You are on page 1of 9

Article

Journal of
Copyright © 2017 American Scientific Publishers
Nanoscience and Nanotechnology
All rights reserved Vol. 17, 3864–3872, 2017
Printed in the United States of America www.aspbs.com/jnn

Influence of Thermal Treatments on the Reducibility and


Catalytic Properties of Pd Supported Over
Ce06Zr04Ox /SiO2 and Ce073Tb027Ox /SiO2 for
Methane Oxidation
Victor Ferrer1 ∗ , Angeli Barroso1 , Jean Bracho1 , Dora Finol1 , Francisco Domínguez1,
Alexander Moronta1, Miguel Ramos2 , Fabrizio Puleo3 , and Leonarda F. Liotta3 ∗
1
Instituto de Superficies y Catálisis, Facultad de Ingeniería, Universidad del Zulia, P.O. Box 15251, Maracaibo 4003A, Venezuela
2
Instituto Zuliano de Investigaciones Tecnológicas (INZIT), La Cañada de Urdaneta, Apartado 331, Maracaibo, Venezuela
3
Istituto per lo Studio dei Materiali Nanostrutturati (ISMN-CNR), via Ugo La Malfa 153, I-90146, Palermo, Italy

Pd/Ce06 Zr04 Ox /SiO2 and Pd/Ce073 Tb027 Ox /SiO2 catalysts were prepared by incipient wetness
impregnation method using aqueous solutions of Ce(NO3 3 · 6H2 O, ZrO(NO3 2 · nH2 O, Tb(NO3 3 ·
5H2 O and Pd(NO3 2 · nH2 O precursors. The catalysts were characterized by X-ray diffraction
(XRD), Temperature Programmed Reduction (TPR-H2 , CO Chemisorption, Oxygen Storage Capac-
ity (OSC) and Infrared spectroscopy of adsorbed CO (FTIR-CO). The catalytic activity was mea-
sured in the methane oxidation reaction. XRD patterns showed that after two activity cycles, the
solids crystallinity was not modified, while TPR results revealed that reduction maxima were shifted
to lower temperatures, indicating an improvement in sample reducibility. The CO/Pd ratio diminished
with increasing the reduction temperature, suggesting both, Pd inward diffusion/covering by Ce3+
ions (geometrical effects) and/or increase in Pd particles size. The IR spectrum of adsorbed CO
reported the presence of cationic palladium species interacting with CO regardless of reduction
temperature as well as carbonate species and weak interactions with ceria and silica. Zr and Tb
containing samples presented a high and stable OSC after undergoing various OSC cycles. The
catalytic activity was enhanced after two reaction cycles due to the formation of large PdO particles.
T50 values indicated that reduction treatments prior to methane reaction generated more active cat-
alytic species than those obtained by oxidation treatment; however, such active species were not
stable at high temperatures.
Keywords: Methane Oxidation, Mixed Oxides, OSC, Pd Catalysts, Thermal Treatments.

1. INTRODUCTION are considered as precursors of climate change.3 In big


Heterogeneous catalysis is focused to resolve one of the cities, where vehicles concentration in low areas is high,
most serious problems of anthropogenic origin: the atmo- the accumulation of these pollutants in the lowest atmo-
spheric pollution generated by vehicles.1 In a complete sphere layer produces thermal inversion, generating photo-
combustion, the hydrocarbons are transformed into carbon chemical smog, acid rain, global warming and ozone layer
dioxide (CO2  and water (H2 O); however, due to ther- depletion.4
modynamic reasons or mechanical flaw, the hydrocarbons Three-way catalysts (TWC) are considered as the most
oxidation (e.g., gasoline or gasoil) is incomplete, gen- successful pollutants post-treatment for gasoline and diesel
erating carbon monoxide (CO), nitric oxides (NOx  and motors.5 6 In heated TWC in stationary condition, con-
non-burnt hydrocarbons (HC).2 The influence of these pol- versions of 95% and higher are reached for CO, NOx
lutants over air quality has been widely studied, and they and HC. TWC can transform CO, NOx and HC into
CO2 , H2 O and molecular nitrogen.7 The TWC are con-

Authors to whom correspondence should be addressed. formed by noble metals such as Pd, Pt, Rh and a redox

3864 J. Nanosci. Nanotechnol. 2017, Vol. 17, No. 6 1533-4880/2017/17/3864/009 doi:10.1166/jnn.2017.14000


Ferrer et al. Influence of Thermal Treatments on the Reducibility and Catalytic Properties

