You are on page 1of 10

Applied Catalysis B: Environmental 65 (2006) 21–30

www.elsevier.com/locate/apcatb

Synthesis and characterisation of La1xCaxFeO3 perovskite-type oxide


catalysts for total oxidation of volatile organic compounds
Bibiana P. Barbero a,1, Julio Andrade Gamboa b, Luis E. Cadús a,*
a
Instituto de Investigaciones en Tecnologı́a Quı́mica (INTEQUI), UNSL, CONICET, Casilla de Correo 290, 5700 San Luis, Argentina
b
Comisión de Energı́a Atómica (CNEA), 8400 San Carlos de Bariloche, Rı́o Negro, Argentina
Received 21 May 2005; received in revised form 28 September 2005; accepted 21 November 2005
Available online 2 February 2006

Abstract
La1xCaxFeO3 perovskite-type oxides with x = 0, 0.2 and 0.4 were prepared by the citrate method and characterised by means of X-ray
diffraction (XRD), X-ray fluorescence (XRF), surface area measurement BET, X-ray photoelectron spectroscopy (XPS), Fourier transformed
infrared spectroscopy (FT-IR), laser Raman spectroscopy (LRS), oxygen temperature-programmed desorption (O2-TPD) and temperature-
programmed reduction (TPR). The citrate method shows to be simple and appropriate to obtain single phases avoiding segregation and/or
contamination. Moreover, controlling the calcination temperature, specific surface areas adequate for catalysts to be used in oxidation reactions are
achieved. The structure refinement by using the Rietveld method indicates that the partial calcium substitution modifies the orthorhombic structure
of the LaFeO3 perovskite towards a less distorted one. From XRF and XPS, a slight surface enrichment in lanthanum and calcium was detected.
XRD, FT-IR and TPR results indicated that the electronic debalance caused by the partial substitution for La3+ by Ca2+ is compensated by an
oxidation state increase of a part of Fe3+ to Fe4+. O2-TPD results revealed that at a substitution level higher than x = 0.2, oxygen vacancies are also
formed to preserve the electroneutrality. Finally, an improvement of the catalytic activity in propane and ethanol combustion was observed on the
substituted perovskites. Correlating this with the characterisation results, the active sites would be associated to the Fe4+ ions.
# 2006 Elsevier B.V. All rights reserved.

Keywords: Lanthanum iron perovskite; Calcium; XRD; XRF; XPS; FT-IR; Raman spectroscopy; O2-TPD; TPR; Rietveld refinement; Propane combustion;
Ethanol combustion

1. Introduction as Pt and Pd, which are more expensive and do not resist
operating at high temperatures [1–3]. Perovskite oxides have
An increasing interest has been shown in catalytic general formula ABO3, where the 12-coordinated A sites may be
combustion processes during the last decades since they are occupied by rare-earth, alkaline-earth, alkali or other large ions
a convenient way for emission prevention (the control of and the 6-coordinated B sites are usually filled with transition
nitrogen oxides NOx and unburned hydrocarbons in heat and metal cations. A large number of metallic cations can occupy the
power generation plants) as well as clean-up (volatile organic A and the B sites. Furthermore, the great stability of the
compounds removal, automobile exhaust converters) [1]. perovskite framework allows partial substitution at the A sites
Perovskite-type oxides display prominent catalytic activities and/or the B sites modifying the catalytic, redox and structural
in many fields such as the total oxidation of methane and of properties. The substitution at A site with ions having lower
volatile organic compounds [2]. This activity, coupled with a valence can allow the formation of structural defects such as
high thermal stability, postulates to pervoskite-type oxides as anionic or cationic vacancies and/or a change in the oxidation
potential catalysts in substitution of very active noble metals such state of the transition metal cation to maintain the electro-
neutrality of the compound. When the oxidation state of B cation
increases, the relative ease of the redox process generates larger
quantities of available oxygen at low temperature and the overall
* Corresponding author. Fax: +54 2652 426711.
oxidation activity enhances. Moreover, the oxygen vacancies
E-mail addresses: bbarbero@unsl.edu.ar (B.P. Barbero),
lcadus@unsl.edu.ar (L.E. Cadús). favour the catalytic activity in oxidation reaction because they
1
Fax: +54 2652 426711. increase the lattice oxygen mobility.

0926-3373/$ – see front matter # 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcatb.2005.11.018
22 B.P. Barbero et al. / Applied Catalysis B: Environmental 65 (2006) 21–30

Perovskites with lanthanum in A site and transition metals in ing Ni-filtered Cu Ka radiation (l = 0.15418 nm). The data
B site (LaBO3) with B = Mn, Co, Fe, Ni have received much were collected at 0.028 with a counting time of 5 s per step, in
attention. Partial substitution of divalent ions, in particular Sr2+, the (2u) range from 208 to 908. The crystalline phases were
for lanthanum has been studied producing interesting results. identified by reference to the PDF data employing standard
For example, La1xSrxFeO3 perovskites with x = 0.1 and 0.2 spectra software. The lattice parameters and the structure of the
are more active in propane oxidation than the corresponding catalysts have been estimated from Rietveld’s powder structure
unsubstituted LaFeO3 [4]. The calcium substitution in A site refinement analysis [11].
has been studied from the structural viewpoint [5–7] but few
results about the catalytic applications have been reported. 2.2.3. X-ray fluorescence (XRF)
Ciambelli et al. [8] investigated a La1xCaxFeO3 series as The composition of every sample was determined by X-ray
catalysts for methane combustion and they observed that the fluorescence on a PW1400 Philips instrument. The calibration
catalytic performance of the substituted samples has not been was made by using calibration samples prepared with known
better than that of unsubstituted sample. In contrast, we have amounts of La2O3, Fe2O3 and CaCO3.
studied the effect of calcium substitution in La1xCaxCoO3
perovskites for propane oxidation and it has been beneficial [9]. 2.2.4. X-ray photoelectron spectroscopy (XPS)
On the basis of these results, in this work we reported a XPS analyses were performed on a SSX 100/206
study carried out on La1xCaxFeO3 perovskites (x = 0, 0.2, 0.4) photoelectron spectrometer from Surface Science Instruments
which have been evaluated in total oxidation of propane with a monochromatized microfocused Al X-ray source.
and ethanol. The aspects related to the synthesis process have Spectra were registered after purging the samples at ambient
been carefully controlled. The structural and redox features temperature in vacuum. The residual pressure in the analysis
are determined from the results of the textural and physico- chamber during the analysis was about 106 Pa. The flood gun
chemical characterisation. Furthermore, an explication of the energy was adjusted at 10 eV. The data treatment was
catalytic behaviour is proposed. performed with appropriate software. Binding energies were
calibrated with regard to the C(C, H) component of the C 1s
2. Experimental peak fixed at 284.8 eV. The atomic ratios were calculated
by using the atomic sensitivity factors provided by the
2.1. Catalyst preparation manufacturer.

