present article is that, more complex: the first-order approximation adopted here
after a second section devoted to pointing out the prop- is, however, compatible with the scope of the present arti-
erties of the superoxide radical ion as a probe of surface cle. E and ∆ (figure 1) contribute to determine the gyy and

 J.C. Baltzer AG, Science Publishers


190 M. Anpo et al. / Generation of superoxide ions at oxide surfaces

the gzz values, respectively: in the absence of the splitting field felt by the adsorbed anion and is therefore a suitable
∆ the relaxation time of the radical is too small to give an parameter for measuring the extent of the surface electro-
observable linewidth to the EPR spectrum. Since E  ∆, static interaction.
gzz is the g-tensor element most sensitive to the cationic The value of gzz depends on the ionic charge of the
cation on which the O− 2 ion radical is adsorbed (or, more
precisely, on its charge to radius ratio) and values rang-
ing from 2.12–2.13 (monovalent cations) to 2.02–2.04 for
hexavalent cations [3] have been reported in the literature.
The merit of this method, however, goes beyond the mere
identification of the cation charge. This has been shown, for
instance, in the case the superoxide adsorbed on MgO [8].
In figure 2, a spectrum of O− 2 on MgO is shown. The gzz
region reveals a clear and well resolved heterogeneity in the
adsorbed species, despite the fact that only one positively
charged chemical entity (the divalent Mg2+ ) is present at
the surface. The various components of the gzz region have
to be ascribed [8] to superoxide ions adsorbed on surface
Mg2+ ions differing in their local coordination. The surface
field of a cation increases when the number of anions sur-
rounding the cation decreases. The highest gzz value is thus
ascribed to superoxide species adsorbed on five-coordinate
Mg2+ ions, whereas the lowest value is due to particularly
well exposed three-coordinate cations exerting a stronger
Figure 1. Energy levels scheme for the superoxide O−
2 radical ion ad-
electric field as a result of their low coordination. Such an
sorbed on a surface. interpretation, first advanced on the basis of an empirical

Figure 2. A typical EPR spectrum of the O−


2 radical ion adsorbed on MgO and exhibiting heterogeneity of the gzz component.
M. Anpo et al. / Generation of superoxide ions at oxide surfaces 191

correlation [8] between the gzz values and the Madelung 3.1. Electron transfer from isolated transition metal ions
constant of different families of surface cations [9], has re- embedded in an insulating matrix
cently been confirmed by quantum chemical calculations
on the O− 2 /MgO system [10]. The population of the vari- An insulator like MgO is not capable of direct electron
ous gzz components observed in figure 2 is not the same transfer towards dioxygen, since all the electrons in the full
in all cases of O−
2 on MgO, as it depends on the procedure valence band (the top of which is separated by a large gap of
adopted for the superoxide generation (see section 3.3 and about 8 eV from the bottom of the empty conduction band)
figure 8). lies at lower energy than that of an hypothetical adsorbed
superoxide ion. The properties of the insulating oxide can
be modified by introducing a transition metal ion into the
3. The formation of adsorbed superoxide species via framework of the oxide. This, in electronic terms, corre-
direct oxide–dioxygen electron transfer sponds to modifying the band structure of the solid by the
When molecular oxygen is contacted with the surface of formation of new localized states which usually lie within
a metal oxide, the formation of O− 2 will in principle occur
the band gap of the insulator. Interesting results have been
if the energy of the adsorbed negative moiety lies below the obtained by doping MgO with Co2+ ions, due to the mutual
Fermi level of the solid. Since the electron affinity of oxy- solubility of MgO and CoO over the whole range of mo-
gen is positive but rather small (0.44 eV), the electrostatic lar concentrations. Fully dehydrated (but not prereduced)
contribution to the stabilisation of the anionic species on the MgO–CoO solid solutions directly adsorb oxygen giving
positive ions at the surface plays a fundamental role. The rise, depending on the adsorption temperature, to different
energies relating to the various steps of the electron transfer EPR spectra [11]. Two stable species have been identi-
from an oxide to molecular oxygen have recently been ana- fied at 77 K, transforming at 120–140 K into two other
lyzed by quantum chemistry calculations [10], for the case species, stable up to room temperature (figure 3(a)). Both
of O− 2 formation on magnesium oxide containing surface
the low temperature species and one of the room temper-
trapped electrons (Fs centers). Although the stability of the ature species (species A in figure 3(a)) are stabilized on
electrons in the surface traps is substantial, the large electro- cobalt ions, as shown by the complex superhyperfine struc-
static energy related to the O− 2 Mg
2+
interaction becomes ture due to the interaction between the unpaired electron
the driving force that allows the electron transfer. With this and the 59 Co nucleus (nuclear spin I = 7/2, multiplicity
in mind, it is easy to divide pure or mixed oxides into two (2I + 1) = 8 lines). As indicated by a detailed analysis of
classes. The former class (section 3.1 below) groups the their spin-Hamiltonian parameters [12], their nature can be
relatively few systems capable of electron transfer towards roughly indicated as Co3+ O− 2 . The ionic model, however,
oxygen without any preliminary reductive treatment, while is not totally adequate, and models derived from coordi-
the latter (sections 3.2 and 3.3 below) encompasses oxides nation chemistry have to be adopted to explain the metal
which must be prereduced prior to oxygen adsorption. ion–oxygen chemical bond satisfactorily [12]. The fourth

