Sei sulla pagina 1di 9

Applied Catalysis A: General 495 (2015) 54–62

Contents lists available at ScienceDirect

Applied Catalysis A: General


journal homepage: www.elsevier.com/locate/apcata

Pd catalysts for total oxidation of methane: Support effects


James B. Miller ∗ , Manaswita Malatpure
Department of Chemical Engineering, Carnegie Mellon University, Pittsburgh, PA 15213, United States

a r t i c l e i n f o a b s t r a c t

Article history: We prepared a series of Pd catalysts for total oxidation of methane on different supports: Al2 O3 ,
Received 18 December 2014 ZrO2 –CeO2 and CeO2 . We characterized the catalysts’ structures by XRD and their surface chemistries
Received in revised form 27 January 2015 by temperature programmed reduction (TPR) and oxidation (TPO); and measured their activities for
Accepted 31 January 2015
methane oxidation in temperature programmed reaction experiments (TPRx). By fitting a kinetic model
Available online 7 February 2015
to the TPRx results, we extracted apparent activation barriers (Eapp ) and pre-exponents (Aapp ) for methane
oxidation over each of the catalysts. Activity correlates strongly with Aapp , which we interpret as a relative
Keywords:
measure of the density of PdO–Pd* (* is an O-vacancy) site pairs. TPR results support this view: PdO in low
Supported Pd catalysts
Total methane oxidation
activity (low Aapp ) catalysts, such as Pd/CeO2 , is relatively resistant to reduction, meaning that formation
Temperature programmed reaction of the Pd* partner of the site pair is difficult—and the number of site pairs that can form is low.
Temperature programmed reduction © 2015 Elsevier B.V. All rights reserved.

1. Introduction apparent reaction pre-exponent, which we interpret as a measure


of the relative density of site pairs on the catalyst surface. Results
Supported palladium catalysts for total hydrocarbon oxida- of temperature-programmed reduction experiments support this
tion have important applications in power generation, where they view.
enable methane combustion at relatively low temperature for
avoiding formation of thermal NOx ; in emissions control, includ- 2. Methods
ing automotive catalysts; and as the active elements of sensors
that detect combustible gasses. A number of reviews have been 2.1. Catalyst preparation
published on supported Pd catalysts for methane oxidation [1–5].
At a mechanistic level, the catalytic chemistry of methane oxida- Table 1 lists the catalysts studied in this work. They were pre-
tion on Pd catalysts is complex. Oxidation is usually described as pared by impregnating Al2 O3 (Strem 13-2525, ‘Al’), ZrO2 –CeO2
a Mars–van Krevelen cycle [6], in which a pair of sites—PdO and (Alfa Aesar 39216, ‘ZrCe’) or CeO2 (Alfa Aesar 12925, ‘Ce’) oxides
Pd* (where * is an ‘O-vacancy’)—act in concert to activate the sta- with solutions of Pd(NO3 )2 (10 wt% in 10% HNO3 , Aldrich 380040,
ble C H bond. The activity of the catalyst depends on a number ‘NO3 ’) or Pd(NH3 )4 (NO3 )2 (tetraamine palladium nitrate, 10 wt%,
of factors, but the density and activity of site pairs are among the Aldrich 377384, ‘TAPN’) precursors. For each preparation, approx-
most important [6,7]. These characteristics are functions of the imately 1 g of oxide was first dried at 400 K for 30 min in a vacuum
size, morphology and surface chemistry of PdO crystallites that oven. Then, precursor solution was stirred into the dried powder in
reside on the support surface, which in turn depend on the details 20 ␮L increments until the powder started to become sticky. The
of catalyst preparation [6,8–11], operating conditions [10] and, precursor was decomposed by calcination in a box furnace at 773 K
significantly, the interactions of PdO with its support [7,12–15]. for 1 h; the heating rate of the furnace is approximately 40 K/min.
Different supports can stabilize PdO crystallites in different sizes Table 1 summarizes the Pd contents of the catalysts calculated
and morphologies [16]. The support can even be an active partic- from the amount of precursor solution added to their supports. The
ipant in the catalytic chemistry through exchange of its O atoms 1.0 Pd(NO3 )/Ce sample was prepared by two impregnations, with
with PdO [17,18]. In this work, we compare Pd catalysts pre- an intermediate calcination at 773 K. We also evaluated a com-
pared on three different supports, Al2 O3 , ZrO2 –CeO2 , and CeO2 . mercially available 1% Pd/Al2 O3 (Alfa Aesar, ‘1.0 Pd/Al(AA)’) as a
Microkinetic analyses of the temperature-programmed methane reference material.
oxidation data show that activity correlates strongly with Aapp , the
2.2. X-ray diffraction