La1xCaxFeO3 perovskites with x = 0, 0.2, and 0.4 were 2.2.5. Raman spectroscopy
prepared by the citrate method [10]. La(NO3)36H2O (Fluka), The Raman spectra were collected by using two different
Ca(NO3)24H2O (Fluka), Fe(NO3)39H2O (Aldrich) and citric spectrometers: (a) JASCO TRS600SZP multichannel mono-
acid (Mallinckrodt) were used as reagents. The aqueous chromatic spectrometer, and (b) Dilor LabRam Instruments. In
solutions of the metal nitrates were added to an aqueous both cases, the samples were pressed into self-supporting
solution of citric acid with a 10% excess over the number of wafers and the spectra were recorded at ambient temperature.
ionic equivalents of cations. The resulting solution was agitated The excitation source was: (a) an Ar ion laser (514.5 nm)
for 15 min and concentrated slowly by evaporating water under operated at a power of 50 mW, and (b) a He–Ne laser
vacuum in a rotavapor at 75 8C until the formation of a gel. This (632.8 nm) operated at a power of 10 mW.
gel was dried in an oven, increasing slowly the temperature up
to 250 8C and keeping it overnight, to yield a solid amorphous 2.2.6. FT-IR spectroscopy
citrate precursor. The obtained precursor was milled and then FT-IR spectra were registered by using a Nicolet Protegé 460
decomposed in air at 400 8C for 30 min and calcined in air at spectrometer in KBr pellets. The spectra were the result of
700 8C for 2 h. The perovskite catalysts are expressed by averaging out 32 scans obtained at ambient temperature in
La1xCaxFeO3 where x = 0, 0.2, and 0.4, although the actual wavelength ranging from 4000 to 225 cm1.
composition may be non-stoichiometric in relationship to the
oxygen and may not be of single phase. 2.2.7. Oxygen temperature-programmed desorption
(O2-TPD)
2.2. Catalyst characterisation O2-TPD experiments were performed in a quartz reactor
using a TCD as detector. In each analysis, 500 mg samples were
2.2.1. Specific surface area pre-treated with helium gas increasing the temperature from
The specific surface area (SSA) of the catalysts was ambient temperature up to 700 8C at 10 8C min1. The samples
calculated by the BET method from the nitrogen adsorption were oxidised with a 20% O2/He mixture at a total flow rate of
isotherms obtained at 77 K on samples outgassed at 250 8C 30 ml min1 at 700 8C for 30 min. Then, they were cooled
using a Micromeritics Accusorb 2100E apparatus. down to ambient temperature in the oxidising mixture and
flushed by a stream of purified He for 30 min. The desorption
2.2.2. X-ray diffractometry (XRD) was carried out in the same conditions as the pre-treatment,
XRD patterns were recorded at room temperature by using a maintaining the temperature at 700 8C until the baseline of the
Rigaku diffractometer operated at 30 kV and 20 mA, employ- chromatograph was stabilised.
B.P. Barbero et al. / Applied Catalysis B: Environmental 65 (2006) 21–30 23

Table 1
Specific surface areas, elemental composition from XRF (nominal values in
parentheses), and Fe4+ amount from the TPR results
Catalyst SBET (m2/g) La (%) Fe (%) Ca (%) Fe4+ (%)
LaFeO3 18.5 61.8 (57.2) 20.2 (23.0) 0 (0) 2.8
La0.8Ca0.2FeO3 13.0 56.1 (49.8) 21.9 (25.0) 3.7 (3.6) 15.9
La0.6Ca0.4FeO3 16.7 47.5 (41.0) 24.0 (27.5) 6.1 (7.9) 21.5