Figure 3. EPR spectra of oxygen radical species on MgO–CoO: (a) in the presence of a low pressure of oxygen; (b) after evacuation. The process is
reversible.
192 M. Anpo et al. / Generation of superoxide ions at oxide surfaces

species (B in figure 3(a)) is overlapped by species A and


has been identified as a “classic” superoxide ion on Mg2+
this being formed by spillover of O− 2 from the less stable
of the cobalt adducts upon increasing the temperature from
77 to 298 K.
The most noticeable feature of the signal of the Co3+ O−2
species in figure 3(a) is that it disappears upon evacuation
at room temperature and consequent oxygen release, leav-
ing only species B at the surface (figure 3(b)). Readmission
of oxygen at room temperature restores the cobalt–oxygen
adduct. This reversible adsorption of oxygen indicates that
a fraction, at least, of the surface cobalt centres behaves
as oxygen carriers exactly like those of iron or cobalt ions
in natural and synthetic oxygen carriers. The basic rea-
son for this unique property of the MgO–CoO surface is
that the symmetry of the Co2+ coordination environment
(C4v ) reproduces that of the molecular analogues mentioned
above [13].

3.2. Electron transfer from reducible (semiconducting)


oxides

Oxides capable of electron transfer towards dioxygen af-


ter a preliminary reductive treatment (thermal treatment un-
der vacuum or chemical reduction in mild conditions) are in
many cases semiconductors, showing deviations from stoi-
chiometry and/or possessing cations stable in more than one
oxidation state. In an n-type semiconductor, for instance,
the donor centres are occupied electron levels lying in the
band gap, quite close to the lower limit of the conduction
band, and the Fermi level of the solid lies between the
donor levels and the conduction band itself. At relatively
low temperature (for example, at ambient temperature) the
electrons occupying the donor levels are easily excited into
the conduction band. The electron transfer towards oxygen
occurs because the energy level of the LUMO orbital of O2
lies below the energy of the Fermi level of the solid. Elec-
tron transfer causes a decrease in the electron population in
the conduction band and a concomitant decrease of the solid
conductivity. In the interfacial region of the solid a posi-
tive charge develops, forming, with the negative charge of Figure 4. EPR spectra of superoxide radical ions adsorbed on: (a) TiO2
(anatase); (b) TiO2 supported WO3 .
the adsorbed species, an electrical double layer [14]. This
causes an increase of the electrical work function or, in
other words, generates an energy barrier (Schottky barrier) tion. When such a partially reduced solid is contacted with
which increases with progressive electron transfer, thereby oxygen at room temperature the grey colour immediately
limiting the electron transfer itself. In terms of energy band vanishes, the solid no longer perturbs the quality factor of
schemes, this can be seen as a bending of the bands of the the EPR cavity, and also its IR transparency improves. In
solid with consequent vanishing of the energy difference such conditions, the sample exhibits an EPR spectrum (fig-
between the energy level of the surface adsorbed oxygen ure 4(a)), readily assigned to O−2 ions stabilized on Ti
4+