∗ Corresponding author. Tel.: +1 920 397 5855; fax: +1 412 268 7139. X-ray diffraction patterns were acquired for the supports and
E-mail address: jbmiller@andrew.cmu.edu (J.B. Miller). the calcined catalysts using a Pananalytical X’pert Pro X-ray

http://dx.doi.org/10.1016/j.apcata.2015.01.044
0926-860X/© 2015 Elsevier B.V. All rights reserved.
J.B. Miller, M. Malatpure / Applied Catalysis A: General 495 (2015) 54–62 55

Table 1 at 10 K/min and then held at 873 K for 15 min before cooling at
Supported Pd catalysts studied in this work. Except for 1.0 Pd/Al(AA), Pd contents are
20 K/min to 300 K in O2 /He flow. The reactor was purged with
calculated from preparation. Pd dispersion and crystallite size are estimated from
the results of pulse CO chemisorption experiments. He and the sample was cooled to 233 K before changing the flow
to 50ccm 10.8%H2 /Ar (Valley National) for TPR. The sample was
Calculated Dispersion Crystallite
heated to 873 K at 10 K/min and then held at 873 for 15 min. Next,
Pd (wt%) size (nm)
the sample was cooled in 10.8% H2 /Ar at 20 K/min to 423 K and
1.0 Pd/Al(AA) 1.02* 0.23 4.8 10 K/min to 288 K. Flow was changed to Ar and a second TPO was
1.4 Pd(NO3 )/Al 1.35 0.21 5.3
conducted. We report results of the TPR and the second TPO only;
1.4 Pd(TAPN)/Al 1.44 0.32 3.6
1.1 Pd(NO3 )/ZrCe 1.13 0.20 5.7 the first TPO served as a pre-treatment to deliver a baseline, oxi-
0.5 Pd(TAPN)/ZrCe 0.54 0.41 2.7 dized state for the TPR experiment.
0.5 Pd(NO3 )/Ce 0.52 0.10 11.1
1.0 Pd(NO3 )/Ce 0.95 – –
3. Results
*
From manufacturer’s certification.
3.1. X-ray diffraction