promoter (typically cerium oxide) supported on a ceramic and catalytic activity of Ce06 Zr04 Ox and Ce073 Tb027 Ox
monolith of high surface area and silicon content. It has mixed oxides supported on SiO2 and impregnated with
been reported that fluorite-structured solid solutions, based Pd was studied. The use of silica as support was pre-
on mixtures of CeO2 /La2 O3 /Pr2 O3 and CeO2 /Ln2 O3 /Pr2 O3 ferred to alumina in order to prevent the formation under
where Ln = Tb, Sm, Gd are effective catalysts for methane reducing conditions of the inactive CeAlO3 phase.30 The
combustion.8 samples were characterized by X-ray fluorescence (ED-
Methane combustion has been of special interest since XRF), X-ray powder diffraction (XRD), Temperature pro-
natural gas, which consists mainly of methane, is a com- grammed reduction (TPR), CO chemisorption, Oxygen
mon fuel used worldwide. Since methane is the most storage capacity (OSC) and Fourier transformed infrared
refractory hydrocarbon, it is often used as the model spectroscopy of adsorbed CO (FTIR-CO). The methane
hydrocarbon compound for activity tests.910 The use of oxidation reaction was studied under stoichiometric con-
Pd has been known as the most active component for the ditions throughout two reaction cycles, and after reduction
combustion of methane among noble metal catalysts.11 treatments.
The use of CeO2 as an additive in the so-called three
way catalysts for automotive exhaust treatment have been
widely investigated because it has been found effective
2. EXPERIMENTAL SECTION
in the promotion of various catalytic reactions including, 2.1. Catalysts Preparation
CO2 activation12 , CO oxidation13 and CO/NO removal.14 SiO2 (Baker) was previously calcined at 400  C for 2 h in
CeO2 has the capacity to shift between CeO2 under oxi- order to remove any water, carbonate and any other volatile
dizing conditions and Ce2 O3 under reducing conditions, impurities, then it was impregnated with aqueous solu-
respectively, and the ease of formation of labile oxygen tions of Ce(NO3 3 ·6H2 O, ZrO(NO3 2 ·nH2 O or Tb(NO33 ·
vacancies and, particularly, the relatively high mobility 5H2 O. All these precursors have purity ≥99.0 and were
of bulk oxygen species. However, the chemical proper- provided by Acros Organics. Cerium and zirconium or ter-
ties of CeO2 can be affected when it is subjected to both bium nitrates were mixed in appropriate amounts in order
high temperature and reducing conditions,15 this can lower to obtain a mass ratio of Ce/Tb = Ce/Zr = 2.3 (molar ratio
OSC values and catalytic activity. Therefore, a great deal Ce/Tb = 2.70 and Ce/Zr = 1.5) and a mixed oxide compo-
of attention has been given to doping of CeO2 with iso- sition of 17 wt.%. The so prepared samples were calcined
valent/aliovalent cations to enhance the OSC properties, at 700  C for 4 h and labeled as Ce073 Tb027 Ox /SiO2 (CTS)
especially to improve the bulk reducibility at lower tem- and Ce06 Zr04 Ox /SiO2 (CZS).
peratures and to stabilize the structure against sintering.16 The corresponding Pd catalysts, Pd/CZS and Pd/CTS,
It has been reported that the addition of a dopant like Zr or with a metal loading of 0.5 wt% were prepared by incip-
Tb to ceria, improves its textural stability,17 18 prevents Ce ient wetness impregnation method using Pd(NO3 2 · nH2 O
sintering19 and enhances the mobility of lattice oxygen.15 solution (Aldrich, 99.9%) in HCl 0.1 N. Finally, the cata-
The introduction of smaller isovalent non-reducible cations lysts were calcined at 500  C for 2 h.
like Zr4+ and Hf4+ into the ceria lattice enhances the OSC For comparison purposes, CeO2 /SiO2 (CS) and 0.5%Pd/
by creating intrinsic oxygen vacancies thereby increasing CeO2 /SiO2 (Pd/CS) samples were also prepared and used
the oxygen mobility by facilitating the Ce3+ /Ce4+ redox for certain characterizations.
process.20 Tests carried out on CeO2 –ZrO2 mixed oxides
showed better thermal stability and OSC than that of pure 2.2. Techniques
ceria6 21–24 and it has been established that the addition of The elemental concentrations of the samples were quanti-
zirconium is important to stabilize performance after cal- fied by energy dispersive X-ray fluorescence spectroscopy
cination and to provide more reactive oxygen ions from (ED-XRF) using a Shimadzu EDX-700HS spectrometer,
the surface and the bulk.25 Recently, the use of variable operating at 50 kV and 30 mA.
valence dopants into the CeO2 lattice has attracted much The X-ray powder diffraction (XRD) patterns were
attention.2627 Among the reducible elements, terbium (Tb) acquired with a Bruker D8 Focus diffractometer operated
and praseodymium (Pr) are particularly suitable for mak- at 40 kV and 40 mA, with Cu K radiation ( = 15418 Å)
ing solid solutions with ceria.28 Other mixed oxides such at a rate of 1.1 2 min−1 with 0.02 step size and time
as Ce08 Tb02 Ox have shown good redox behavior and cat- per step of 0.5 s.
alytic activity.29 TPR measurements were carried out in a stainless steel
For this reason, the studies related to the improve- reaction line coupled to a thermal conductivity detec-
ment of chemical and catalytic properties of TWC are of tor (TCD). 50 mg of calcined catalyst were loaded in a
great relevance. Textural and redox properties of TWC quartz reactor and subjected to a cleaning/oxidation treat-
are closely related with operation conditions;23 however, ment that consisted in heating the sample from room tem-
the effect of catalytic cycles as well as thermal treatments perature (RT) up to 550  C (10  C min−1  in a flow of
over TWC properties is still under study. In this research, O2 (5%)/He (30 cm3 min−1 ) for 1 h, cool down to 150  C,
the influence of thermal treatments over the reducibility switching the gas flow to Ar (Praxair, 99.999%) and finally

J. Nanosci. Nanotechnol. 17, 3864–3872, 2017 3865


Influence of Thermal Treatments on the Reducibility and Catalytic Properties Ferrer et al.

