You are on page 1of 11

Applied Catalysis B: Environmental 24 (2000) 243–253

AMnO3 (A=La, Nd, Sm) and Sm1−x Srx MnO3 perovskites as


combustion catalysts: structural, redox and catalytic properties
Paolo Ciambelli b , Stefano Cimino a , Sergio De Rossi c , Marco Faticanti c , Luciana Lisi d ,
Giuliano Minelli c , Ida Pettiti c , Piero Porta c,∗ , Gennaro Russo a , Maria Turco a
a Dipartimento di Ingegneria Chimica, Università ‘Federico II’, Napoli, Italy
b Dipartimento di Ingegneria Chimica e Alimentare, Università di Salerno, Salerno, Italy
c Centro di Studio del CNR su ‘Struttura e Attività Catalitica di Sistemi di Ossidi’ (SACSO), c/o Dipartimento di Chimica, Università La
Sapienza, Piazzale A. Moro 5, 00185 Rome, Italy
d Istituto di Ricerche sulla Combustione, CNR, c/o Dipartimento di Ingegneria Chimica, Piazzale V. Tecchio 80, 80125 Napoli, Italy

Received 13 June 1999; received in revised form 3 September 1999; accepted 3 September 1999

Abstract
Catalytic combustion of methane has been investigated over AMnO3 (A = La, Nd, Sm) and Sm1−x Srx MnO3 (x = 0.1,
0.3, 0.5) perovskites prepared by citrate method. The catalysts were characterized by chemical analysis, XRD and TPR
techniques. Catalytic activity measurements were carried out with a fixed bed reactor at T = 623–1023 K, space veloc-
ity = 40 000 N cm3 g−1 h−1 , CH4 concentration = 0.4% v/v, O2 concentration = 10% v/v.
Specific surface areas of perovskites were in the range 13–20 m2 g−1 . XRD analysis showed that LaMnO3 , NdMnO3 ,
SmMnO3 and Sm1−x Srx MnO3 (x = 0.1) are single phase perovskite type oxides. Traces of Sm2 O3 besides the perovskite phase
were detected in the Sm1−x Srx MnO3 catalysts for x = 0.3, 0.5. Chemical analysis gave evidence of the presence of a significant
fraction of Mn(IV) in AMnO3 . The fraction of Mn(IV) in the Sm1−x Srx MnO3 samples increased with x. TPR measurements
on AMnO3 showed that the perovskites were reduced in two steps at low and high temperature, related to Mn(IV) → Mn(III)
and Mn(III) → Mn(II) reductions, respectively. The onset temperatures were in the order LaMnO3 > NdMnO3 > SmMnO3 . In
Sm1−x Srx MnO3 the Sr substitution for Sm caused the formation of Mn(IV) easily reducible to Mn(II) even at low temperature.
Catalytic activity tests showed that all samples gave methane complete conversion with 100% selectivity to CO2 below 1023 K.
The activation energies of the AMnO3 perovskites varied in the same order as the onset temperatures in TPR experiments
suggesting that the catalytic activity is affected by the reducibility of manganese. Sr substitution for Sm in SmMnO3 perovskites
resulted in a reduction of activity with respect to the unsubstituted perovskite. This behaviour was related to the reduction
of Mn(IV) to Mn(II), occurring under reaction conditions, hindering the redox mechanism. ©2000 Elsevier Science B.V. All
rights reserved.
Keywords: Catalytic methane combustion; Perovskites; Redox properties

1. Introduction

Catalytic combustion has been proposed as a


∗Corresponding author. Fax: +39-3-6-490-324. method for promoting the effective oxidation of
E-mail address: porta@axrma.uniromal.it (P. Porta). fuel/air lean mixtures with low emissions of NOx , CO

0926-3373/00/$ – see front matter ©2000 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 6 - 3 3 7 3 ( 9 9 ) 0 0 1 1 0 - 1
244 P. Ciambelli et al. / Applied Catalysis B: Environmental 24 (2000) 243–253