Figures A1, A2 and A3 display diffraction patterns for Al2 O3 -


diffractometer in theta–2theta (–2) configuration. X-rays with
supported catalysts, ZrO2 –CeO2 supported catalysts and CeO2
1.54 Å wavelength were generated from a Cu source operated at
supported catalysts. No diffraction features that could be indexed
35 kV and 10 mA. Diffraction patterns were recorded over a 2 range
to either Pd or PdO were identified in any of the Pd-catalysts. This
from 10◦ to 90◦ with 0.0525◦ step size.
result reflects small primary crystallite size, or perhaps simply rela-
tively low Pd loadings [19]. Furthermore, Pd addition did not change
2.3. CO pulse chemisorption the structure of any of the of the parent oxides. Al2 O3 and Pd/Al cat-
alysts all displayed diffraction patterns characteristic of ␥-Al2 O3 .
Pd dispersion was characterized by pulse CO chemisorption ZrO2 –CeO2 and Pd/ZrCe catalysts’ diffraction patterns reflect a mix-
using a Micromeretics 2920 instrument. The instrument is config- ture of monoclinic and tetragonal ZrO2 . No features that could be
ured as a flow-through, fixed-bed quartz U-tube reactor enclosed indexed to CeO2 were identified in the patterns of ZrO2 –CeO2 or the
in a high-temperature furnace. Samples are held in place within the Pd/ZrCe catalysts. CeO2 and Pd/Ce catalysts all displayed patterns
reactor by a small quartz wool plug. For the chemisorption exper- characteristic of a highly crystalline CeO2 .
iment, 0.10–0.15 g of catalyst was first pre-treated in 50 ccm 10%
H2 /Ar (Airgas) for 2 h at 523 K and then cooled to 323 K in 50 ccm 3.2. CO pulse chemisorption
He (Airgas). A series CO pulses (9.991% CO/He, Airgas) were injected
into the He flow; a thermal conductivity detector (TCD) monitored Pd dispersions and crystallite sizes of the catalysts are reported
relative CO content of the reactor effluent during each pulse. Injec- in Table 1. Dispersions ranged from 0.10 and 0.41, correspond-
tions were terminated when the size of the detected pulses no ing to crystallite sizes between 2.7 and 11 nm. Dispersions in
longer changed. Dispersion (fraction of Pd atoms at the surfaces this range are typical of supported Pd catalysts made by solu-
of crystallites) and crystallite size were estimated from CO uptake tion impregnation of Pd salts [9,19–22]. For both the Al2 O3 and
assuming adsorption stoichiometry of 1:1 CO:Pd. ZrO2 –CeO2 supports, catalysts prepared from the TAPN precur-
sor display higher dispersion then those prepared from the NO3
2.4. Temperature-programmed oxidation, reduction and reaction precursor. Others have reported dependencies of dispersion on
precursor choice in incipient wetness preparation of supported
Supports and catalysts were characterized by temperature- Pd catalysts [9,22]; while beyond the scope of this work, we note
programmed reaction (TPRx), oxidation (TPO) and reduction (TPR) that the differences likely reflect differences in the decomposition
using a Micromeretics 2950HP instrument. This instrument is also kinetics of the precursors. We measured the dispersion of only
configured as a flow-through, fixed-bed quartz U-tube reactor one of our two ceria-supported catalysts; it displayed the lowest
enclosed in a high-temperature furnace. Samples are held in place dispersion within the sample set. Others have observed a depend-
within the reactor by a small quartz wool plug. A thermocouple is ence of Pd dispersion on support composition [12,20]. Even within
located at the top (inlet) of the catalyst bed. Mass flow controllers cerias, Pd dispersion has been observed to depend on the details
regulate the flow of reaction and treatment gases. An integrated of support morphology and chemistry. For example, Satsuma et al.
TCD reports a signal proportional to the difference between the report that high surface area cerias interact strongly with Pd to
thermal conductivities of the reactor inlet and outlet streams. keep Pd crystallites small [15]. On the other hand, the interaction
Before entering the TCD, TPRx product gas flows through a cold between Pd and highly crystalline (like ours), low surface area,
trap to remove water produced during the oxidation reaction. deeply reduced cerias is weaker, allowing larger Pd structures to
For the TPRx experiment, a ∼0.1–0.2 g sample was first pre- form.
treated in 50 ccm air (zero grade, Valley National) at 873 K for
15 min. After cooling to 323 K, the flow was changed to 50 ccm 3.3. TPRx
2% CH4 in air (Matheson). The catalyst temperature was increased
at 10 K/min to 873 K, where it was held for 15 min; we refer to Figure 1 shows the TPRx results for 1.0 Pd/Al(AA) as a plot of TCD
this as the 1up portion of the experiment. To complete the first signal vs. reaction temperature. The 0.0 to ∼0.05 TCD signal scale
reaction cycle, the catalyst was cooled at 10 K/min to 323 K in the corresponds to fractional CH4 conversions, x, of 0.0–1.0. Results for
CH4 /air mixture (1dn). A second cycle was started immediately the first reaction cycle (1up and 1dn) are plotted in blue, with a
without changing flow from the CH4 /air mixture (2up and 2dn). solid line for the increasing temperature portion of the cycle (1up)
For supported Pd samples, a third cycle, preceded by a new air and a dotted line for the decreasing temperature portion (1dn).
pre-treatment, was also performed (3up and 3dn). The second cycle is plotted in red (2up and 2dn) and the third
For TPO characterization, a ∼0.1 g sample was first cooled to cycle in green (3up and 3dn). For this catalyst, complete conver-
278 K flowing Ar. At 278 K, the flow was changed to 50 ccm of sion is achieved across all three cycles. The catalyst exhibits highest
10.0% O2 /He (Valley National). The sample was heated to 873 K activity—conversion at any reactor temperature is highest—during
56 J.B. Miller, M. Malatpure / Applied Catalysis A: General 495 (2015) 54–62