cool down to room temperature. Pd containing samples the evacuation continued for 15 min. After finishing the
were cooled down at a flow rate of 30 cm3 min−1 of Ar pre-treatment, the wafer was moved to the analysis region
(Praxair, 99.995%) to −80  C in a cold trap. Then, the flow and the spectrum was taken. Then, a continuous flow of
rate was changed to 30 cm3 min−1 of H2 (5.22%)/Ar (Prax- CO (Praxair, 99.5%) was admitted to a pyrex-glass cell
air) and the cold trap was removed, starting to compute for 10 min at atmospheric pressure. The wafer was evac-
H2 consumption from −80  C to room temperature. After- uated in a flow of Ar for 30 min and finally, the spectrum
wards, the sample was heated up to 850  C using a heating was obtained. The spectrum of the pre-treated catalyst was
rate of 10  C min−1 . TPR experiments were also car- used as reference and subtracted from the spectra of the
ried out over exhausted samples subjected to two reaction catalyst exposed to CO.
cycles (denoted as the corresponding notation followed by
−ex), previously dried at 120  C for 30 min in Ar to elim- 2.3. Catalytic Activity
inate water formed during reaction. The methane oxidation reaction was performed from RT
CO chemisorption was carried out in the TPR line up to 700  C, in a quartz reactor, using 50 mg of catalyst
described above. 50 mg of sample were submitted to the in a fixed bed. Before reaction, catalyst was submitted to
oxidation treatment mentioned above (550  C, O2 /He). The the oxidation treatment described above. The reactor was
sample was cooled down until the reduction temperature fed with a gas mixture (102.7 cm3 min−1  using a flow
(300 or 500  C) and reduced in H2 (5%)/Ar flow for 1 h. of methane (2.7 cm3 min−1 , oxygen (5.4 cm3 min−1  and
Afterwards, the sample was evacuated in He gas (Aga, helium as balance (94.6 cm3 min−1  in stoichiometric pro-
99.999%) for 30 min at the same reduction temperature. portion (CH4 /O2 = 05, space velocity = 14342 h−1 . The
When the treatment finished, the sample was cooled down methane composition was analyzed at the reactor exit with
to room temperature. Finally, 1.47 mol of CO (Praxair, a Perkin Elmer Clarus 500 gas chromatograph, equipped
99.5%) were injected repeatedly until no adsorption was with a FID and a molecular sieve 5A capillary column.
detected. The CO chemisorption was reported as mol CO The exhausted samples (−ex) were obtained by a sec-
adsorbed per gram of sample. ond reaction cycle over the same catalyst. With the pur-
For OSC measurements, the same TPR and CO pose to study the effect of reduction treatments over the
chemisorption line was used. First, the sample (50 mg) was catalytic activity, after the oxidation treatment, Pd/CZS
submitted to the oxidation treatment used in TPR tests. and Pd/CTS were submitted to a flow of a H2 (5%)/Ar
Afterwards, the temperature was decreased up to reduction (60 cm3 min−1  at 300 or 500  C for 1 h before the
temperature (300 or 500  C) in O2 (5%)/He flow. At this reaction. The methane oxidation reaction rate was cal-
temperature, the flow was switched to He (30 cm3 min−1 , culated at 300  C considering the reactor in differential
Praxair, 99.999%) for 30 min. After evacuation, a flow conditions.
of H2 (5.22%)/Ar (60 cm3 min−1  was admitted into the
reactor for 1 min at the reduction temperature, and then
switched to He flow for 10 min at the same temperature. 3. RESULTS AND DISCUSSION
Finally, successive O2 (5%)/He pulses of 40 L (STP) at 3.1. Catalysts Characterization
300  C were injected repeatedly until no changes in the The chemical composition of prepared samples measured
value of O2 peak area was detected. The result is reported by ED-XRF spectroscopy is reported in Table I. The ana-
as Cycle 1. This procedure was repeated 2 or 3 times for lytic composition was close to the nominal one ±10%.
the same sample and the OSC values were reported as The ceria, CZ and CT nominal loading was chosen
cycles 2 and 3, respectively. equal to 17 wt% in order to form a monolayer over the
FTIR spectra of self-supported wafers (≈24 mg cm−2  silica surface.31
were recorded by accumulation of 100 scans with a res- The diffraction patterns are reported in Figures 1 and 2.
olution of 2 cm−1 in a Shimadzu Prestige-21 equipment All samples showed diffraction peaks at 28.5, 33.1, 47.5
provided with a MCT detector cooled with liquid N2 . The and 56.3. The first peak has been attributed to the crys-
wafers were loaded in a moveable glass sample holder, tallographic plane (1 1 1) of fluorite structure of CeO2 ,5
equipped with an iron magnet, inserted in a conventional and the others correspond to (2 0 0), (2 2 0) and (3 1 1)
pyrex-glass cell (NaCl window). The iron magnet allowed CeO2 reflection planes, respectively.56 The slight devi-
the transfer of the catalyst sample from the heating region ation of 28.5 signal to higher angles for CZS sample
to the infrared light beam. Prior to CO adsorption, the (Fig. 1(b)) compared to CS sample (Fig. 1(a)) can be
wafer was dried at 120  C for 1 h in Ar (20 cm3 min−1 , associated to CeO2 –ZrO2 phases reflection with different
Praxair, 99.995%). Then, the sample was submitted to a composition23 that could indicate the CeZr mixed oxide
reduction pre-treatment (H2 , 20 cm3 min−1 , Aga, 99.999%) formation.6 In contrast, a deviation of (1 1 1) CeO2 plane
from 120  C until the reduction temperature (300 or for CTS (Fig. 1(d)) was not observed, suggesting that Tb
500  C) for 1 h at atmospheric pressure, followed by evac- introduction into CeO2 lattice do not generate a noticeable
uation for 30 min in Ar flow at the same temperature. structural change in the CeO2 crystalline lattice that could
After evacuation, the temperature was decreased to RT, and be detected by XRD.

3866 J. Nanosci. Nanotechnol. 17, 3864–3872, 2017


Ferrer et al. Influence of Thermal Treatments on the Reducibility and Catalytic Properties

Table I. Chemical composition (wt.%) determined by ED-XRF spectroscopy, H2 consumption (cm3 H2 g−1 CeO2  and CO/Pd ratio for different
reduction treatments of prepared samples.