and unburnt hydrocarbons [1,2]. Therefore, by this limiting to some extent their application in catalysis.
process neither potentially harmful emissions are pro- In order to improve the catalytic activity it is thus nec-
duced nor expensive secondary treatments required essary to produce such materials with higher surface
to meet the stringent demands on atmospheric pollu- areas using suitable precursors which may give, under
tion limitation [3]. Catalytic combustion operates at mild heat treatment, the desired catalysts. For this pur-
low temperature compared to the conventional flame pose the citrate method was proposed by Zhang et al.
combustion thus reducing the formation of thermal [13] to prepare high surface area perovskites. This
NOx starting at about 1773 K [1,4]. method involves addition of citric acid to the precursor
Catalytic materials to be used in the combustor must nitrate solution. From citrate complexes, through sev-
posses several properties such as high surface area, eral decomposition steps leading to the elimination of
thermal stability, high durability with respect to oxi- the residual CO3 2− and NO3 − ions, perovskite struc-
dation activity and product selectivity. Noble metals ture is obtained. It was found that this method allows
and metal oxides have been reported as active cat- to obtain more homogeneous dispersion of the pre-
alysts for methane total oxidation [1]. Noble metals cursor salts. Therefore, a lower calcination tempera-
based catalysts are the most active ones even at low ture than other methods is needed to obtain perovskite
temperature, but they are expensive and their thermal structure, thus avoiding sintering phenomena.
stability is poor, due to sintering and volatility. On This paper reports on the catalytic activity of
the other hand transition metal oxides are cheaper, but large surface area AMnO3 (A = La, Nd, Sm, Sr)
less active and undergo sintering at moderate temper- perovskites, prepared by citrate precursors, in the
ature [1,2,5]. Much attention has been paid recently combustion of methane, with the aim to understand
to perovskite-type oxides, of general formula ABO3 the effect of the rare earth A cation on their catalytic
(where A and B are usually rare earth and transition behaviour. Structural and redox properties were also
metal cations, respectively), as catalysts for total oxi- investigated.
dation of hydrocarbons, due to their high activity and
thermal stability [1,2,6].
Many metallic elements are stable in the ABO3 2. Experimental
perovskite structure provided that their cationic
radii fit well the sizes of the 12-coordinated A LaMnO3 , NdMnO3 and Sm1−x Srx MnO3 (x = 0.0,
and 6-coordinated B sites, e.g. rA > 0.90 Å, and 0.1, 0.3, 0.5) catalysts were prepared according to the
rB > 0.51 Å. Moreover, the high stability of the per- method described by Zhang et al. [13]. A concentrated
ovskite structure allows the partial substitution of solution of metal nitrates was mixed with an aqueous
either A and B cations by other metals with differ- solution of citric acid. The molar ratio of citric acid
ent oxidation state and the consequent generation of to total metal cations was fixed at unity. Water was
structural defects (e.g. anionic or cationic vacancies) evaporated from the mixed solution at 343 K and the
[7–9]. The effect of the nature of the B cation on sol was further dehydrated to yield a solid amorphous
the physico-chemical and catalytic properties of lan- citrate precursor. The product was ground and fired
thanum based perovskites has been widely studied, at 573 K for 1 h, then reground and calcined for 5 h
perovskites containing manganese or cobalt having at 823 K. The final calcination, after regrinding, was
been found the most active in methane combustion performed for 5 h at 1073 K.
[10]. It was reported that the substitution of B by lower The content of all metals (Table 1) was determined
valence cations in LaMnO3 and LaCoO3 perovskites by inductively coupled plasma emission spectroscopy
generally promotes methane oxidation [1], whereas (ICP). To determine the relative concentration of
the effect of rare earth has been less investigated, only Mn(III) and Mn(IV), part of the sample was dissolved
few studies concerning ACoO3 perovskites having in a known excess of standard ammonium-ferrous
been reported [11,12]. (Mohr’s salt) sulfate solution acidified with sulfuric
Perovskite oxides are generally prepared by ceramic acid. The excess of Fe(II) was then titrated with stan-
methods that require very high temperature and pro- dard potassium permanganate solution and the num-
duce materials with surface area lower than 1 m2 g−1 , ber of milliequivalents of iron oxidized by Mn(III)
P. Ciambelli et al. / Applied Catalysis B: Environmental 24 (2000) 243–253 245

Table 1
Sm1−x Srx MnO3 , NdMnO3 and LaMnO3 perovskitesa
Sm1−x Srx MnO3 NdMnO3 LaMnO3

x = 0.0 x = 0.1 x = 0.3 x = 0.5

Phases P P P + Sm P + Sm P P
a 5.39 5.36 5.45 5.43 5.44 5.52
b 5.71 5.74 5.49 5.43 5.79
c 7.68 7.56 7.52 7.50 7.55 13.33
V 236 233 225 221 238 351
Tc 0 +95 +90 +80 +100 +165
µ 5.14 5.07 5.90 5.33 5.90 5.72
SABET 19 20 14 13 20 20
SAcalc 29 26 36 36 33 13
H2 low T peak (T)max 0.20 (733) 0.20 (738) 0.47 (788) 0.47 (793) 0.17 (611) 0.21 (713)
H2 high T peak (T)max 0.44 (1023) 0.41 (1053) 0.27 (1056) 0.25 (1043) 0.50 (1063) 0.38 (1063)
Ea 17.1 20.8 18.6 18.4 19.3 23.3
Pre-exponential factor 0.21 1.30 0.65 0.63 0.35 8.1
D 290 330 240 250 260 730
Mnexp 22.2 21.1 21.6 22.0 21.5 20.0
Mnnom 21.5 22.1 23.1 24.6 22.0 22.4
Srexp 5.3 10.5 18.5
Srnom 3.7 10.9 19.6
Smexp 53.6 53.9 43.6 32.8
Smnom 58.9 54.1 44.7 33.7
Ndexp 53.4
Ndnom 57.9
Laexp 46.6
Lanom 56.7
Mn(IV)/Mntot 0.28 0.35 0.64 0.64 0.23 0.35
t 0.92 0.93 0.97
a Phases detected by XRD (P: perovskite, Sm: Sm O ). Lattice parameters: a/Å, b/Å, c/Å, V/Å3 . Curie temperature, T /K and magnetic
2 3 c
moment, µ/µB . Surface area (m2 g−1 ) determined by BET, SABET , and calculated (see text), SAcalc . Hydrogen uptake by TPR at low
T, H2 low T peak , (mol H2 mol−1 Mn), with T taken at the maximum of the peak, (T)max /K, in parentheses. Hydrogen uptake by TPR
at high T, H2 high T peak , (mol H2 mol−1 Mn), with T taken at the maximum of the peak, (T)max /K, in parentheses. Activation energy,
Ea (kcal mol−1 ). Pre-exponential factor (l m−2 h−1 ) × 106 . Crystallite size, D/Å. Experimental and nominal percentage for each element.
Fraction of Mn(IV)/Mn total. Tolerance factor, t.