2.2.8. Temperature-programmed reduction (TPR)


TPR experiments were performed in the same apparatus
used for O2-TPD on the samples after the O2-TPD. The
reducing atmosphere was a 5% H2/N2 mixture at a total flow
rate of 30 ml min1. The temperature was increased at Fig. 1. XRD results of La1xCaxFeO3 perovskites and PDF data.
10 8C min1 from ambient temperature to 700 8C. The
calibration of the TCD was performed by injection of know
amounts of N2 in the H2/N2 flow, thus simulating H2 characteristic of orthorhombic symmetry (PDF 37-1493), space
consumption. group Pnma (#62). The lattice parameters estimated by the
Rietveld method are a = 5.564 Å, b = 7.849 Å, and c = 5.551 Å
2.3. Catalytic test (Table 2). The X-ray patterns of the substituted perovskites are
also monophasic and they almost show the same lines that
The catalysts (300 mg, 0.5–0.8 mm particle diameter) LaFeO3, although a XRD line shift to higher 2u angle values
diluted with glass particles of the same size in a ratio 1:5 when the calcium amount increases is observed. The lattice
were tested in a fixed-bed tubular reactor operated at parameters for the substituted samples are presented in Table 2.
atmospheric pressure. The temperature, measured with a The structure was refined assuming a rhombohedral symmetry,
coaxial thermocouple, was increased from 150 to 200 8C to space group R3c (#167).
attain total conversion of VOC (propane or ethanol). The feed The unit cell volume (normalised to account for the
was a mixture of 2:20:78 for C3H8:O2:He or 1:20.8:78.2 for difference in Z) decreases with the increase of the calcium
C2H5OH:O2:He. The total flow rate was 100 ml min1 amount (Table 2).
measured at ambient temperature. The reactants and reaction
products were alternately analysed on-line by gas chromato- 3.3. Bulk composition: X-ray fluorescence
graphy. The conversion of VOC, X%, is defined as the
percentage of VOC feed that has reacted, i.e.: XRF results are shown in Table 1. The experimentally
determined lanthanum content is higher than the expected one
ðVOCÞin  ðVOCÞout while iron amount is approx. 13% lower than the theoretical
X ð%Þ ¼  100
ðVOCÞin one in all the cases. Calcium amount detected in La0.8Ca0.2-
FeO3 is in agreement with the theoretical value but in
The specific activity (conversion/m2) is calculated as the VOC La0.6Ca0.4FeO3 the experimental result is quite lower than the
conversion per surface area unit. theoretical value.
Data obtained under differential conditions with propane
conversions below 10% were used in the calculus of the
reaction rates. The apparent activation energies were estimated Table 2
from Arrhenius type plots (ln r versus 1/T). Lattice parameters for La1xCaxFeO3 perovskite oxides
Sample a (Å) b (Å) c (Å) a (8) V (Å3)/Z
3. Results
LaFeO3 5.564 7.849 5.551 – 61
La0.8Ca0.2FeO3 5.524 – – 60.1 60
3.1. Specific surface area La0.6Ca0.4FeO3 5.490 – – 60.2 59

Table 1 shows the specific surface areas measured by the


BET method. The specific surface areas of LaFeO3 and Table 3
La0.6Ca0.4FeO3 are similar within the experimental error XPS results
whereas that of La0.8Ca0.2FeO3 is lower. Binding energies (eV) Atomic ratios
La 3d5/2 Ca 2p Fe 2p La/Fe Ca/Fe Ca/(La + Ca)
3.2. X-ray diffraction
LaFeO3 833.8 709.9 2.22 0 0
La0.8Ca0.2FeO3 833.5 347.0 710.1 1.40 0.54 0.28
XRD results revealed the formation of perovskite-type
La0.6Ca0.4FeO3 833.2 346.4 710.0 1.05 0.83 0.44
single phases (Fig. 1). The diffractogram of LaFeO3 sample was
24 B.P. Barbero et al. / Applied Catalysis B: Environmental 65 (2006) 21–30

3.4. Surface composition: X-ray photoelectron 107 cm1 (medium), 113 cm1 (medium), 277 cm1 (weak),
spectroscopy 432 cm1 (very strong), 455 cm1 (shoulder), 497 cm1
(medium), 608 cm1 (medium), 711 cm1 (strong),
Table 3 shows the XPS results. La 3d5/2 peak is detected at 739 cm1 (weak), 918 cm1 (very weak), and 1108 cm1
833.5  0.3 eV. According to the reported [12–14], this (weak). Bontempi et al. [17] obtained a very similar spectrum
corresponds to La3+. Ca 2p peak is at 346.7  0.3 eV and it for LaNiO3 perovskite, which has rhombohedral structure. On
is Ca2+ [13]. Fe 2p peak appears at 710.0  0.1 eV. This the other hand, our spectrum could also be compared with that
binding energy would be characteristic of Fe3+ [13,14]. Binding of La0.5Ca0.5MnO3 recorded at 8 K by Granado et al. [18] who
energy shifts due to calcium substitution are not observed. assigned it orthorhombic symmetry. The assignment to one or
(La + Ca)/Fe atomic ratios almost double the theoretical value. other symmetry by comparison with the previously published
Furthermore, Ca/(La + Ca) atomic ratio is slightly higher than data is not straightforward, because of the different experi-
the theoretical one. This indicates a surface enrichment in mental conditions (laser line, laser power, etc.) used in each
lanthanum and calcium. work. However, it is worth noting in the Raman spectra of the
substituted samples, the band intensities decrease when calcium
3.5. Raman spectroscopy amount increases.

Raman spectra of La1xCaxFeO3 perovskites are shown in 3.6. FT-IR spectroscopy


Fig. 2a (spectra acquired with Jasco spectrometer) and Fig. 2b
(by using LabRam Instruments). The spectra obtained with Fig. 3 presents the IR spectra of the La1xCaxFeO3
Jasco spectrometer (Fig. 2a) are not well defined. Very few perovskites. All of them show two strong and well-defined
Raman spectra for LaFeO3 perovskite have been found in absorption bands, typical of perovskite oxides ABO3 [19–21].
literature. Some of them, similar to ours, have been reported by Because the Fe–O bonds of the BO6 octahedral units are
Popa et al. [15,16] and Bontempi et al. [14]. According to their undoubtedly stronger than those of the 12-coordinated La(III)–
assignments, the crystal symmetry of LaFeO3 is orthorhombic. O units, one may predict that the BO6 units dominate the
In the Raman spectra of the substituted samples, the band at spectroscopic behaviour [21]. The bands at higher wavenum-
417 cm1 disappears as well as the other weak ones and the bers (500–730 cm1) are assigned to the stretching modes of
band at 635 cm1 decreases and widens.
The spectra shown in Fig. 2b are very different from those in
Fig. 2a. LaFeO3 spectrum presents well-defined bands at