Fig. 1. TPRx of 1.0 Pd/Al(AA), 0.1002 g sample size, displayed as TCD signal (relative
methane conversion) vs. reactor temperature. Blue lines are for the first reac-
tion cycle (1), red for the second cycle (2) and green for the third (3). Increasing
temperature portions of each cycle (up) are plotted as solid lines; decreasing tem- Fig. 2. TPR of 1.0 Pd/Al(AA). Solid blue line is the signal from the increasing tem-
perature portions (dn) as dotted lines. The catalyst is most active during 1up. A perature portion of the experiment. Features are assigned to PdO reduction (273 K),
‘low-temperature feature’ is observed during 1up and 3up. (Color version available Pd H decomposition (347 K) and interactions of H2 with the apparatus (640 K). Dot-
online.) ted red line is the signal from the decreasing temperature portion of the experiment.
The feature at 335 K corresponds to re-formation of Pd H. (Color version available
online.)
the 1up portion of the experiment. Conversion is significantly lower
in 1dn than in 1up, but activity changes from 1dn through the
end of the experiment are small. The need for supported Pd to
be “conditioned” or “stabilized” by cyclic exposure to different
reaction environments has been widely reported [1,6,10,19,21,23];
during conditioning, PdO crystallites achieve their steady state
morphology and surface chemistry. The 1up and 3up portions of
the experiment—those that were immediately preceded by an air
treatment—display a low-temperature feature at around 480 K. We
will consider the significance of this feature in the Discussion sec-
tion; here, we simply note that it likely reflects a transition in PdO
surface chemistry.
Results for 1.4 Pd(NO3 )/Al appear in Fig. A4. The TPRx response
of this sample is similar to that of 1.0 Pd/Al(AA): complete conver-
sion in the temperature range of the experiment, highest activity
in the 1up portion of the experiment, and appearance of the low-
temperature feature in 1up and 3up. The TPRx response of 1.4
Pd(TAPN)/Al, shown in Fig. A5, is similar to those of the other Pd/Al
catalysts, with the exception that it does not display a ‘condition-
ing period’—CH4 conversions are similar across all three cycles. The
state of the as-prepared catalyst must not be very different from the
steady state at reaction conditions. Complete consideration of sta-
bilization differences between 1.4 Pd(NO3 )/Al and 1.4 Pd(TAPN)/Al
is beyond the scope of this report, but we note that they are likely Fig. 3. TPR (increasing temperature portion of the experiment) of all catalysts stud-
linked to differences in the decomposition/thermal evolution of the ied in this work. Temperatures of features that correspond to the PdO→Pd transition
TAPN and Pd(NO3 )2 precursors. are indicated. Dotted lines are results for the bare oxides.
Figures A6 and A7 display TPRx results for 0.5 Pd(TAPN)/ZrCe
and 1.1 Pd(NO3 )/ZrCe. Like the Pd/Al catalysts, both Pd/ZrCe cat- 3.4. TPO/R
alysts achieve full conversion well before 873 K. Unlike Pd/Al, the
activity of Pd/ZrCe increases with the number of reaction cycles. Figure 2 displays the TPR of 1.0 Pd/Al(AA). Three primary fea-
Neither Pd/ZrCe sample exhibits the low-temperature feature that tures appear. The large positive feature at 273 K is evidence of
was characteristic of the Pd/Al catalysts. hydrogen consumption; it has been linked to reduction of sur-
TPRx results for 0.5 Pd(NO3 )/Ce and 1.0 Pd(NO3 )/Ce appear face PdO [6,12,16]. The negative feature at 347 K is attributable
in Figs. A8 and A9. The Ce-supported catalysts are clearly less to hydrogen evolution upon decomposition of Pd hydride [16].
active than the others—neither delivers complete CH4 conversion The broad feature near 640 K appears in blank runs and most
in the temperature range of the experiment. Both Pd/Ce samples likely results from interaction of H2 with the thermocouple’s metal
exhibit activity loss as the number of reaction cycles increases. Like sheath. During the decreasing temperature portion of the experi-
Pd/ZrCe, the low-temperature feature is not present in TPRx over ment, a hydrogen consumption feature, related to re-formation of
Pd/Ce. the hydride, appears at 335 K.
We performed 2 cycle TPRx experiments on the bare oxides Others have successfully used the low-temperature reduction
(not shown in this report). All three exhibited measurable activ- feature to interpret oxidation activity of supported Pd catalysts
ity at high temperatures. At 873 K, CH4 conversion is ∼30% over [6,12,16]; therefore, in Fig. 3 we compare TPR results for all catalysts
both Al2 O3 and ZrO2 –CeO2 ; and ∼12% over CeO2 . during the increasing temperature portion of the experiment at
J.B. Miller, M. Malatpure / Applied Catalysis A: General 495 (2015) 54–62 57