and the Fermi level of the solid [14]. cations at the solid surface. More than one gzz compo-
An example of this behaviour is given by titanium diox- nent characterises the spectrum: the most abundant species
ide, when outgassed at 883 K. Non-stoichiometry is en- corresponds to a gzz value of 2.023, and less abundant
hanced because of oxygen loss and the solid turns to pale species correspond to shoulders at 2.027 and 2.020 indicat-
grey. The conduction band becames populated by a cer- ing a weakly heterogeneous distribution of the superoxide
tain quantity of electrons, causing a loss of transparency in adsorption sites.
the IR, perturbations of the Q factor of the cavity in the An example of the role of adsorbed superoxide as probe
EPR and by modifications of the profile of the UV absorp- of cationic sites is found when comparing the results de-
M. Anpo et al. / Generation of superoxide ions at oxide surfaces 193

scribed above for titanium dioxide with those arising from


the same treatment of WO3 supported on TiO2 (TW), in
which a surface layer of hexavalent tungsten partially cov-
ers the surface of titania [15]. When the supported oxide is
reduced under vacuum the binary TW sample reacts with
oxygen at room temperature leading to the EPR spectrum
in figure 4(b), whose features are different from those of the
spectrum in figure 4(a). In particular, in the gzz region only
one sharp component is present, at g = 2.019: it should
be noticed that, with respect to the other components (gyy
and gxx ) this gzz component is lower than in the previous
case. Although the gzz values observed in the two cases
are quite close, it seems reasonable to infer that O− 2 in the
TW case (figure 4(b)) is not stabilized on the same Ti sites
as are available on bare TiO2 but, most probably, is stabi-
lized on Wx+ ions of the WO3 overlayer. The g value is in
fact in agreement with those reported for superoxide ions
on WO3 /SiO2 [16].
The reduction of molecular oxygen at the surface of
slightly reduced transition metal oxides described above
is an important step in the complex series of redox re-
actions occurring during oxidation processes taking place
on mixed or supported oxide systems with at least one re-
ducible oxide phase. The mechanism was first described
by Mars and van Krevelen in the 1950s for the selective
oxidation of aromatic hydrocarbons [17]. A good candidate
for catalysing selective oxidation reactions is an oxide with
a certain propensity to lose oxygen and a good electronic
conductivity to ensure efficiency for the described redox Figure 5. EPR spectra of: (a) activated ZrO2 containing surface Zr3+
process. The capability of superoxide formation after mild ions; (b) the same sample after contact with O2 and surface adsorbed
reducing treatments is thus a property linked with the effi- superoxide formation.
cency of a given solid for catalysing oxidation reactions.
A behaviour similar to that of TiO2 is shown by zirco- of charge carriers delocalized in the solid and available for
nium dioxide (an oxide which is the object of increasing electron transfer in given conditions.
attention as a catalyst or catalyst support). Figure 5(a) re-
ports the EPR spectrum of a sample of zirconia after activa-
tion in vacuo at 873 K: the weak signals at g values lower 3.3. Electron transfer from localized electron defect
than the free-spin value (gk = 1.953 and g⊥ = 1.978) centres
are due to surface Zr3+ ions [18]. Contact with oxygen
at 423 K gives rise to the intense orthorhombic signal re- It is known since the early sixties that irradiation of
ported in figure 5(b), readily assigned again to superoxide some oxides with highly energetic radiation generates de-
ions. The gzz value is clearly compatible with a tetrava- fects consisting of electrons trapped in anionic vacancies,
lent adsorption site, so that the assignment to the O− 2 /Zr
4+ the so-called F centres or colour centres [19]. Such cen-
system is straightforward. The electron transfer does not tres can also be generated by addition of a low ionisation
occur at room temperature. This indicates that the process energy metal to the solid: if the addition is performed at a
is slightly activated. The intensity of the superoxide species reasonably moderate temperature (from RT to 600 K), the
(figure 5(b)) overwhelms that of the Zr3+ signal in the centres are localized at the solid surface [20,21]. In fig-
starting sample (figure 5(a)). This latter, moreover, does ure 6(a) the EPR spectrum of MgO containing F centres
not completely disappear after superoxide formation (fig- is reported. The visible ones are one-electron centres, but
ure 5(b)). This indicates that the majority of the oxygen two-electron diamagnetic centers are also present. The sur-
molecules are not reduced by Zr3+ ions and that the reduc- face colour centres are unstable towards oxygen and O− 2
ing centres, if localized, are not paramagnetic. It seems, radical ions are formed by oxygen adsorption at low tem-
however, more probable, due to the semiconducting prop- perature as indicated by the spectrum in figure 6(b). The
erties of the oxide, that the electrons transferred to the sur- electronic state corresponding to F centres lies in the band
face adsorbed dioxygen are, in the partially reduced solid, gap of the oxide. As mentioned before, the electron transfer
delocalized over the conduction band. The formation of to oxygen occurs due to the contribution of both the elec-
superoxide species can thus be used to monitor the amount tron affinity of oxygen and the Madelung stabilisation of
194 M. Anpo et al. / Generation of superoxide ions at oxide surfaces