H2 consumption CO/Pd ratio


Sample Pd CeO2 ZrO2 Tb4 O7 SiO2 (cm3 H2 g−1 CeO2  (Reduction temperature 300  C or 500  C)

CS – 17.0 – – 83.0 65 –
CZS – 13.09 3.46 – 83.45 48 –
CTS – 11.25 – 9.22 79.53 60 –
Pd/CS 0.45 16.92 – – 82.63 39 0.378 [0.313]
Pd/CZS 0.46 13.03 3.44 – 83.07 37 0.695 [0.501]
Pd/CTS 0.45 11.20 – 9.18 79.17 49 0.646 [0.398]

Notes: Values in square brackets correspond to CO/Pd ratio after reduction at 500  C. No CO adsorption was detected for the supports.

Based on the methodology proposed by Kozlov et al.,32 Figures 3 and 4 show the TPR profiles of calcined and
considering the (1 1 1) peak position, the predominant exhausted (ex) samples, ceria-based oxides and supported
Ce–Zr phase for CZS sample could be Ce08 Zr02 O2 . This Pd catalysts, respectively. The TPR profile of CS sample
result is similar to that obtained for this sample if it is (Fig. 3(a)) reported two signals with maxima at 542 and
considered the complete incorporation of zirconium into 837  C that correspond to the reduction of surface and bulk
the CeO2 lattice (≈Ce073 Zr027 O2 . For CZS-ex sample Ce4+ ions.6 33 CZS (Fig. 3(b)) and CTS (Fig. 3(d)) samples
(Fig. 1(c)), the peak intensity corresponding to the (1 1 1) showed the typical CeO2 reduction signals but at lower
CeO2 plane increased after two reaction cycles, which temperatures, suggesting the Zr and Tb incorporation into
can be related with a better crystallinity possibly due to the CeO2 lattice.6 29
the catalyst sintering. However, for the CTS-ex sample The reduction signal observed at 489  C for CZS sam-
(Fig. 1(e)), no significant changes in the intensity of the ple (Fig. 3(b)), previously assigned to reduction of surface
same peak was registered, indicating a good thermal sta- Ce4+ ions, becomes more intense than the correspond-
bility. Accordingly, Hornés et al.27 reported that doping ing signal for CTS sample (Fig. 3(d)) sample registered
ceria with Tb helps to stabilize to some extent the sys- at 474  C, suggesting that Ce–Zr mixed oxide improves
tems toward thermal sintering. No Pd peaks were detected the redox properties more than Ce–Tb mixed oxide. For
in the XRD patterns (Fig. 2), suggesting that PdO parti- CeTb mixed oxides it has been reported that Tb ions
cles are finely dispersed,5 6 although we cannot exclude could be stabilized in two different oxidation states, i.e.,
that due to the low palladium content any PdO feature is 3+ and 4+ ,26 34 so if part of Tb would be reduced, this
below the detection limit. The changes in the intensity of would affect also the cerium oxidation state and, therefore,
the signal at 28.5 when Pd is incorporated suggest mod- its reduction degree. On the other hand, by comparing the
ifications in the cristallinity of the samples that could be reduction profile of CS and CTS samples (Figs. 3(a, d)),
due to the redispersion of the ceria or mixed oxide when a shift to lower temperatures can be observed, suggest-
this is impregnated with the Pd precursor solution and sub- ing that Tb incorporation improves the redox properties
sequent calcination. of CeO2 .
33.1º
28.5º

47.5º
28.5º

56.3º
33.1º

47.5º

56.3º

(e) (e)
Intensity (a.u.)

(d)
Intensity (a.u)

(d)

(c)
(c)

(b)
(b)

(a)
(a)

25 30 35 40 45 50 55 60 65 70 25 30 35 40 45 50 55 60 65 70
2θ 2θ

Figure 1. X-ray diffraction patterns for (a) CS; (b) CZS; (c) CZS-ex; Figure 2. X-ray diffraction patterns for (a) Pd/CS; (b) Pd/CZS;
(d) CTS and (e) CTS-ex samples. (c) Pd/CZS-ex; (d) Pd/CTS and (e) Pd/CTS-ex samples.

J. Nanosci. Nanotechnol. 17, 3864–3872, 2017 3867


Influence of Thermal Treatments on the Reducibility and Catalytic Properties Ferrer et al.

790
along with reduction of surface Ce4+ ions.33 The signals
440 (e)
observed between 200 and 400  C (inset graphs) corre-
754 spond to reduction of PdOx species strongly interacting
474 with support.6 The negative reduction signals at 44 and
(d)
59  C (Figs. 4(c and e)) is attributed to -PdHx phase
TCD signal (a.u.)