and Mn(IV) was obtained by the difference between K is a constant equal to 0.9, λ the wavelength of the
the total milliequivalents of Fe(II) added and the mil- X-ray used, β the effective line width of the X-ray
liequivalents of excess iron determined by titration. reflection under observation, calculated by the expres-
Phase analysis, lattice parameters and particle sion β 2 = B2 − b2 [where B is the full width at half
sizes determination were performed by X-ray powder maximum (FWHM)], b the instrumental broadening
diffraction (XRD) using a Philips PW 1029 diffrac- determined through the FWHM of the X-ray reflec-
tometer with Ni-filtered Cu K␣ radiation. Lattice tion at θ = 14◦ of SiO2 having particles larger than
parameters were calculated by means of the UNIT- 1000 Å, θ the diffraction angle of the (1 1 1) X-ray
CELL program. 1 Particle sizes were evaluated by reflection (θ 1 1 1 = 12.8◦ ) for samarium and neodim-
means of the Scherrer equation D = Kλ/β cos θ after ium compounds, and of the (1 0 2) X-ray reflection
Warren’s correction for instrumental broadening [14]. (θ 1 0 2 = 11.5◦ ) for the lanthanum perovskite.
BET surface areas (SABET ) of the materials were
1 T.J.B. Holland and S.A.T. Redfern, J. Appl. Crystallogr., 30 measured by N2 adsorption at 77 K using a volumetric
(1997) 84. all glass apparatus.
246 P. Ciambelli et al. / Applied Catalysis B: Environmental 24 (2000) 243–253

The surface area values were also calculated mixed at atmospheric pressure to obtain inlet concen-
starting from particle sizes using the expression trations of 0.4% methane, 10% O2 , N2 as balance.
SAcalc = 30 000/rd, where r is the radius of crystal- The feed and product streams were analyzed by on
lites (supposed spherical) deduced from the crystallite line HP 6890 gaschromatograph equipped with ther-
sizes (D) measured by means of X-ray line broaden- mal conductivity and flame ionization detectors and
ing (r = D/2), and d is the bulk density derived from with Porapak Q and molecular sieve 5A columns. A
the formula: d = (MZ)/(VN), where M is the molar silica gel water trap placed before the gaschromato-
mass of the formula unit, Z the number of formula graph allowed to dry the product stream. For each test
units per unit cell, V the measured unit cell volume, the methane conversion was calculated as the average
N the Avogadro’s constant. of at least three measurements. Carbon balance was
Magnetic susceptibilities were measured by the closed to within ±5% in all catalytic tests.
Gouy method over the temperature range 100–300 K
and at different magnetic-field strengths. Correction
was made for the diamagnetism of the samples. 3. Results and discussion
Temperature-programmed desorption (TPD) of O2
and temperature-programmed reduction (TPR) with 3.1. Oxides
H2 were performed using a Micromeritics TPD/TPR
2900 analyser equipped with a TC detector and cou- Table 1 reports the main features of the perovskite
pled with a Hiden HPR 20 mass spectrometer. The samples in terms of phases present, chemical composi-
sample (30 mg) was preheated in flowing air at 1073 K tion, surface area, structural and magnetic properties.
for 2 h before each TPD or TPR test. In TPR analyses a Fig. 1 shows that LaMnO3 , NdMnO3 , SmMnO3
2% H2 /Ar mixture (25 cm3 min−1 ) was used to reduce and Sm0.9 Sr0.1 MnO3 samples are single phase
the sample by heating 10 K min−1 up to 1073 K. Wa-
ter produced by the sample reduction was condensed
in a cold trap before reaching the detectors. Only H2
was detected in the outlet gas confirming the effective-
ness of the cold trap. In O2 TPD analyses the sample
was heated 10 K min−1 up to 1073 K in flowing He
(25 cm3 min−1 ). Only O2 was detected in the outlet
gas of TPD measurements.
Catalytic combustion experiments were carried out
with a downflow quartz annular reactor electrically
heated in a three zone tube furnace. The annular
cross-section was chosen to obtain a small equiv-
alent diameter that enables to control the catalyst
temperature and to reduce the temperature gradients
within the bed. Catalyst particles in the range of
180–250 ␮m were diluted 1 : 10 in quartz powder of
the same dimension and placed on a porous quartz
disk. The narrowing of the reactor diameter both
in the pre- and in the post-catalytic zone and the
presence of ␣-Al2 O3 pellets upside the catalytic bed
limited the occurrence of homogeneous reactions.
The temperature of the catalytic bed was measured
by a K-type thermocouple. The space velocity was
Fig. 1. XRD spectra for all samples. (a) LaMnO3 , (b) NdMnO3 ,
40 000 N cm3 g−1 h−1 in all tests, the temperature was (c) SmMnO3 , (d) Sm0.9 Sr0.1 MnO3 , (e) Sm0.7 Sr0.3 MnO3 , (f)
in the range 623–1023 K. The gaseous flow rates were Sm0.5 Sr0.5 MnO3 . Asterisks indicate the most intense lines of
measured by Brooks 5850 mass flow controllers and Sm2 O3 .
P. Ciambelli et al. / Applied Catalysis B: Environmental 24 (2000) 243–253 247