the adsorbed anion on positive centres at the surface. This into a silicalite framework (in the case of the TS-1 titanium
latter contribution is by far larger than the former [10]. silicate) exhibit different gzz features from superoxide ad-
Electron transfer to oxygen from localized defects has sorbed on titanium ions introduced in the zeolitic channels
also been observed in the past in the case of variously acti- by ion exchange [23], thus allowing unambiguous discrim-
vated zeolitic systems (mainly activated by irradiation with ination for isomorphous substitution.
high energy photons) [22]. More recently, similar inves-
tigations have been performed to understand the nature of
the reduced centers in new synthetic molecular sieves and 4. Photoinduced electron transfer
to confirm the isomorphous substitution of a given element
in the framework of such systems [23–25]. For instance,
An alternative method to generate adsorbed superoxide
superoxide radical ions adsorbed on Ti4+ ions incorporated
ions, when the solid-to-dioxygen electron transfer does not
occur spontaneously, consists of irradiating the solid in the
presence of molecular oxygen. This method was employed
in early work on zeolites [26,27] using both γ and UV ir-
radiation to check the magnitude of the electric field in the
framework as a function of the metal ions exchanged in the
structure. More recently superoxide photoformation was
used to characterise the efficiency of photocatalytic sys-
tems based on highly dispersed titanium oxide [28]. The
spectrum in figure 7 has been obtained by UV irradiation
at 77 K, in the presence of molecular oxygen, of a sam-
ple of titanium dioxide highly dispersed on porous Vycor
glass. The intensity of the spectrum increases linearly with
irradiation time. The g values of the spectrum are reported
in the figure. The gzz value is, as expected, very close to
the value recorded for superoxide on polycrystalline TiO2
(figure 4). It is worth mentioning that highly dispersed
titanium dioxide is normally photoluminescent due to the