478 decomposition and is related with the presence of large


803 PdOx particles that could be formed after the second reac-
(c) tion cycle.5 35
489
The overall hydrogen consumption registered for the
808
(b) so far discussed mixed oxides and Pd supported catalysts
(Table I) is well comparable with our previous investi-
837 gation on ceria-zirconia mixed oxides, ranging between
(a)
37–65 cm3 H2 g−1 CeO2 .23
542
×½ CO chemisorption results are shown in Table I. No CO
adsorption was detected for supports, so the adsorbed CO
100 200 300 400 500 600 700 800 900 is referred to Pd. For Pd/CZS sample reduced at 300  C,
Temperature (ºC) a CO/Pd ratio of 0.695 was reported while an increase
Figure 3. TPR profiles for (a) CS; (b) CZS; (c) CZS-ex; (d) CTS and in reduction temperature to 500  C decreases this ratio
(e) CTS-ex samples. (0.501). The same trend was registered for Pd/CTS and
Pd/CS samples. Bernal et al.36 attributed similar behavior
The catalytic activity cycles modified the redox prop- to strong metal support interaction effect (SMSI) to explain
erties of CZS-ex (Fig. 3(c)) and CTS-ex (Fig. 3(e)), as the suppress of CO chemisorption with the increase of
it is confirmed by the shift to lower temperature of the reduction temperature due to high electronic mobility of
reduction maxima for exhausted samples, excepted for the reduced cerium particles (Ce3+ ) that could produce a
signal registered at 790  C in CTS-ex sample that was migration of these particles towards Pd metallic phase,
shifted to a higher temperature. The slight reduction sig- forming a thin layer that cover the Pd surface. Therefore,
nals registered for all mixed oxide between 500–700  C it is possible to consider the results of adsorbed CO to
could be related with the reduction of the mixed oxide in the strong interaction of Pd with the redox promoter gen-
different interaction with silica support. For TPR profiles erated by the reduction treatment at 500  C, causing the
of Pd/CZS (Fig. 4(b)) and Pd/CTS (Fig. 4(d)) samples, partial blockage of Pd sites for CO chemisorption.37 How-
an important change over the reducibility was observed ever, when comparing the CO/Pd ratio of Pd/CS, Pd/CTS
with Pd incorporation. Even more, the presence of Tb or and Pd/CZS samples, it is evident that the presence of the
Zr modified the reducibility when comparing TPR profiles mixed oxide can avoid the loss of chemisorption capac-
of Pd/CZS, Pd/CTS and Pd/CS (Fig. 4(a)) samples. The ity at both reduction temperatures. Another hypothesis that
reduction signals registered at T < 200  C can be assigned could explain the decrease in CO/Pd ratio can be the sin-
to reduction of crystalline PdOx particles of different sizes5 tering of Pd particles produced by the increase of reduction
temperature.
OSC values of prepared samples are reported in Table II.
81
821 CTS and CZS supports registered low OSC values for the
(e)
first cycle compared to samples that contain Pd, which
383 showed an increase in OSC values with repetitive cycles.
44 94
758 For CTS sample, the OSC values were similar at both
TCD signal (a.u.)

(d) reduction temperatures, indicating that an increase in the


298
826
reduction temperature up to 500  C did not increase the
76 (c)
343 Table II. Oxygen storage capacity (mol O2 g−1  for different cycles
59 and after reduction at 300 or at 500  C.
82
817
(b) Sample Cycle 1a Cycle 2b Cycle 3c
58 268 760
CTS 4.1 (4.2) –d –d
(a) CZS 1.3 (10.8) 0.4 (27.1) 0.2 (25.4)
Pd/CTS 23.3 (20.9) 24.8 (32.2) 26.6 (34.1)
Pd/CZS 32.5 (30.6) 37.6 (39.6) 40.0 (42.7)
–100 0 100 200 300 400 500 600 700 800 900
Temperature (ºC) Notes: Values in round brackets correspond to OSC value for samples reduced
at 500  C. a Sample submitted to OSC procedure; b Sample resulting of Cycle 1
Figure 4. TPR profiles for (a) Pd/CS; (b) Pd/CZS; (c) Pd/CZS-ex; subjected to a new OSC procedure; c Sample resulting of Cycle 2 subjected to a
(d) Pd/CTS and (e) Pd/CTS-ex samples. new OSC procedure; d Test not carried out.

3868 J. Nanosci. Nanotechnol. 17, 3864–3872, 2017


Ferrer et al. Influence of Thermal Treatments on the Reducibility and Catalytic Properties

oxygen vacancies for this sample. The reduction treat- bands were registered for the support reduced at 300  C:
ment at 300  C for CZS sample did not promote the the first one at 1867 cm−1 with a shoulder at 1994 cm−1
oxygen vacancies formation that could explain the low and the second one at 1625 cm−1 , while for the support
quantity of oxygen chemisorbed, and, hence, after 3 repet- reduced at 500  C a low intensity band at 2003 cm−1 was
itive cycles, the support redox capacity is almost null. obtained. The band observed at 1625 cm−1 for the sup-
When the reduction temperature was increased to 500  C, port reduced at 300  C is attributed to the presence of
higher OSC values were obtained, that can be related with carbonates.38 Thomas et al.39 carried out FTIR-CO tests
a higher reduction grade of mixed oxide and formation over Ce068 Zr032 O2 mixed oxide, reporting bands at 2167
of more oxygen vacancies.33 Pd/CZS sample showed the and 1575 cm−1 correlated to the weak interaction of CO
highest OSC for all cycles and both reduction tempera- with cerium surface sites and carbonate formation, respec-
tures (42.7 mol O2 g−1 for cycle 3). Considering the tively. This result could indicate that the bands at 2003,
possibility of O2 spillover intrinsic to Pd,23 the presence 1994 and 1867 cm−1 corresponds to the weak CO interac-
of Zr could be facilitating the O2 mobility into the CeO2 tion with ceria. Similar bands (2003 cm−1 with a shoulder
lattice in a higher proportion than Tb. It can be observed at 2025 cm−1  were reported for CTS sample reduced at
the slight improvement in OSC values with the number 500  C (Fig. 5(B)-2). When CTS support was reduced at
of cycles, especially for Pd/CZS sample at both reduction 300  C (Fig. 5(B)-1), it is worth noting the low intensity
temperatures, and for CZS and Pd/CTS samples reduced at of the CO absorption bands. This behavior was opposite
500  C. This behavior could be due to an organization of to that registered for CZS support, which reported well-
the solid solution (cubic and tetragonal phases) after suc- defined absorbance signals in both reduction temperatures.
cessive redox treatments and the formation of bulk Ce3+ This result could be suggesting that Tb incorporation gen-
cations involving a thorough reorganization of the mixed erates surface species of different nature to those obtained
oxide network.21 for CZS sample.
FTIR spectra of adsorbed CO in samples reduced at 300 FTIR-CO spectra for Pd/CZS sample previously
or 500  C are reported in Figure 5. Infrared spectra for reduced at 300 or 500  C are shown in Figure 5(C).
CZS sample are shown in Figure 5(A). Two absorption The absorbance band observed at 2157 cm−1 corresponds

Figure 5. IR spectra of CO adsorbed at 25  C for (A) CZS; (B) CTS; (C) Pd/CZS and (D) Pd/CTS samples, reduced at 300  C (1) or 500  C (2).