perovskite-type oxides. For the Sm1−x Srx MnO3 cat- Note that the elemental contents (reported in Table
alysts with x = 0.3 and x = 0.5 the XRD pattern re- 1) expected for the given formulae are in reasonable
veals, in addition to perovskite, the presence of two agreement with the experimental ones. Some more sig-
peaks at 2θ = 28.8◦ (d = 3.10 Å) and at 2θ = 30.6◦ nificant difference between nominal and experimental
(d = 2.92 Å) which correspond to the two strongest values is indeed found for the rare-earth elements in
lines of hexagonal Sm2 O3 (d = 3.09 Å, intensity the bicomponent oxides, and this may indicate that the
70; d = 2.94 Å, intensity 100) (a). NdMnO3 (b) and A-sublattice could be more cation defective than the
Sm1−x Srx MnO3 (c) oxides exhibit the orthorhombic B-sublattice for (Sm,Nd,La)MnO3 .
perovskite Pbnm structure, whereas LaMnO3 crys- The evaluation of the unit cell volume (reported in
tallizes in a primitive rhombohedral (non-primitive Table 1 and in Fig. 2) shows that the Sm1−x Srx MnO3
hexagonal) lattice (d) 2 . solid solution system presents a lattice contraction
Redox titration showed that in all AMnO3 samples at the increase of x, which is the result of the com-
there is a substantial fraction of manganese as Mn(IV) bined effect of the substitution of bigger Sr(II) ions
(35% for LaMnO3 , 23% for NdMnO3 and 28% for for Sm(III) in the 12-coordinated A sites and of
SmMnO3 , see Table 1). Thus, in agreement with the smaller Mn(IV) for Mn(III) in the 6-coordinated
results found by other authors for LaMnO3 [15–17], B sites {rSr(II) = 1.44 Å and rSm(III) = 1.24 Å,
also in our case the (La,Nd,Sm)-Mn perovskites can rMn(IV) = 0.53 Å and rMn(III) = 0.645 Å, for dodec-
be regarded as nonstoichiometric oxidized AMnO3+δ ahedral and octahedral coordination, respectively
materials. [18]}. Note also that NdMnO3 shows higher lat-
By taking into account the observed Mn(IV) con- tice volume than SmMnO3 , in agreement with the
tent and the concentration of each other element with relative rare-earth cation sizes [rSm(III) = 1.24 Å,
its corresponding charge, the following formulae rNd(III) = 1.27 Å].
have been derived for each LaMnO3 , NdMnO3 and
Sm1−x Srx MnO3 (x = 0.0, 0.1, 0.3, 0.5) catalyst:
LaMn(III)0.65 Mn(IV)0.35 O3.18
NdMn(III)0.77 Mn(IV)0.23 O3.12
SmMn(III)0.72 Mn(IV)0.28 O3.14
Sm0.9 Sr0.1 Mn(III)0.65 Mn(IV)0.35 O3.13
Sm0.7 Sr0.3 Mn(III)0.36 Mn(IV)0.64 O3.17
Sm0.5 Sr0.5 Mn(III)0.36 Mn(IV)0.64 O3.07
By normalizing to the three oxygens formula and
supposing, as found by Van Roosmalen et al. [16], the
presence of an equal amount of cation vacancies in
the A and B sites, the following cation defective per-
ovskites may be obtained (the corresponding formula
weight, FW, is given in parenthesis for each sample):
Nd0.96 Mn(III)0.74 Mn(IV)0.22 O3 (FW = 239.2)
La0.94 Mn(III)0.61 Mn(IV)0.33 O3 (FW = 230.2)
Sm0.96 Mn(III)0.69 Mn(IV)0.27 O3 (FW = 245.2)
Sm0.86 Sr0.1 Mn(III)0.62 Mn(IV)0.34 O3 (FW = 238.8)
Sm0.67 Sr0.28 Mn(III)0.34 Mn(IV)0.61 O3 (FW = 225.5)
Sm0.49 Sr0.49 Mn(III)0.35 Mn(IV)0.63 O3 (FW = 218.5)