Fig. 2. Raman spectra acquired with (a) Jasco spectrometer and (b) LabRam
Instruments. Fig. 3. FT-IR spectra of La1xCaxFeO3 perovskites.
B.P. Barbero et al. / Applied Catalysis B: Environmental 65 (2006) 21–30 25

the BO6 octahedra and those at lower wavenumbers (250–


500 cm1), to the deformation modes of these same polyhedra
[19].
IR spectrum of LaFeO3 presents an intense band around
567 cm1 with a shoulder at 623 cm1. The main band can be
assigned to antisymmetric stretching vibration of the BO6
octahedra [20,21] whereas the shoulder could be assigned to the
symmetric stretching vibration of these octahedra. The bands in
the low-wavenumbers region (250–500 cm1) correspond to
the deformational modes of the octahedra. Bands associated to
a La(III)–O stretching motion can also appear in this region
[21]. Couzi and Huong [20] reported a similar spectrum of
LaFeO3.
In the IR spectra of calcium-substituted samples, the same
main bands that those for LaFeO3 are found although a slight
shift towards higher wavenumbers when the calcium amount
increases is observed. Furthermore, the shoulders at 623 and
453 cm1 are undetected in the substituted samples and a band
around 877 cm1 appears whose intensity increases when the
calcium amount increases.

3.7. Temperature-programmed reduction

Fig. 4 shows the curves of the TPR experiments. TPR profile


of LaFeO3 shows two very weak signals with the maxima at
about 255 and 385 8C. Ciambelli et al. [8,22] and Spinicci et al.
[23] reported a similar curve. TPR profiles of substituted
samples show an intense reduction signal at about 355 8C and
Fig. 5. O2-TPD results.

another weak signal at 410 8C. The intensity of the first signal
increases when the calcium amount increases. A further
reduction starts at temperature above 620 8C and although it is
not completely resolved under the experimental conditions
used in this work, the signal intensity increases with the
calcium substitution.

3.8. Temperature-programmed desorption of oxygen

Fig. 5 shows the O2-TPD results. The curves present a first


peak between 150 and 350 8C whose maximum shifts towards
higher temperatures when the calcium amount increases.
Furthermore, a second signal above 400 8C is observed only for
the substituted samples. Nitadori and Misono [4] observed
curves with the same shape and trend in their O2-TPD results of
La1xSrxFeO3 perovskites. The first signal is generally denoted
as a oxygen species and they correspond to oxygen species
adsorbed on surface oxygen vacancies. The second signal at
higher temperature is known as b oxygen species. These
oxygen species release from the oxide lattice [24–26].

3.9. Catalytic activity

The results are presented in Figs. 6–8 and Table 4. Catalytic


performance was evaluated using two different molecules:
propane and ethanol. The main products during the total
Fig. 4. TPR profiles of La1xCaxFeO3 perovskites. oxidation of propane were CO2 and water. At reaction
26 B.P. Barbero et al. / Applied Catalysis B: Environmental 65 (2006) 21–30

Fig. 8. Specific activity (VOC conversion per surface area unit) as a function of
Fig. 6. Catalytic activity in total oxidation of propane (solid line, closed the reaction temperature. Identification as in Fig. 6.
symbols) and ethanol (dotted line, open symbols) on (^) LaFeO3, ( )
La0.8Ca0.2FeO3, and ( ) La0.6Ca0.4FeO3.

Table 4
Catalytic activity results in propane oxidation
Catalyst Tonseta (8C) T50b (8C) T95b (8C) Eac (kcal/mol) rwd (mmol g1 h1) rsd (mmol m2 h1)
LaFeO3 225 385 428 26.2 76.96 4.16
La0.8Ca0.2FeO3 220 392 445 22.9 88.93 6.84
La0.6Ca0.4FeO3 220 378 420 23.4 108.75 6.51
a
Temperature of reaction start.
b
Temperatures corresponding to 50% and 95% conversion, respectively.
c
Apparent activation energies.
d
Reaction rates per catalyst mass unit and per surface area unit, respectively, at reaction temperature = 300 8C.

temperatures higher than 370 8C, propylene was also detected with the increase of the calcium amount (Fig. 6). At low
and above 420 8C, traces of CO and C2H4 were found. Other reaction temperatures, all the catalysts have a similar behaviour
oxygenated products such as acrolein, acrylic acid, acetaldehyde during propane oxidation and when the reaction temperature
were not found. The total oxidation of ethanol occurs through an increases, different catalytic performance is observed with
intermediate, acetaldehyde. In Fig. 7, ethanol conversion and different calcium amount.
selectivities to acetaldehyde and CO2 as a function of the reaction Considering the reaction temperature corresponding to 50%
temperature on LaFeO3 are shown. The curves for the other and 95% conversion of VOC (T50 and T95, Table 4), the activity
catalysts have the same shape that for LaFeO3. The carbon for propane oxidation increases in the order: La0.8Ca0.2-
balance was closed in 5% for both reactions. FeO3 < LaFeO3 < La0.6Ca0.4FeO3 while for ethanol oxidation
Propane oxidation starts at around 220 8C on all the catalysts the activity increases with increasing x. However, when the
while ethanol oxidation begins at lower reaction temperature VOC conversion per surface area unit (VOC conversion/m2) is
compared (Fig. 8), the substituted samples show performances
very similar among them and higher than that of LaFeO3, both
for propane and ethanol oxidation.
By analysing the kinetic parameters (apparent activation
energies and reaction rates, Table 4), the best catalytic
behaviour of the substituted perovskites against LaFeO3
perovskite is also observed.