J. Nanosci. Nanotechnol. 17, 3864–3872, 2017 3869


Influence of Thermal Treatments on the Reducibility and Catalytic Properties Ferrer et al.

to CO adsorbed on cationic Pd species (Pd2+ –CO), the Table III. Methane reaction rate at 300  C (mol CH4 s−1 g−1  and
band at 2085 cm−1 can be assigned to CO species singly T50 of calcined and exhausted catalysts after reduction at 300 (R300) or
at 500  C (R500).
bonded to Pd0 atoms and the bands at 1975 and 1936 cm−1
correspond to bridged CO species bonded to zero-valent rCH4
Pd atoms (Pd02 CO).40 41 In this sample, the increase in Sample T50 ( C) T50 R300 ( C) T50 R500 ( C) (mol CH4 s−1 g−1 )
reduction temperature did not modify the spectra, indicat- CS nr nm nm 55
ing that the surface species have good thermal stability. CTS nr nm nm 110
FTIR-CO spectra for Pd/CTS sample reduced at 300 or CZS nr nm nm 221
500  C (Fig. 5(D)) present absorbance bands at 2083 and Pd/CTS 519 [457] 466 501 74 (258)
2087 cm−1 that correspond to linear CO adsorbed over Pd/CZS 440 [355] 395 498 294 (331)
Pd0 surface atoms. The low spectrum resolution did not Notes: Values in square brackets correspond to T50 of exhausted samples. Values
allow assigning other bands, but some contribution in the in round brackets corresponds to the reaction rate for exhausted samples. nr: Not
reached; nm: Not measured.
range of 2000–1850 cm−1 could be related to bridge CO
between two Pd atoms.42 The surface species present in
Pd/CTS sample have a good thermal stability. sense, large PdOx particles present in exhausted samples
can dissociate methane more readily.35 The methane oxida-
tion reaction rate (rCH4  at 300  C is reported in Table III.
3.2. Catalytic Activity
At this temperature, conversions were lower than 10%,
The curves of methane conversion (%) versus temperature
allowing to compute the reaction rate, because the reactor
for supports, Pd/CTS and Pd/CZS calcined and exhausted
is operating in differential conditions. The rCH4 increased
(ex) catalysts are reported in Figure 6. The supports
when Pd catalysts were exposed to two reaction cycles.
showed relatively low catalytic performance being not able
This trend was more evident in Pd/CTS catalyst, which
to reach 50% of conversion, although at low tempera-
rCH4 enlarged from 74 to 258 mol CH4 s−1 g−1 . For
ture (around 300–350  C) some CH4 conversion (<10%)
Pd/CZS catalyst, the increment in rCH4 was around 12%
was registered. T50 (Temperature corresponding to 50%
(from 294 to 331 mol CH4 s−1 g−1 . These results could
of methane conversion) values of exhausted samples were
indicate the presence of PdOx species of different nature;
lower than those of calcined counterparts (Table III). This
these findings are supported by the changes observed in the
behavior could be related to the formation of large PdOx TPR profiles of calcined and exhausted catalysts (Fig. 4).
particles,5 generated by the decomposition of -PdHx
In addition, the interaction between Pd and Ce would pro-
phase observed in TPR profiles of exhausted samples mote the presence of catalytic species of higher activity.
(Figs. 4(c, e)). Colussi et al.43 conducted tests of methane
Infrared tests showed the existence of this interaction due
combustion in Pd catalysts supported on alumina doped to the presence of absorbance bands above 2100 cm−1 ,
with various rare earths (Ce, La, Pr, Tb) and concluded that
related to the formation of Pd cationic species.
the high activity observed corresponds to the PdO phase, Even though, it has been reported that PdOx species are
which was identified by thermogravimetric analysis (TGA) highly active for methane combustion, the values found for
and temperature programmed oxidation (TPO). In this CZS and CTS supports (221 and 110 mol CH4 s−1 g−1 ,
respectively) indicates an important contribution in the
reaction rate of the corresponding Pd catalyst. Compar-
100 CZS ing the T50 of calcined and exhausted catalysts, it is evi-
CTS
Pd/CZS dent that Pd/CZS catalyst shows better performance than
80 Pd/CTS Pd/CTS catalyst, highlighting the effect of Zr incorpora-
Pd/CZS-ex
CH4 Conversion (%)

Pd/CTS-ex tion into the ceria lattice that generates a high oxygen stor-
15 age capacity, reported for Zr containing samples (Table II).
60
It has been reported that CeZr and CeTb mixed oxides
10
improve catalytic, thermal and redox properties compared
40 5 to pure CeO2 because these mixed oxides present a high
capacity to release oxygen at low temperatures.42 44 In
0
250 350 350 this regard, it is clear the improvement in the methane
20
oxidation reaction rate at 300  C for CeZr and CeTb
mixed oxides compared to CS catalyst (rCH4 = 55 mol
0 CH4 s−1 g−1 .
0 100 200 300 400 500 600 700 The curves of methane conversion (%) versus temper-
Temperature (ºC) ature for catalysts previously reduced before reaction are
Figure 6. Conversion versus temperature curves for methane oxida- reported in Figure 7. None of catalysts previously reduced
tion after oxidation treatment for () CZS; (•) CTS; () Pd/CZS; () at 300 or at 500  C reached 100% conversion. This find-
Pd/CTS; () Pd/CZS-ex and (♦) Pd/CTS-ex catalysts. ing disagrees with the high activity reported for the same

3870 J. Nanosci. Nanotechnol. 17, 3864–3872, 2017


Ferrer et al. Influence of Thermal Treatments on the Reducibility and Catalytic Properties