2 X-Ray Powder Data File, ASTM cards: (a) 19-1114 for Sm O ;


2 3
(b) 25-565 for NdMnO3 ; (c) 25-747 for SmMnO3 ; (d) 32-484 for Fig. 2. Cell volume as a function of the substitutional parameter
LaMnO3 ; (e) 39-1190 for La0.8 Sr1.2 CuO3.4 . x. 䊊 Sm1−x Srx MnO3 , 䊐 NdMnO3 .
248 P. Ciambelli et al. / Applied Catalysis B: Environmental 24 (2000) 243–253

so-called ‘double-exchange’ interaction between near-


est Mn(III) and M(IV) paramagnetic cations [19].
The effective magnetic moment for all samples (µ
in the range 5.2–5.9 ␮B ) is higher than that expected
from the spin-only values of the paramagnetic species
present in the sample [Mn(III), Mn(IV)]. Note that
for LaMnO3 a rather high value (µ = 5.4 ␮B ) has also
been observed by Jonker [20].

3.2. TPR and O2 TPD measurements

No other reactions except the reduction of the


catalysts occur during the TPR experiments, only
H2 consumption having been detected both by TCD
and mass spectrometer. TPR profiles of NdMnO3 ,
LaMnO3 , SmMnO3 and Sm1−x Srx MnO3 perovskites
are shown in Fig. 4. The overall signal is made of two
main contributions for all catalysts. Furthermore, the
complexity of the low temperature peak in the TPR
Fig. 3. Inverse magnetic susceptibility as a function of temperature. curves of Sm1−x Srx MnO3 samples with 0.0 ≤ x ≤ 0.3
䊉 LaMnO3 , 䊊 NdMnO3 , 䊐 SmMnO3 , 4 Sm0.9 Sr0.1 MnO3 , N suggests the contribution of at least two signals also in
Sm0.7 Sr0.3 MnO3 , 䉬 Sm0.5 Sr0.5 MnO3 . this temperature region. Both SmMnO3 and NdMnO3
samples show very sharp signals at high temperature
The crystallite sizes (D1 0 2 for the hexagonal with maximum at 1023 and 1063 K, respectively. On
LaMnO3 material, and D1 1 1 for the samarium and the contrary, LaMnO3 sample shows a broad signal
neodimium orthorhombic perovskites) are reported in at high temperature occurring under isothermal con-
Table 1. Their values are 730 Å for LaMnO3 , and in ditions, suggesting a slower reduction process. The
the range 220–330 Å for all others catalysts. onset temperature of the reduction (Table 1) is in the
Surface areas (SABET ), reported in Table 1, are order LaMnO3 > NdMnO3 > SmMnO3 .
in the range 13–20 m2 g−1 . Also the surface areas XRD analysis performed at the end of TPR exper-
(SAcalc ) estimated on the basis of the radius of crys- iments showed the formation of MnO, A2 O3 and/or
tallite sizes and of the X-ray density (d ≈ 7 g cm−3 ) A(OH)3 (A = La, Sm, Nd) for LaMnO3 , NdMnO3 ,
are reported in Table 1. Note that the values of SABET SmMnO3 and Sm0.9 Sr0.1 MnO3 catalysts, in agree-
and SAcalc are of the same order of magnitude. The ment with results reported in [21,22] for La–Mn
difference between SABET and SAcalc may be due to based perovskites. The X-ray spectra of the re-
the fact that, in the calculation of the surface area, all duced Sm1−x Srx MnO3 samples with x = 0.3 and 0.5,
particles were supposed to have a spherical shape, and apart from the X-ray lines correspondent to MnO,
this could not be true in our samples. matched the pattern of a compound described by
Fig. 3 shows the inverse atomic susceptibility, N. Nguyen et al. as La0.8 Sr1.2 CuO3.4 [23]. It was
1/χ at versus T. Note that only LaMnO3 does not not surprising to find that, considering the analogy
show a linear behavior of 1/χ at for the whole range among Cu(II)–Mn(II) and La(III)–Sm(III) species,
of temperatures. All samples but SmMnO3 exhibit the product of complete reduction of our samples
a ferromagnetic behavior. The extrapolated Curie at higher strontium content is similar to the parent
points, TC , and the effective
√ magnetic moments per La0.8 Sr1.2 CuO3.4 compound. Fig. 5 reports, as exam-
formula unit, µ = 2.83 χat T , are reported in Table ple, the XRD patterns of Sm0.5 Sr0.5 MnO3 (original,
1.The ferromagnetic behavior of AMnO3 perovskites after the first reduction step and after complete reduc-
is indeed predicted by theory and is caused by the tion).
P. Ciambelli et al. / Applied Catalysis B: Environmental 24 (2000) 243–253 249