4. Discussion

4.1. Mirokinetic model of methane oxidation

To formally compare the activities of the catalysts, we fit a


kinetic model to the reaction data. We use the model of Fujimoto
et al. [6], which describes methane oxidation as a sequence of the
following elementary steps:
K1
O2 + ∗ ←→ O2 ∗
k2
O2 + ∗ −→ 2O∗
K3
CH4 + ∗ ←→ CH4 ∗
k4
CH4 ∗ + O∗ −→ CH3 ∗ +HO∗


Fig. 4. TPO of 1 Pd/Al(AA). Solid green line is the signal from the increasing temper- →
ature portion of the experiment. The feature at 559 K is assigned to consumption of → CO2 ∗, CO3 ∗ . . .
O2 during oxidation of surface Pd. Dotted red line is the signal from the decreasing
temperature portion of the experiment. (Color version available online.) K5
2HO∗ ←→ H2 O + O ∗ +∗
K6
temperatures below 500 K. The location of the maximum tempera- CO2 ∗ ←→ CO2 + ∗
ture of the PdO reduction feature depends on the catalysts’ support; K7
reduction maxima increase in the order Pd/Al < Pd/ZrCe < Pd/Ce. CO3 ∗ ←→ CO2 + O∗
Yue, Zhou et al. [12,16] reported a reduction maximum for Pd/Al2 O3
that was ∼24 K below that of Pd/ZrO2 –CeO2 /Al2 O3 ; differences In this sequence, O* represents surface PdO and * represents Pd*,
between our Pd/Al and Pd/ZrCe catalysts are similar. a surface Pd atom or ‘O-vacancy’. Methane adsorbs at a vacancy
Figure 4 is the TPO of 1.0 Pd/Al(AA). The broad negative feature (step 3) and is oxidized in rate-limiting steps 4 as H atoms are
at around 650 K is characteristic of the support itself. Of greater rel- removed from the C atom by successive reaction with O*. Adsorbed
evance to activity is the feature at 559 K, which has been attributed species produced during the oxidation steps (HO*, CO2 * and CO3 *)
to oxidation of Pd to PdO [16,24]. No notable features appear in the are in equilibrium with their gas phase counterparts (steps 5, 6 and
decreasing temperature portion of the experiment. Figure 5 com- 7). Note that the sequence includes cyclic consumption (step 4)
pares the increasing temperature portions of the TPO experiments and regeneration (steps 1–2, 5 and 7) of O*—a Mars–van Krevelen
for all catalysts. Trends are less clear than in the TPR experiments. cycle. The requirement that both types of sites, Pd* for adsorption
1.0 Pd/Al(AA) and the two Pd/ZrCe catalysts display significant oxi- of methane and PdO for oxidation of adsorbed methane, be present
dation features between 500 and 650 K; features below ∼500 K in for the oxidation reaction to proceed is central to understanding
the Pd/ZrCe catalysts are related to their support. Prepared Pd/Al the activity of the catalyst.
and Pd/Ce catalysts have weak or no oxidation features in the Applying the steady state approximation in the case where HO*
500–650 K interval. is the most abundant surface species to the reaction sequence, Fuji-
moto derived a rate law for methane oxidation [6]:

k4 K3 K5 [CH4 ]1.0 [O2 ]0.0 kapp [CH4 ]


r= =
[H2 O]1.0 [H2 O]