Figure 6. (a) EPR spectrum of surface F+ s centres on MgO; (b) EPR


spectrum of adsorbed superoxide formed upon contact of oxygen with
the surface containing F centers. The higher intensity of the superoxide
spectrum indicates that a fraction of the surface trapped electrons are kept Figure 7. EPR spectrum of superoxide ions formed upon irradiation of
in diamagnetic centres. TiO2 highly dispersed on porous Vycor glass.
M. Anpo et al. / Generation of superoxide ions at oxide surfaces 195

following charge transfer process



Ti4+ O2−  [Ti3+ O− ]∗
hν 0
0
The luminescence hν is quenched by adsorption of oxygen.
Oxygen addition thus affords an alternative route for the
transfer of the excitation energy. A photoformed electron is
transferred and stabilized on the adsorbed molecule, which
is reduced to superoxide radical ion. Analogous results
have been found in the case of the photocatalytic oxida-
tion of propene occurring on divanadium pentoxide highly
dispersed on a porous Vycor glass (PVG) support [29].

5. Surface intermolecular electron transfer

Although the formation of O− 2 at the surface of an insu-


lating oxide by direct electron transfer is limited by the low
energy of the electrons in the solid, it has been shown that
several hydrocarbons (propene and toluene, for instance)
are oxidized by molecular oxygen in mild conditions at the
surface of MgO with formation of oxidized products and
cleavage of carbon–carbon bonds. In the case of propene,
for example, the contact with oxygen at the surface of MgO
causes the formation of adsorbed acetates, formates and car-
bonates at ambient temperature and total oxidation at higher
temperature. This fact is the fingerprint for an electrophilic
oxidation taking place through a superoxide ion or an O−
species [1]. Superoxide O− 2 ions were actually observed at
the surface of thoroughly outgassed MgO in the presence
of hydrocarbon–oxygen mixtures. This fact has been in-
terpreted introducing the concept of surface intermolecular
electron transfer (SIET) [30–33]. The fundamental factor
promoting the observed reactivity is the basic nature of
the MgO surface. Particularly low-coordinate surface ox-
ide O2− ions are able to split the hydrocarbon molecule Figure 8. EPR spectra of superoxide radical ions on MgO formed upon:
heterolytically forming a proton (stabilized as OH− ) and a (a) propene–O2 surface interaction; (b) adsorption of oxygen on the
carbanion: hydrogen-precovered surface.

− −
(surf) → R(ads) + OH(surf)
R–H + O2− The same mechanism has been shown to apply when
superoxide species are formed on alkaline earth oxides
The carbanion, which has indeed been observed by UV
(MgO and CaO) by adsorption of oxygen on a hydrogen-
spectroscopy [34], is unstable towards molecular oxygen
precovered surface (figure 8(b)). In this case the superoxide
and readily transfers an electron to dioxygen itself:
ion is formed by electron transfer from a surface adsorbed
R− − − −
(ads) –OH(surf) + O2(g) → O2 + OH(surf) + R
hydride. This comes from the heterolytic dissociation of
molecular hydrogen on a pair of Mg2+ O2− ions having par-
A minor fraction of the superoxide ions formed is captured ticularly low coordination and hence high polarizing power.
and stabilised by low coordination Mg2+ ions (near to a The whole process is summarized as follows:
surface hydroxyl) and becomes EPR visible (figure 8(a)).
The remaining species react, probably in their protonated
form HO2 , with the radical R leading to the cleavage of
the molecule and the formation of oxidized products. The
above described SIET mechanism shows how an oxida-
tion reaction involving superoxide species can take place
without direct activation of molecular oxygen by electron The above reaction scheme has been recently confirmed
transfer from the solid. In this case the driving force of the by IR investigations which show the formation of a hy-
process is merely the basic strength of the oxide ions at the dride band on contact with H2 and its disappearance when
surface. oxygen is subsequently admitted [35]. The superoxide ion
196 M. Anpo et al. / Generation of superoxide ions at oxide surfaces

resulting from the above process exhibits the EPR spectrum


in figure 8(b). As in the case of figure 8(a), the magnetic
interaction with a nearby hydroxyl group is witnessed by
the hydrogen superhyperfine doublets visible on the gyy and
gzz components, as discussed in a previous paper [36]. The
difference in the gzz regions in figures 8(a) and (b) (indi-
cating that different species are formed) is related to the
different chemistry of the two processes described above,
and will be discussed elsewhere [37].