Fig. 4. (a) TPR profiles of LaMnO3 (full line), NdMnO3 (line with circles) and SmMnO3 (dotted line) perovskites. (b) TPR profiles of
Sm1−x Srx MnO3 perovskites. x = 0.1 (full line), x = 0.3 (dotted line), x = 0.5 (line with circles).

main peaks was chosen to separate them. The values


of the total H2 /Mn ratio, higher than 0.5 for all cat-
alysts, confirm the presence of a fraction of Mn(IV)
in the samples, as detected by the chemical analy-
sis. According to literature data [21,22], the reduction
steps Mn(IV) → Mn(III) → Mn(II) occur in LaMnO3
perovskite, the former at low temperature and the lat-
ter at high temperature. The same sequence could be
roughly hypothesized also for Nd and Sm containing
perovskites. XRD patterns of the samples taken after
the first reduction step show the signals of a less crys-
talline perovskite phase (Fig. 5 for Sm0.5 Sr0.5 MnO3 ,
as example), but with increased cell parameters. The
lattice expansion can be explained by the reduction
of the Mn(IV) fraction to Mn(III) which has a larger
ionic radius than Mn(IV), thus confirming the above
interpretation of TPR curves. According to the above
hypothesis, the first TPR signal should correspond to
the reduction Mn(IV) → Mn(III) and the second to
the reduction Mn(III) → Mn(II). Therefore, a value of
Fig. 5. XRD spectra of Sm0.5 Sr0.5 MnO3 . (a) sample calcined at H2 /Mn ratio of 0.5 should be expected for the sec-
1073 K, (b) after TPR first peak, (c) after TPR up to 1073 K. As- ond peak. However, except for NdMnO3 , lower val-
terisks and circles indicate the strongest X-ray lines corresponding ues were evaluated for the other samples. Thus, some
to Sm2 O3 and MnO, respectively. X-ray lines for SmMnO3 (c) reduction to Mn(II) can likely occur even at low tem-
and for La0.8 Sr1.2 CuO34 (e) reference compounds are given at the
perature.
bottom and at the top, respectively.
All Sr substituted perovskites show (Fig. 4) com-
plex TPR profiles due to the contribution of several
In Table 1 the evaluated H2 uptakes, relevant to signals, with shape and intensity influenced by the ex-
the low and high temperature peaks, are reported. The tension of Sm substitution by Sr. Indeed, the signal at
minimum value reached by the signal between the two low temperature increases while that at high temper-
250 P. Ciambelli et al. / Applied Catalysis B: Environmental 24 (2000) 243–253

ature decreases with the Sr content. A similar effect showing that the perovskite phase was restored after
due to Sr substitution was observed for LaMnO3 per- re-oxidation in air.
ovskites by Irutsa et al. [24]. In the low temperature O2 TPD spectra (not reported) show two peaks for
signal (T < 833 K) at least two components can be ob- all catalysts, the former with the maximum in the tem-
served, whose intensity is strongly affected by Sr con- perature range 513–733 K and the latter with the max-
tent. A signal at about 623 K, well evident for x = 0.1, imum in the range 916–1073 K. The amount of oxy-
decreases with increasing Sr and, at the same time, gen related to the low temperature peak is very small
the intensity of the signal at about 773 K markedly in- for all catalysts while that evolved at high tempera-
creases with Sr substitution. ture ranges from 0.024 to 0.20 mol O2 per total mol
The fraction of Mn(IV), evaluated from the total H2 of transition metal cation, the maximum value being
uptake, increases with Sr substitution, in agreement given by LaMnO3 . The first peak, referred to as ␣
with the results of chemical analysis. Some reduction peak, was attributed to oxygen species weakly bound
of manganese to Mn(II) likely occurs in the low tem- to the surface while the second peak (β peak) was re-
perature region also for the Sr substituted samples as lated to the reduction of B cations to lower oxidation
shown by the very high value of the H2 uptake cor- state [8] in the ABO3 structure. The values of O2 re-
responding to the first peak, increasing with the Sr leased both at low and at high temperature suggest
fraction. These results suggest that the addition of Sr that the oxygen evolution must be related only to the
enhances the formation of Mn(IV) easily reducible to catalyst surface since larger amounts of O2 should be
Mn(II) which is already present, in lower amount, in expected for a phenomenon involving the bulk of the
the SmMnO3 perovskite. The Sr substitution promotes sample. XRD analysis performed on the samples after
the reducibility of Mn(IV) probably because it causes the TPD experiments revealed that no phase transfor-
a lower crystallinity of the perovskite, as suggested mation occurs confirming the previous hypothesis.
by Irutsa et al. [24] for La1−x Srx MnO3 perovskites, The reversibility of the O2 release process was ver-
and/or because it causes the formation of small crys- ified by performing a second TPD experiment after a
tallites of manganese oxides not detectable by XRD. further treatment of the sample 2 h in flowing air at
Manganese oxides were reported by Arnone et al. [25] 1073 K. The complete superimposition of the first and
to be easier reducible than manganese-containing per- the second TPD profile for all catalysts confirms the
ovskites. As a consequence, the complexity of the low reversibility of the process. XRD spectra performed
temperature signal could be related either to the oc- after O2 TPD cycles accordingly showed that the struc-
currence of two mono-electronic steps by reduction of ture of catalysts was unchanged.
Mn(IV) within the perovskite first to Mn(III) and then
to Mn(II), or to the contribution of some manganese
oxides. However, we are well aware that the prepara- 3.3. Catalytic activity measurements
tion conditions used for this work (previous prolonged
calcinations at lower temperatures and then final cal- Preliminary tests performed in the absence of cata-
cination at 1073 K for 5 h) should provide a good crys- lyst showed that homogeneous reactions are negligible
tallization of possible manganese oxides (as it occurs under the experimental conditions investigated.
for the perovskite phase), thus easily detectable by The results of catalytic activity measurements are
XRD. We are thus more inclined to suggest the two reported in Figs. 6 and 7. All catalysts give complete
mono-electronic reduction steps of Mn(IV) in the per- conversion of methane below 1023 K with 100% se-
ovskite phase. lectivity to CO2 . After the first testing cycle, all cat-
After TPR experiments all samples were treated in alysts were cooled down to room temperature and a
air flow at 1073 K and reduced again under the same second cycle was performed. The results of the sec-
conditions of the first experiments obtaining very sim- ond cycle were the same of the first one, suggesting
ilar results. This indicates that the catalysts undergo a that the catalysts do not undergo modification nor de-
reversible reduction process. This behavior was also activation in the reaction conditions investigated.
confirmed by XRD analysis performed on a sample NdMnO3 is the less active catalyst in the whole
re-oxidized in air at 1273 K after TPR experiments temperature range. LaMnO3 is less active than
P. Ciambelli et al. / Applied Catalysis B: Environmental 24 (2000) 243–253 251