4. Discussion

4.1. On the synthesis and characterisation of


La1xCaxFeO3 perovskites

La1xCaxFeO3 perovskites with x = 0, 0.2, and 0.4 were


Fig. 7. Ethanol conversion and selectivities to acetaldehyde and CO2 as a prepared by the citrate method which is recognised as an
function of the reaction temperature on LaFeO3. appropriate method to obtain pure solids since the precursor gel
B.P. Barbero et al. / Applied Catalysis B: Environmental 65 (2006) 21–30 27

is highly homogeneous. Furthermore, specific surface areas structure distortion. However, the perovskite structure can
adequate for materials to be used as catalysts in oxidation change due to either stoichiometry modifications as well as by
reactions are obtained since considerably lower calcination the effect of thermal treatments or changes of the oxygen partial
temperatures than those employed in ceramic methods are pressure in the surrounding atmosphere [20]. Moreover, as it
sufficient to obtain the suitable phase. The perovskites can be deduced from the comparison of the spectra obtained
synthesised in this work have specific surface areas between with two different spectrometers (Fig. 2a and b), the results
13 and 18 m2/g (Table 1). Similar SSA (19 m2/g) was obtained depend strongly on the experimental conditions employed in
by Daturi et al. [27] for LaFeO3 prepared as aerogels via the so- the measurements. While the spectrum shown in Fig. 2a can be
called supercritical drying technique. In contrast, Ciambelli assigned to an orthorhombic structure [15–18], the spectrum of
et al. [8] prepared a series of La1xCaxFeO3 perovskites with Fig. 2b could correspond to a rhombohedral [17] or
specific surface areas much lower than those obtained by us (3– orthorhombic structure [18]. It has been reported in the
6 m2/g range). Although the same preparation method was literature [28] that the orthorhombic phase is the stable phase at
employed, they calcined at 800 8C (100 8C higher than us). This ambient temperature for lanthanum orthoferrite and phase
could be the reason for the decrease of the specific surface area. transition to rhombohedral structure occurs at about 977 8C.
As it has been shown in the study on La1xCaxCoO3 perovskites Probably, a local overheating of the sample by laser effect
[9], an increase of the calcination temperature from 700 to caused the phase change of the samples analysed with the
800 8C strongly reduces the specific surface area. Another LabRam Instrument (Fig. 2b). Further on these differences, it is
difference with regard to our samples is the SSA variation with important to note that in both cases (Fig. 2a and b) the band
the calcium content. Ciambelli et al. [8] found that the specific intensity decreases when the calcium amount increases. This
surface areas increased for low x values (x = 0.1–0.2) and then could be due to the disorder induced by the calcium
decreased but we observed the contrary tendency. substitution. Considering that a cubic symmetry perovskite
The synthesis success was corroborated by XRD (Fig. 1) should not show any Raman bands [18,20], a nearly cubic
where a single perovskite phase was observed for all the symmetry in the calcium-substituted perovskites would be
samples. No segregated phase or contamination was detected. obtained confirming the one observed by XRD.
This confirms how appropriate is the synthesis method for In summary, La1xCaxFeO3 perovskites with adequate
avoiding the phase segregation and obtaining pure phases. specific surface area to be applied in oxidation reactions can
Nevertheless, the presence of other phase(s) in very low be obtained by a simple method such as the citrate method,
concentration and/or as amorphous phase cannot be discarded controlling the calcination temperature. The catalysts are single
because of the limitations of the XRD technique. perovskite phases with a slight surface enrichment in
The elemental composition results from XRF and XPS lanthanum and calcium. The calcium substitution in the A
indicate lanthanum excess for all of the samples. The excess sites of the LaFeO3 perovskite produces a structural change
detected by XPS is a bit higher than that by XRF. Considering from the orthorhombic structure towards a less distorted one.
that the result from XRF is related to the bulk composition Furthermore, XRD and Raman spectroscopy results presented
while that from XPS is about the surface composition, some here clearly show that the perovskite structure changes by
lanthanum-rich phase could exist on the surface. Another composition modifications as well as by exposition at high
relevant insight is that the (La + Ca)/Fe surface atomic ratio is temperatures. The study of the surrounding atmosphere on the
almost constant for all of the catalysts. This indicates that there perovskite structure is beyond the aim of this work. However,
is no substitution limit of calcium for lanthanum in we can alert about how careful we should be when correlating
La1xCaxFeO3 perovskites, at least when x varies between 0 the crystalline structure of a perovskite with catalytic activity,
and 0.4. Thus, the obtained catalysts should consist mainly of since it is expected that the reaction atmosphere (especially in
the single perovskite phase with a surface enrichment in the case of oxidation reactions) can modify the crystalline
lanthanum and calcium. structure.
The X-ray diffractograms of La1xCaxFeO3 perovskites
studied in this work show that the diffraction lines are broaden 4.2. Effect of calcium on the electroneutrality
and their intensities decrease when the calcium amount of the perovskites
increases. This makes that the less intense diffraction lines
become undetected. The prevailing lines in the La0.6Ca0.4FeO3 The calcium substitution causes a notable change on the
perovskite are in agreement with those of the PDF 75-0541 of lattice parameters determined by Rietveld method (Table 2).
the cubic LaFeO3 perovskite, although they are shifted to The shift of diffraction lines to higher 2u angle values when the
higher 2u angles. Then, the calcium substitution could induce calcium amount increases, indicates a lower interplanar
changes in the crystalline structure modifying the orthorhombic distance and, consequently, a decrease of the cell volume.
structure tending to a nearly cubic structure. The structure This is corroborated by the results of unit cell volumes
refinement by means of Rietveld method confirms that the (Table 2). The ionic radius of Ca2+ 12-coordinated (1.34 Å)
substituted perovskite structure are less distorted than that of [29] is very similar to that of La3+ 12-coordinated (1.36 Å) [29];
LaFeO3. so, the cell volume should be very slightly affected by the
The structural modification should be easily detected by calcium substitution. Nevertheless, since calcium has lower
Raman spectroscopy, which is a very sensitive technique to oxidation state than lanthanum, an increase of the iron
28 B.P. Barbero et al. / Applied Catalysis B: Environmental 65 (2006) 21–30