6. Interaction with hydrogen peroxide

A new method to generate superoxide at the surface of


oxides has recently been proposed by some of us [38]. It Figure 9. EPR spectrum of O−2 superoxide ions formed at the surface of
consists of the treatment of an oxide with a solution of zirconium dioxide by interaction with hydrogen peroxide.
hydrogen peroxide, followed by drying the solid and heat-
ing under vacuum at a suitable temperature. This method isation of an adsorbed superoxide has been already widely
derives from the observation that many oxides catalyse discussed [38,40] and it is based on the following series of
the decomposition of hydrogen peroxide, a reaction which elementary steps:
occurs via radical intermediates, some of which are, in
H2 O2 → 2OH·
particular conditions, stabilized at the surface. The main
goal of the research was to monitor the reaction interme- OH· + H2 O2 → H2 O + HO2 ·
diates during heterogeneous catalytic oxidation by H2 O2 . The HO2 · hydroperoxyl radical is a weak acid and is also
Indeed, an increasing importance has recently been as- the protonated form of the superoxide radical ion which
sumed by the processes of heterogeneous oxidation of var- easily forms at the surface of the oxide according to the
ious substrates (alcohols, olefins, aromatic rings) on ze-
following reaction:
olitic catalysts such as TS-1 titanium silicate [39]. All
HO2 · −→ H+ + O−
these processes, however, occur in competition with the MeO
2
spontaneous decomposition reaction of H2 O2 to water and
molecular oxygen, the only process taking place by con- The unavoidable contact of the oxide with an aqueous
tact of H2 O2 with the solid in the absence of the organic phase may be a limit of this method with some solids. In
substrate. Adsorbed superoxide is not the only intermedi- certain cases, however, this method has clear advantages
ate species formed in the H2 O2 –oxide interaction but it is over other procedures. This is the case, for instance, with
by far the one most easily observed. In the case of the zeolites. Superoxide ions in zeolites have been widely in-
H2 O2 –MgO system stable hydroxyl OH radicals were also vestigated in the past employing as the method of forma-
observed [40]. tion γ irradiation or UV irradiation of the thoroughly de-
A typical example of superoxide generation via catalytic hydrated solid in the presence of oxygen, a method which
decomposition of hydrogen peroxide is that observed in the involves the creation of defects in the aluminosilicate struc-
case of zirconium dioxide. When this solid is contacted ture [26,27]. Some of us have recently shown [41] that
with an H2 O2 /H2 O solution the decomposition of hydrogen superoxide species can be formed under much milder con-
peroxide is readily catalysed as is shown by the abundant ditions by the mere interaction of a hydrogen peroxide so-
development of oxygen. The solid (dried at 343 K) exhibits lution with the zeolite. Outgassing the system at relatively
an intense EPR spectrum (figure 9), readily assigned to a low outgassing temperatures (370–470 K) after this treat-
superoxide O− 2 radical ion adsorbed on a Zr
4+
ion at the ment, causes the appearance of a superoxide spectrum. The
surface. The spectrum is closely similar to that reported in spectrum is then observed up to temperatures between 670
figure 5(b). There is no doubt, therefore, about the forma- and 770 K. This indicates that the stability of the intraze-
tion of superoxide on zirconia also after contact with H2 O2 . olitic superoxide ions is higher than that of the “classic”
This indicates, quite surprisingly, that the surface exposed superoxide species adsorbed on the surface of a bulk ox-
Zr4+ ions capable of O− 2 coordination are similar in the ide, which is usually destroyed by treatments under vacuum
two cases, one of which relates to ZrO2 which has been at 470–500 K.
strongly dehydrated at 870 K (figure 5(a)) and the other to Figure 10 reports a series of spectra recorded in a NaY
ZrO2 mildly dehydrated at 343 K (figure 9). The absence zeolite after treatment with hydrogen peroxide and succes-
of any appreciable heterogeneity in the gzz component of sive dehydration at increasing temperature. The progressive
the superoxide spectrum on ZrO2 indicates furthermore that change in the EPR spectrum monitors the different electro-
only one family of surface ions acts as sites for superoxide static situation present in the zeolite cavity containing the
coordination. The reaction mechanism leading to the stabil- superoxide as the temperature is raised and provides in-
M. Anpo et al. / Generation of superoxide ions at oxide surfaces 197