Fig. 6. CH4 conversion as a function of temperature. 䊉 LaMnO3 , Fig. 8. Arrhenius plots for LaMnO3 , NdMnO3 and SmMnO3
N NdMnO3 and 䊏 SmMnO3 perovskites. perovskites. Symbols as in Fig. 6.

Fig. 7. CH4 conversion as a function of temperature for


Sm1−x Srx MnO3 perovskites. 䊏 x = 0,0, 䊐 x = 0,1 䊊 x = 0.3, 4 Fig. 9. Arrhenius plots for Sm1−x Srx MnO3 perovskites. Symbols
x = 0.5. as in Fig. 7.

SmMnO3 up to 773 K, the reverse at higher temper- The values of the apparent activation energy and
ature. Sm1−x Srx MnO3 perovskites show comparable of the pre-exponential factor, estimated from the Ar-
catalytic activities, slightly increasing with Sr con- rhenius plots (Figs. 8 and 9) on the base of a CH4
tent, but they are less active than the unsubstituted first-order rate equation and on the hypothesis of
perovskite. isothermal PFR behavior are reported in Table 1 for
252 P. Ciambelli et al. / Applied Catalysis B: Environmental 24 (2000) 243–253

all catalysts. The curves reported in Figs. 6 and 7, cal- We did not find a similar effect for the Sm1−x Srx MnO3
culated with the estimated kinetic parameters, show system.
a satisfactory agreement with the experimental data.
The highest value of apparent activation energy was
found for LaMnO3 , while lower values were estimated 4. Conclusions
for Nd and Sm containing perovskites. Comparable
values of apparent activation energy were reported in The main points put forward in this paper can be
[26] for ACoO3 catalysts, lanthanum based perovskite summarized as follows: (i) all examined samples, but
showing the highest one. Sm0.7 Sr0.3 MnO3 and Sm0.5 Sr0.5 MnO3 , were found to
As reported in [27] two different mechanisms can be be single perovskite phases; (ii) Mn(IV) in addition to
hypothesized for methane oxidation: the first, involv- Mn(III) was found in all catalysts, the Mn(IV) frac-
ing chemisorbed oxygen, is controlling at low tem- tion increasing with x in the Sm1−x Srx MnO3 cata-
perature, whereas the second, involving lattice oxy- lysts; (iii) all samples gave complete methane con-
gen, occurs at high temperature. TPR experiments version with 100% selectivity to CO2 below 1023 K;
showed that the reduction of manganese starts at lower (iv) the catalytic activity paralleled the reducibility of
temperature for NdMnO3 and SmMnO3 with respect Mn(IV) in the various catalysts; (v) the substitution
to LaMnO3 . The easier reducibility of NdMnO3 and of Sr for Sm resulted in a reduction of activity, this
SmMnO3 could be related to the low value of activa- behaviour could tentatively be related to reduction of
tion energy for these two samples. This agrees with Mn(IV) to Mn(II) under reaction, hindering the redox
the results reported by Futai et al. [12] who associ- mechanism.
ated the easy reducibility of ACoO3 perovskites to the
high activity in CO oxidation and reported the same
order of activation energy for La, Sm and Nd per- References
ovskites. However, LaMnO3 perovskite, where man-
[1] M.F. Zwinkels, S.G. Järås, P.G. Menon, Catal. Rev.-Sci. Eng.
ganese is more stable towards reduction, is capable
35 (1993) 319.
to desorb larger quantities of lattice oxygen as shown [2] K. Eguchi, H. Arai, Catal. Today 29 (1996) 379 .
by TPD, thus being the most active sample at high [3] R.L. Garten, R.A. DallaBetta, J.C. Schlatter, in: G. Ertl, H.
temperature. Knözinger, J. Weitkamp (Eds.), Handbook of Heterogeneous