oxidation state from Fe3+ to Fe4+ can occur or oxygen vacancies can be assigned to the total reduction of iron towards the
can be generated to maintain the structure electroneutrality. The metallic state. This deep reduction would be accomplished by
Fe4+ formation in substituted or unsubstituted AFeO3 the destruction of the perovskite phase. Due to the intensity of
perovskites has been reported [22]. The ionic radius of Fe4+ that signal increases when the calcium content increases, it is
(0.585 Å) is smaller than that of Fe3+ (0.645 Å); then, this could possible to suggest that the calcium substitution decreases the
be the reason for the cell volume decrease. stability of the perovskite structure. This can be also deduced
The FT-IR results are in line with these observations. from the FT-IR results since the shift of the bands toward higher
Particularly, the FT-IR band intensity appearing at around wavenumbers (Fig. 2) can be attributed to a weakening of the
877 cm1 increases when the calcium amount increases. In metal–oxygen bonds.
La1xSrxCoO3 perovskites, the appearance of a new band at A decisive measurement to confirm the Fe4+ formation can
around 670 cm1 whose intensity increased with the increase of be obtain by using Mössbauer spectroscopy (work in progress).
strontium ions content was observed by Berger et al. [30]. They The singlet with isomeric shift of 0.12 mm/s detected in the
assigned this band to the increase of Co–O constant force as a spectra at 298 K is in good agreement with a signal of Fe4+.
result of charge increase of cobalt ions (Co4+ ions formation) to In summary, the XRD, FT-IR and TPR results indicate that
conserve the electroneutrality of the perovskite structure. the electronic debalance caused in the perovskites by the partial
According to this, we can assign the band at 877 cm1 to substitution for lanthanum (3+) with calcium (2+) is
structural modifications produced by the substitution with compensated by the oxidation of a part of Fe3+ ions to Fe4+.
calcium and probably associated to the Fe4+ ions formation. Taking into account that the ionic radius of Fe4+ is smaller than
The results from TPR also indicate the Fe4+ formation. The that of Fe3+ and that La3+ and Ca2+ have similar ionic radii, it is
curves shown in Fig. 4 represent the hydrogen consumption due expected that theptolerance
ffiffiffi factor defined by Goldschmidt [31]
to sample reduction. To determine the final state of the sample, t ¼ ðRA þ RO Þ= 2ðRB þ RO Þ, decreases when the calcium
when the reduction finished, the reducing gaseous mixture was substitution increases. The perovskite structure exists in oxides
shifted to helium and the sample was cooled down to ambient within the t range 0.75 < t < 1 [2]. The value 1 is for cubic
temperature in helium flow. Immediately after, it was analysed perovskite and it decreases when the structure is distorted.
by XRD and it was observed that the perovskite phase was Then, the increase of value t when increasing the calcium
preserved. Only in the diffractogram of the reduced La0.6Ca0.4- substitution level would indicate a tendency of the structure
FeO3 sample, the most intense peak of metallic iron was weakly towards a nearly cubic one, confirming that previously
detected. Furthermore, a slight peak shift towards lower 2u observed by XRD and Raman.
angles was found (Fig. 9). This indicates that the reduction If the increase of the iron oxidation state is sufficient to
causes an expansion of the unit cell volume which is predictable conserve the perovskite electroneutrality, the formed Fe4+
from the difference between the ionic radii of Fe4+ and Fe3+. amount would increase linearly with the calcium amount. In
These results conduce us to assign the two first signals to the order to verify this, the Fe4+ amount was estimated from the
reduction of Fe4+ to Fe3+. Ciambelli et al. [8] observed two TPR results (Table 1) and it was plotted as a function of the
overlapped peaks in the TPR curves of La1xCaxFeO3 calcium amount (Fig. 10). Furthermore, considering that
perovskites and they assigned the shoulder at lower temperature the shift of the XRD lines is proportional to the Fe4+ amount,
to more easily reducible surface Fe4+ ions. The good resolution the position of the XRD main peak was also plotted as a
of the two peaks in our TPR profiles induce us to propose a function of the calcium amount in Fig. 10. As it can be seen, the
reduction of Fe4+ to Fe3+ in two steps: Fe4+ ! Fe4+/ relationship is not linear. Even if it is considered that
Fe3+ ! Fe3+ rather splitting due to a diffusion phenomena. the calcium amount determined by XRF is lower than the
The reduction step starting at 620 8C, which is not completely theoretical value, the relationship is not linear. Then, it is
resolved under the experimental conditions used in this study, predictable that besides Fe4+, oxygen vacancies are formed to