80
Pd/CZS
increase of reduction temperature could be related with the
A
occurrence of two steps:
(i) The transformation of PdOx to Pd0 , existing a mini-
60 mum PdOx fraction which particle size depends of reduc-
CH4 conversion (%)

tion temperature,
(ii) Afterwards, the high reaction temperature favors a
40 growing of PdOx fraction in the Pd0 /PdOx phase, which
can occur due to the availability of oxygen in the reaction
mixture.
20 Therefore, it is suggested that a similar behavior in Pd/CZS
catalyst can happen, where the reduction treatment and
subsequent exposure to reaction mixture could reoxidize
0 Pd particles, generating species that catalyze the methane
0 100 200 300 400 500 600 700
combustion. It was not possible to reach 100% conver-
Temperature (ºC)
sion for any catalysts under the reduction temperatures
80 employed, while complete conversions were obtained over
B Pd/CTS
the same samples only subjected to oxidation treatment
(Fig. 6). For Pd/CZS and Pd/CTS catalysts previously
60
reduced at 300  C (Table III), the T50 results were substan-
CH4 conversion (%)

tially lower than those reported for Pd/CZS and Pd/CTS,


indicating that the catalytic species generated by the previ-
ous reduction treatment are active but no thermally stable
40
because complete conversions at 700  C were not obtained.

20 4. CONCLUSIONS
XRD results showed that successive oxidation/activity
cycles did not greatly modify the samples crystallinity.
0
The shift of reduction maxima to lower temperatures in
0 100 200 300 400 500 600 700 the mixed oxides and the presence of large PdOx parti-
Temperature (ºC) cles indicated that previous reduction treatment generates
species more easy to reduce with high activity. CO/Pd ratio
Figure 7. Conversion versus temperature curves for methane oxidation of the samples decreased with the increase in reduction
for (A) Pd/CZS and (B) Pd/CTS reduced at 300  C () or at 500  C (•).
temperature, suggesting a Pd encapsulation by ceria and/or
an increase in Pd particles. FTIR-CO results of supports
samples subjected only to oxidation treatment (Fig. 6). reported the presence of carbonates species, while Pd/CZS
In particular, catalysts reduced at 300  C reported higher sample registered bands up to 2100 cm−1 related to CO
conversion than those reduced at 500  C. This tendency adsorbed on Pdn+ species attributed to Pd–CeO2 interac-
could indicate that the reduction treatment at 500  C would tions which could be responsible for the improvement in
produce a migration of Ce3+ ions to Pd particles.33 45 catalytic activity. Pd/CZS sample showed the highest OSC
Pd/CZS catalyst reduced at 300  C shows the lowest T50 and catalytic activity due to the formation of Ce–Zr mixed
(395  C) while for reduction at 500  C a T50 of 498  C oxide that promotes the high oxygen mobility into the
was observed (Table III). Meyer et al.46 have suggested CeO2 lattice. The catalytic activity was enhanced after two
that reduction treatments over a Pd/CeO2 sample before reaction cycles due to the formation of large PdO parti-
methane oxidation reaction modify the nature of Pd parti- cles. T50 values indicated that reduction treatments carried
out before catalytic tests, generated species catalytically
cles. They proposed that before exposure the catalyst to the
more actives than those obtained by the oxidation treat-
reaction mixture, the sample was submitted to an oxida-
ment, however, these active species were not stables at
tion and reduction treatments, that is, at the beginning the
high temperatures.
Pd oxidized was reduced and finally reoxidized when the
catalyst was in contact with the reaction mixture. These Acknowledgments: This research was partially sup-
oxidation-reduction-oxidation cycles can change the nature ported by the Consejo de Desarrollo Científico y
of PdOx towards a more active phase that can generate Humanístico of La Universidad del Zulia (CONDES-
high conversion. LUZ) and Fondo Nacional de Ciencia, Tecnología e Inno-
Considering that the PdOx phase has been reported as vación (FONACIT-Project PEI2011001345). Dr. Francesco
the most active for methane oxidation,47 a possible expla- Giordano (ISMN-CNR) is acknowledged for XRD
nation for the decrease in the catalytic activity with the measurements.

J. Nanosci. Nanotechnol. 17, 3864–3872, 2017 3871


Influence of Thermal Treatments on the Reducibility and Catalytic Properties Ferrer et al.