however, is rather reactive and its reactivity often limits the


applicability of the method as a diagnostic tool. The use
of EPR-active probes complements the use of diamagnetic
probes of the surface crystal field such as the CO mole-
cule whose interaction with the surface can be followed by
changes of the C–O stretching frequency monitored by IR
spectroscopy [43].

References
[1] A. Bielanski and J. Haber, Oxygen in Catalysis (Dekker, New York,
1991).
[2] M. Che and A.J. Tench, Adv. Catal. 31 (1982) 77.
[3] M. Che and A.J. Tench, Adv. Catal. 32 (1983) 1.
[4] D.J. Miller, D. Haneman, Phys. Rev. B 3 (1971) 2918.
[5] B. Fubini, L. Mollo and E. Giamello, Free Rad. Res. 23 (1995) 593.
[6] M. Che and E. Giamello, Stud. Surf. Sci. Catal. 57 B (1990) 265.
[7] W. Känzig and M.H. Cohen, Phys. Rev. Lett. 3 (1959) 509.
[8] E. Giamello, E. Garrone and P. Ugliengo, J. Chem. Soc., Faraday
Trans. I 85 (1989) 1373.
[9] E. Garrone, A. Zecchina and F.S. Stone, Phil. Mag. B 42 (1980) 683.
[10] G. Pacchioni, A. Ferrari and E. Giamello, Chem. Phys. Lett. 255
(1996) 58.
[11] E. Giamello, Z. Sojka, M. Che and A. Zecchina, J. Phys. Chem. 80
(1986) 6084.
[12] Z. Sojka, E. Giamello, M. Che, K. Dyrek and A. Zecchina, J. Phys.
Chem. 92 (1988) 1541.
[13] M. Che, K. Dyrek, E. Giamello and Z. Sojka, Zeit. Phys. Chem.
Neue Folge 152 (1987) 397.
[14] J.R. Morrison, The Chemical Physics of Surfaces (Plenum, New
York, 1977).
[15] M.C. Paganini, L. Dall’Acqua, E. Giamello, L. Lietti, P. Forzatti and
G. Busca, J. Catal. 166 (1997) 195.
[16] R.F. Howe, J. Chem. Soc., Faraday Trans. I 71 (1975) 1689.
Figure 10. EPR spectra of superoxide radical ions formed upon interaction
[17] P. Mars and D.W. van Krevelen, Chem. Eng. Sci. 3 (1954) 41.
between hydrogen peroxide and NaY zeolite and successive outgassing at
[18] C. Morterra, E. Giamello, L. Orio and M. Volante, J. Phys. Chem.
increasing temperature. The outgassing temperature of each experiment is
94 (1990) 3111.
indicated on the left hand side of the corresponding spectrum.
[19] A.J. Tench and R.L. Nelson, J. Colloid and Interface Sci. 26 (1968)
364.
formation on the variation of the electrostatic field in the [20] E. Giamello, A. Ferrero, S. Coluccia and A. Zecchina, J. Phys. Chem.
zeolite which parallels the progressive dehydration. A full 95 (1991) 9385.
interpretation of these preliminary results [42] has not yet [21] E. Giamello, D. Murphy, L. Ravera, S. Coluccia and A. Zecchina,
been accomplished but there is ample evidence already to J. Chem. Soc., Faraday Trans. 90 (1994) 3167.
[22] J.C. Védrine, A. Abou-Kais, J. Massardier and G. Dalmai-Imelik,
show that the use of hydrogen peroxide and its decomposi- J. Catal. 29 (1973) 120.
tion to produce intrazeolitic superoxide ions is a promising [23] A. Tuel, J. Diab, P. Gelin, M. Dufaux, J.F. Dutel and Y. Ben Taarit,
method to improve knowledge of the properties of zeolitic J. Mol. Catal. 63 (1990) 95.
materials. [24] S.B. Hong, S.J. Kim and Y. Sun Hu, J. Phys. Chem. 100 (1996)
15923.
[25] A.M. Prakash, V. Kurshev and L. Kevan, J. Phys. Chem. B 101
(1997) 9794.
7. Conclusions [26] P.H. Kasai, J. Chem. Phys. 43 (1965) 3322.
[27] K.M. Wang and J.H. Lunsford, J. Phys. Chem. 74 (1970) 1512.
Only in a few cases is O− 2 formed on a substantially [28] M. Anpo, N. Aikawa, Y. Kubokawa, M. Che, C. Louis and E.
unperturbed surface. In most cases the surface needs to Giamello, J. Phys. Chem. 89 (1985) 5017.
be activated, prior to O−
2 formation, by a suitable chemical
[29] M. Anpo, T. Suzuki, Y. Yamada, E. Giamello and M. Che, in: New
Developments in Selective Oxidation, eds. G. Centi and F. Trifirò,
and/or thermal treatment causing a significant (though often Stud. Surf. Sci. Catal. 11 (1990) 683.
very small) perturbation of the bare surface. The EPR spec- [30] E. Garrone and F.S. Stone, in: Proc. 8th Int. Conf. Catal., Vol. 4,
trum of surface-adsorbed superoxide ions provides a useful Berlin, 1984 (Verlag Chemie, Weinheim, 1984) p. 441.
tool for monitoring the surface electric field of an oxide [31] E. Garrone, E. Giamello, S. Coluccia, G. Spoto and A. Zecchina, in:
catalyst. Similar investigations can be also performed, by Proc. 9th Int. Conf. Catal., Vol. 4, Calgary, 1988, eds. M.J. Phillips
and M. Ternan (Chem. Inst. Canada, Ottawa, 1988) p. 1577.
EPR, using the NO radical molecule as a probe. The advan- [32] E. Giamello, E. Garrone, S. Coluccia, G. Spoto and A. Zecchina,
tage of employing NO is that the molecule can be directly in: New Developements in Selective Oxidation, eds. G. Centi and F.
adsorbed on the unperturbed surface. The NO molecule, Trifiró, Stud. Surf. Sci. Catal. 11 (1990) 817.
198 M. Anpo et al. / Generation of superoxide ions at oxide surfaces