As concerns Sr substituted samples, the perovskite Catalysis, vol. 4, VHC Weinheim, Germany, 1998, p. 1668.
[4] H. Arai, M. Machida, Catal. Today 10 (1991) 81.
with the lowest Sr content (x = 0.1) shows an ap-
[5] R. Burch, Catal. Today 35 (1997) 27.
parent activation energy larger than that obtained for [6] J.G. Mc Carty, H. Wise, Catal. Today 8 (1990) 231.
SmMnO3 , whereas a further increase of Sr content [7] R.E. Newman, in: A. Navrotsky, D.J. Weidner (Eds.),
(samples with x = 0.3 and 0.5) leads to a decrease of Structure–property relationship in perovskite electroceramics:
the activation energy values. The same trend of the Perovskite: a structure of great interest to geophysics and
material science, American Geophysical Union, Washington,
activation energy values was observed for the onset
DC, 1989, p. 66.
temperature of the reduction in the TPR experiments [8] L.G. Tejuca, J.L.G. Fierro (Eds.), Properties and applications
thus confirming that an easy reducibility of manganese of perovskite-type oxides, Marcel Dekker, New York, 1993.
is correlated to the ability to activate methane at low [9] R.J.H. Voorhoeve, D.W. Johnson Jr., J.P. Remeika, P.K.
temperature. On the other hand, if manganese oxides, Gallagher, Science 195 (1977) 827.
[10] L.G. Tejuca, J.L.G. Fierro, J.M.D. Tascón, Adv. Catal. 36
in addition to the perovskite phase, were formed in
(1989) 237.
the samples with higher level of Sr substitution, a de- [11] A. Baiker, P.E. Marti, P. Keusch, E. Fritsch, A. Reller, J.
crease in the apparent activation energy was to be ex- Catal. 146 (1994) 268.
pected since manganese oxides are known to activate [12] M. Futai, C. Yonghua, Louhui, React. Kinet. Catal. Lett. 31
methane oxidation at lower temperature as compared (1986) 47.
[13] H.M. Zhang, Y. Teraoka, N. Yamazoe, Chem. Lett. (1987)
to corresponding perovskites [25].
665.
Note that a promoting effect of Sr on the catalytic [14] H.P. Klug, L.E. Alexander, X-ray Diffraction Procedures for
activity was related by other authors to the presence of Polycrystalline and Amorphous Materials, Wiley, London,
Mn(IV) and Co(IV) in perovskite structures [6,12,27]. 1962, p. 491.
P. Ciambelli et al. / Applied Catalysis B: Environmental 24 (2000) 243–253 253

[15] B.C. Tofield, W.R. Scott, J. Solid State Chem. 10 (1974) 183. [22] M.L. Rojas, J.L.G. Fierro, L.G. Tejuca, A.T. Bell, J. Catal.
[16] J.A.M. van Roosmalen, E.H.P. Cordfunke, R.B. Helmholdt, 124 (1990) 41.
H.W. Zandbergen, J. Solid State Chem. 110 (1994) 100. [23] N. Nguyen, J. Choisnet, M. Hervieu, B. Raveau, J. Solid
[17] I.G. Krogh Andersen, E. Krogh Andersen, P. Norby, E. Skou, State Chem. 39 (1981) 120.
J. Solid State Chem. 113 (1994) 320. [24] S. Irutsa, M.P. Pina, M. Menendez, J. Santamaria, J. Catal.
[18] R.D. Shannon, Acta Crystallogr. Sect. A 32 (1976) 751. 179 (1998) 400.
[19] J.B. Goodenough, Magnetism and the Chemical Bond, Wiley, [25] S. Arnone, G. Busca, L. Lisi, F. Milella, G. Russo, M. Turco,
New York, 1963, p. 221. Proc. 27th Int. Symp. on Combustion, Boulder, USA, 1998.
[20] G.H. Jonker, J. Appl. Phys. 37 (1966) 1424. [26] H. Arai, T. Yamada, K. Eguchi, T. Seiyama, Appl. Catal. 26
[21] L. Lisi, G. Bagnasco, P. Ciambelli, S. De Rossi, P. Porta, G. (1986) 265.
Russo, M. Turco, J. Solid State Chem. 146 (1999) 291. [27] T. Seiyama, Catal. Rev.-Sci. Eng. 34 (1992) 281.

You might also like