Fig. 10. Position of the XRD main peak, Fe4+ amount from TPR results and
Fig. 9. XRD patterns of La0.6Ca0.4FeO3 perovskite both fresh and after TPR. area under the first O2-TPD peak per surface area unit as a function of the
(*) metallic Fe. calcium amount.
B.P. Barbero et al. / Applied Catalysis B: Environmental 65 (2006) 21–30 29

maintain the electroneutrality, especially when the substitution vacancies play an important role, favourable for the oxidation
level increases. process because they accelerate the dissociation of oxygen
A simple and direct way to determinate oxygen vacancies is molecule on the surface and also, increase the mobility of
by means of O2-TPD experiments. The first peak of the curves lattice oxygen.
shown in Fig. 5 is generally assigned to oxygen adsorbed on To identify which is the phenomenon governing the reaction,
surface oxygen vacancies. Thus, the area under the curve of this we intent to find a linearity between the catalytic behaviour and
signal would be a measure of surface oxygen vacancies amount. the perovskite characterisation results. In Fig. 6 and Table 4 it is
The first signal intensity increases in the order La0.8Ca0.2- observed that the calcium substitution modifies the perfor-
FeO3 < LaFeO3 < La0.6Ca0.4FeO3. However, considering that mance of the LaFeO3 perovskite but in a different way if the
they are ‘‘surface vacancies’’ it would be more convenient to do oxidised molecule is a hydrocarbon (propane) or an alcohol
the comparison taking into account the surface area of the (ethanol). In the propane oxidation, the conversion increases
samples. When the first signal intensity was divided by the in the order La0.8Ca0.2FeO3 < LaFeO3 < La0.6Ca0.4FeO3
surface area, an increase of the surface oxygen vacancy amount whereas the ethanol conversion increases when the calcium
when increasing the calcium amount was observed (Fig. 10). amount increases. Taking into account that the Fe4+ ion amount
This indicates that the electronic debalance caused by the also increases when the calcium amount increases, we can
calcium partial substitution for lanthanum is formerly suggest that the ethanol oxidation proceeds through a redox
compensated by the Fe4+ ion formation and, when the calcium mechanism where the par Fe4+/Fe3+ plays an important role. In
amount increases, oxygen vacancies are also formed. contrast, the propane oxidation could be governed by another
A TPD signal such as that observed above 350 8C, which is mechanism in which the surface oxygen vacancy concentration
not completely resolved under the experimental conditions is relevant. According to the O2-TPD results, the intensity of the
used in this work, has also been observed by Nitadori and a-oxygen signal follows the same order that the propane
Misono [4]. They assigned it to partial reduction of Fe4+ to Fe3+. conversion (La0.8Ca0.2FeO3 < LaFeO3 < La0.6Ca0.4FeO3).
In our experiment, it is not very probable that the second signal However, when the VOC specific conversion (conversion
corresponds to Fe4+ reduction. We carried out the temperature- per surface area unit) is considered, very similar results are
programmed reduction following the TPD after cooling down obtained with both molecules. Furthermore, the calcium-
the sample to ambient temperature under helium flow without substituted samples are more active than the unsubstituted
reoxidation. Since the TPR results shown a clear reduction perovskite but no difference is observed with several
signal, we discarded the possibility that the second TPD peak be substitution levels. This tendency is coherent with the apparent
associated to the Fe4+ reduction. Then, we agree with other activation energy (Ea) values. Ea for LaFeO3 perovskite in
several authors [24–26] who have assigned that signal to propane oxidation was 26.2 kcal/mol while for La0.8Ca0.2FeO3
oxygen coming from the catalyst bulk and probably associated and La0.6Ca0.4FeO3 perovskites were 22.9 and 23.4 kcal/mol,
to inner oxygen vacancies. The desorption of these species respectively. Existing no steric limitations due to the model
could be completed at much lower temperatures during the molecule sizes, it can be assumed that the calcium substitution
temperature-programmed reduction. In fact, a very weak signal generates complex active sites. Probably, these sites are related
below 200 8C is observed in the TPR curves. This signal could to Fe4+ ions that, as discussed above, are generated to conserve
be due to oxygen release not associated to iron reduction. The the electroneutrality but not further than a certain amount.
amount of these vacancies is difficult to estimate but an increase However, the site concentration would not only depend on the
of the intensity when the calcium amount increases is observed. Fe4+ ions content. The catalyst capacity to complete the redox
In conclusion, the partial substitution for lanthanum by cycle by means of the oxygen activation to re-oxidise itself also
calcium in La1xCaxFeO3 perovskites causes an oxidation state plays an important role. The equilibrium between these factors
increase of a part of the iron ions and at higher substitution seems to be determinant for the improvement of the catalytic
levels (x > 0.2) oxygen vacancies are also formed. activity. Probably, thus can be explained the methane
combustion results obtained by Ciambelli et al. [8].
4.3. On the catalytic performance of the La1xCaxFeO3
perovskites
5. Conclusions
It is widely known that the catalytic activity of the perovskite
oxides is related to unusual oxidation states of the transition La1xCaxFeO3 perovskite-type oxides have been obtained
metal ions, the amount of non-stoichiometric oxygen and the by means of a very simple method, such as the citrate method,
structural defects of lattice. Depending on the nature of the with adequate specific surface area for oxide catalysts to be
VOC molecule that wants to be oxidised, one of these aspects used in oxidation reactions, simply controlling the calcination
will be more relevant for the catalytic performance and will temperature. The catalysts are single perovskite phases with a
give an idea about how the reaction occurs. The increase of the slight surface enrichment in lanthanum and calcium. The
iron oxidation state could improve the catalyst performance in a calcium substitution in the A sites of the LaFeO3 perovskite
redox cycle, in which the iron is reduced supplying the produces a structural change from the orthorhombic structure
necessary oxygens for the oxidation reaction and then it is towards a less distortioned one. Furthermore, XRD and Raman
reoxidized taking oxygen from the gas phase. The oxygen spectroscopy results clearly show that the perovskite structure
30 B.P. Barbero et al. / Applied Catalysis B: Environmental 65 (2006) 21–30