[33] E. Garrone, E. Giamello, M. Ferraris and G. Spoto, J. Chem. Soc., [39] M.G. Clerici, G. Bellussi and U. Romano, J. Catal. 129 (1991) 159.
Faraday Trans. 88 (1991) 333. [40] E. Giamello, L. Calosso, B. Fubini and F. Geobaldo, J. Phys. Chem.
[34] E. Garrone, A. Zecchina and F.S. Stone, J. Catal. 62 (1980) 396. 97 (1993) 5735.
[35] E. Knözinger, K.H. Jacob and P. Hofman, J. Chem Soc., Faraday [41] B. D’Anna and E. Giamello, Appl. Mag. Res. 10 (1996) 591.
Trans. 89 (1993) 1101. [42] E. Giamello et al., to be published.
[36] E. Giamello, P. Ugliengo, E. Garrone, M. Che and A.J. Tench, [43] V. Bolis, C. Morterra, B. Fubini, P. Ugliengo and E. Garrone, Lang-
J. Chem. Soc., Faraday Trans. I 85 (1989) 3987. muir 9 (1993) 349.
[37] E. Giamello, manuscript in preparation.
[38] E. Giamello, P. Rumori, F. Geobaldo, B. Fubini and M.C. Paganini,
Appl. Mag. Res. 10 (1996) 173.

You might also like