References and Notes 25. E. Aneggi, C. de Leitenburg, G. Dolcetti, and A. Trovarelli, Catal.
1. J. Q. Wang, M. Q. Shen, Y. An, and J. Wang, Catal. Commun. Today 114, 40 (2006).
10, 103 (2008). 26. M. Balaguer, C. Yoo, H. Bouwmeester, and J. Serra, J. Mater. Chem.
2. R. Heck and R. Farrauto, Appl. Catal. A: Gen. 221, 443 (2001). A 1, 10234 (2013).
3. G. Centi, S. Perathoner, and Z. Rak, Appl. Catal. B: Environ. 41, 143 27. A. Hornés, D. Gamarra, G. Munuera, A. Fuerte, R. X. Valenzuela,
(2001). M. J. Escudero, L. Dazaa, J. C. Conesa, P. Bera, and A. Martínez-
4. S. Manahan, Introducción a la química ambiental, 1st edn., edited Arias, J. Power Sources 192, 70 (2009).
by Reverté, Ciudad de México, México (2007), p. 366. 28. B. M. Reddy, G. Thrimurthulu, and L. Katta, Catal. Lett. 141, 572
5. Z. Zhan, X. Liu, H. He, L. Song, J. Li, and D. Ma, J. Rare Earth (2011).
31, 750 (2013). 29. G. Blanco, J. Pintado, S. Bernal, M. Cauqui, M. Corchado,
6. V. Ferrer, D. Finol, D. Rodríguez, F. Domínguez, R. Solano, A. Galtayries, J. Ghijsen, R. Sporken, T. Eickhoff, and W. Drube,
J. Zárraga, and J. Sánchez, Catal. Lett. 132, 292 (2009). Surf. Interface Anal. 34, 120 (2002).
7. R. Heck and R. Farrauto, Catal. Today 51, 351 (1999). 30. L. F. Liotta, A. Longo, G. Pantaleo, G. Di Carlo, A. Martorana,
8. M. O’Connell and M. A. Morris, Catal. Today 59, 387 (2000). S. Cimino, G. Russo, and G. Deganello, Appl. Catal. B 90, 470
9. W. Liu and M. Flytzanistephanopoulos, J. Catal. 153, 304 (1995). (2009).
10. R. Hicks, H. Qi, M. Young, and R. Lee, J. Catal. 122, 280 (1990). 31. S. Bernal, J. Calvino, G. Cifredo, J. Gatica, J. Pérez Omil, and
11. P. Briot and M. Primet, Applied Catal. 68, 301 (1990). J. Pintado, J. Chem. Soc. Faraday Trans. 89, 3499 (1993).
12. A. Trovarelli, G. Dolcetti, C. de Leitenburg, J. Kaspar, P. Finetti, and 32. A. Kozlov, D. Kim, A. Yezerets, P. Andersen, H. Kung, and
A. Santoni, J. Chem. Soc. Faraday Trans. 88, 1311 (1992). M. Kung, J. Catal. 209, 417 (2002).
13. G. Aguila, F. Gracia, and P. Araya, Appl. Catal. A: Gen. 343, 16 33. H. Yao and Y. Yao, J. Catal. 86, 254 (1984).
(2008). 34. M. Balaguer, C. Solís, and J. Serra, J. Phys. Chem. C 116, 7975
14. V. Ferrer, D. Finol, R. Solano, A. Moronta, and M. Ramos, J. Envi- (2012).
ron. Sci. 27, 87 (2015). 35. S. Yang, A. Maroto, M. Benito, I. Rodríguez, and A. Guerrero, Appl.
15. R. Di Monte, P. Fornasiero, J. Kaspar, P. Rumori, G. Gubitosa, and Catal. B: Environ. 28, 223 (2000).
M. Graziani, Appl. Catal. B: Environ. 24, 157 (2000). 36. S. Bernal, J. Calvino, M. Cauqui, J. Gatica, C. Larese, J. Pérez, and
16. B. M. Reddy, G. Thrimurthulu, L. Katta, Y. Yamada, and S. E. Park, J. Pintado, Catal. Today 50, 175 (1999).
J. Phys. Chem. C 113, 15882 (2009). 37. L. Acuña, F. Muñoz, M. Cabezas, D. Lamas, A. Leyva, M. Fantini,
17. A. Trovarelli, C. De Leitenburg, M. Boaro, and G. Dolcetti, Catal. R. Baker, and R. Fuentes, J. Phys. Chem. C 114, 19687 (2010).
Today 50, 353 (1999). 38. D. Trujillo, F. Ramírez, V. Ferrer, D. Finol, F. Domínguez, and
18. S. Bernal, G. Blanco, M. Cauqui, M. Corchado, C. Larese, M. Ramos, Avances en Química 8, 157 (2013).
J. Pintado, and J. Rodríguez-Izquierdo, Catal. Today 53, 607 (1999). 39. C. Thomas, O. Gorce, C. Fontaine, J. Krafft, F. Villain and G. Djéga,
19. B. Yue, R. Zhou, Y. Wang, and X. Zheng, Appl. Catal. A: Gen. Appl. Catal. B: Environ. 63, 201 (2006).
295, 31 (2005). 40. R. Monteiro, L. Dieguez, and M. Schmal, Catal. Today 65, 77
20. B. M. Reddy, P. Bharali, P. Saikia, S. E. Park, M. W. van den (2001).
Berg, M. Muhler, and W. Gruenert, J. Phys. Chem. C 112, 11729 41. A. Bastos, D. Finol, J. Méndez, F. Domínguez, D. Rodríguez, and
(2008). V. Ferrer, Ciencia 21, 25 (2013).
21. L. Liotta, A. Macaluso, G. Pantaleo, A. Longo, A. Martorana, 42. V. Ferrer and D. Finol, Ciencia 21, 165 (2013).
G. Deganello, G. Marcì, and S. Gialanella, J. Sol–Gel Science and 43. S. Colussi, A. Trovarelli, C. Cristiani, L. Lietti, and G. Groppi, Catal.
Technology 26, 235 (2003). Today 180, 124 (2012).
22. L. Liotta, G. Pantaleo, A. Macaluso, G. Marcì, S. Gialanella, and 44. L. Meng-Fei and Z. Xiao-Ming, Appl. Catal. A: Gen. 189, 15 (1999).
G. Deganello, J. Sol–Gel Science and Technology 28, 119 (2003). 45. L. Liotta, A. Longo, A. Macaluso, G. Pantaleo, A. Venezia, and
23. L. Liotta, A. Macaluso, A. Longo, G. Pantaleo, A. Martorana, and G. Deganello, Appl. Catal. B: Environ. 48, 133 (2004).
G. Deganello, Appl. Catal. A: Gen. 240, 295 (2003). 46. D. Meyer, C. Fraga, A. Alcover, M. Baldini, and F. Zanon, Process
24. P. Fornasiero, G. Balducci, R. Di Monte, J. Kašpar, V. Sergo, Saf. Environ. 87, 315 (2009).
G. Gubitosa, A. Ferrero, and M. Graziani, J. Catal. 164, 173 (1996). 47. S. Oh, P. Mitchell, and R. Siewert, J. Catal. 132, 287 (1991).

Received: 20 May 2016. Accepted: 19 August 2016.

3872 J. Nanosci. Nanotechnol. 17, 3864–3872, 2017

You might also like