changes by composition modifications as well as by exposition [8] P. Ciambelli, S. Cimino, L. Lisi, M. Faticanti, G. Minelli, I. Pettiti, P. Porta,
Appl. Catal. B 33 (2001) 193.
at high temperatures.
[9] N.A. Merino, B.P. Barbero, P. Grange, L.E. Cadús, J. Catal. 231 (2005)
The structure electroneutrality of La1xCaxFeO3 perovskites 232.
is achieved by an oxidation state increase of a part of the iron [10] P. Courty, H. Ajot, C. Marcilly, B. Delmon, Power Technol. 7 (1973) 21.
ions from Fe3+ to Fe4+ and at higher substitution levels [11] R.A. Young, A. Sakthivel, T.S. Moss, C.O. Paiva-Santos, J. Appl. Cryst. 28
(x > 0.2), oxygen vacancies are also formed. (1995) 366.
Finally, an improvement of the catalytic activity in propane [12] T.L. Barr, J. Phys. Chem. 82 (16) (1978) 1801.
[13] C.D. Wagner, W.M. Riggs, L.E. Davis, J.F. Moulder, in: G.E. Muilenberg
and ethanol combustion was observed on the substituted (Ed.), Handbook of X-ray Photoelectron Spectroscopy, first ed., Perkin-
perovskites. Correlating this with the characterisation results, Elmer Corporation (Physical Electronics Division), 1979.
the active sites would be associated to the Fe4+ ions. [14] D. Briggs, M.P. Seah, 2nd ed., Practical Surface Analysis, vol. 1, John
Willey & Sons, 1993.
[15] M. Popa, J. Frantti, M. Kakihana, Solid State Ion. 154/155 (2002)
Acknowledgements
135.
[16] M. Popa, J. Frantti, M. Kakihana, Solid State Ion. 154/155 (2002) 437.
The financial support from CONICET, Universidad Nacio- [17] E. Bontempi, C. Garzella, S. Valetti, L.E. Depero, J. Eur. Ceram. Soc. 23
nal de San Luis, Fundación Antorchas and SeCyT (Argentina), (2003) 2135.
FNRS (Belgium) cooperation project is gratefully acknowl- [18] E. Granado, N.O. Moreno, A. Garcı́a, J.A. Sanjurjo, C. Rettori, I. Torriani,
S.B. Oseroff, J.J. Neumeier, K.J. McClellan, S.W. Cheong, Y. Tokura,
edged. The Raman spectra (a) were acquired thanks to a grant
Phys. Rev. B 58 (17) (1998) 11435.
of Japan International Cooperation Agency to CENACA [19] S.D. Ross, Inorganic Infrared and Raman Spectra, McGraw-Hill, London,
(Argentina). The authors also thank to Pierre Eloy (Université 1972.
Catholique de Louvain, Louvain-la-Neuve, Belgium) for his [20] M. Couzi, P.V. Huong, Ann. Chim. 9 (1974) 19.
help with the XPS analysis. [21] A.E. Lavat, E.J. Baran, Vibrat. Spectrosc. 32 (2003) 167.
[22] P. Ciambelli, S. Cimino, S. De Rossi, L. Lisi, G. Minelli, P. Porta, G.
Russo, Appl. Catal. B 29 (2001) 239.
References [23] R. Spinicci, A. Tofanari, A. Delmastro, D. Mazza, S. Ronchetti, Mater.
Chem. Phys. 76 (2002) 20.
[1] R.L. Garten, R.A. Dalla Betta, J.C. Schlatter, in: G. Ertl, H. Knozinger, J. [24] R.J.H. Voorhoeve, J.P. Remeika, P.E. Freeland, B.T. Mattias, Science 177
Wertkamp (Eds.), Handbook of Heterogeneous Catalysis, vol. 4, VCH, (1972) 353.
Weinheim, Germany, 1998, p. 1668. [25] N. Yamazoe, Y. Teraoka, T. Nakamura, Chem. Lett. (1981) 1767.
[2] L.G. Tejuca, J.L.G. Fierro, Adv. Catal. 36 (1989) 237. [26] T. Seiyama, N. Yamazoe, K. Eguchi, Ind. Eng. Chem. Prod. Res. Dev. 24
[3] L. Simonot, F. Garin, G. Maire, Appl. Catal. B 11 (1997) 167. (1985) 19.
[4] T. Nitadori, M. Misono, J. Catal. 93 (1985) 459. [27] M. Daturi, G. Busca, R.J. Willey, Chem. Mater. 7 (1995) 2115.
[5] M.A. Alario Franco, M.J.R. Henche, M. Vallet, J.M. González Calbet, J.C. [28] L. Bornstein, Numerical Data and Functional Relationships in Science and
Grenier, A. Wattiaux, P. Hagenmuller, J. Solid State Chem. 46 (1983) 23. Technology, vol. III/4a, Springer-Verlag, Berlin, 1970.
[6] M.A. Alario Franco, J.M. González Calbet, M. Vallet Regi, J.C. Grenier, J. [29] R.D. Shannon, Acta Cryst. A 32 (1976) 751.
Solid State Chem. 49 (1983) 219. [30] D. Berger, V. Fruth, I. Jitaru, J. Schoonman, Mater. Lett. 58 (2004) 2418.
[7] M. Vallet Regi, J. González Calbet, M.A. Alario Franco, J.C. Grenier, P. [31] V.M. Goldschmidt, Skr. Nor. Videnk-Akad., Kl. 1: Mat.-Naturvidensk. Kl.
Hagenmuller, J. Solid State Chem. 55 (1984) 251. No. 8 (1926).

You might also like