This is an interesting result; it underscores the well-known inhib-


iting effect of product water on the reaction kinetics [21].
To fit the rate law to our data, it must be integrated for a plug
flow reactor. Au-Yeung et al. report the result of the integration as
[25]:

2FCH4
 1

kapp = ln −x
ns 1−x

where FCH4 is the methane molar feed rate, x is methane conver-


sion and nS is the number of ‘accessible Pd sites’ in the reactor. We
calculate nS from the catalyst’s Pd content and its Pd dispersion
(Table 1). It is important to recognize that nS contains no informa-
tion on the fraction of surface Pd atoms that participate in active
site pairs, PdO–Pd*; we will return to this point shortly.
We assume that kapp has Arrhenius form:

kapp = Aapp exp(−Eapp /RT )

where Aapp and Eapp are the apparent pre-exponent and activa-
Fig. 5. TPO (increasing temperature portion of the experiment) of all catalysts stud-
ied in this work. Temperatures of features that correspond to the Pd→PdO transition
tion barrier associated with the rate constant. Our strategy for
are indicated. interpreting the reaction data is to calculate kapp as a function
58 J.B. Miller, M. Malatpure / Applied Catalysis A: General 495 (2015) 54–62

Table 2
Kinetic parameters, Eapp and lnAapp , and kapp (at 723 K) for 1.0 Pd/Al(AA). Uncer-
tainties reflect the range of kinetic parameters estimated from replicate TPRx of the
same catalyst.

Eapp (kcal/mol) ln(Aapp (s−1 )) kapp @723 K (s−1 )

Cycle 1, up, low 47 ± 1 45 ± 1 210,000 ± 10,000


temperature
Cycle 1, up, high 45 ± 1 39 ± 1 1500 ± 100
temperature
Cycle 1, down 47 ± 1 38 ± 1 220 ± 10
Cycle 2, up 44 ± 1 36 ± 1 150 ± 10
Cycle 2, down 49 ± 1 39 ± 1 160 ± 10

can display this transient. And furthermore, within this sample set,
the mixed form appears to develop only on samples supported by
Al2 O3 . Pd supported on ZrO2 –CeO2 and CeO2 does not display the
low-temperature feature/discontinuity.
Broad linear regions exist in the Arrhenius plots for all four por-
tions of the 1.0 Pd/Al(AA) TPRx experiment (Fig. 6). For all but the
1up portion of the experiment, we fit the data corresponding to
5–50% conversion. Below 5% the data are noisy; above 50% devi-
ations from linearity are observed as model assumptions break
down and supports become active at higher temperatures. The 1up
Fig. 6. Arrhenius analysis of methane oxidation TPRx results for 1.0 Pd/Al(AA). Blue portion of the experiment has two linear regions, one below and
lines are for the first reaction cycle (1) and red for the second cycle (2). Increasing one above the discontinuity. Table 2 summarizes values of Eapp and
temperature portions of each cycle (up) are plotted as solid lines; decreasing tem-
perature portions (dn) as dotted lines. The catalyst is most active during 1up. The
lnAapp extracted from the Arrhenius plots of Fig. 6; for a simple
‘low-temperature feature’ observed in the raw reaction data (Fig. 1) appears as a activity comparison, we calculated kapp (at 723 K) from lnAapp and
discontinuity in the plot of 1up. (Color version available online.) Eapp . Reported uncertainties reflect the range that we typically see
in repeat runs starting with a new charge of the same catalyst.
In the 1up portion of the experiment, kapp is significantly higher
of temperature and extract values of Eapp and Aapp for each cat- at temperatures below the discontinuity. Because Eapp is the same
alyst from (Arrhenius) plots of lnAapp vs. 1/T. Differences in Eapp on both sides of the discontinuity, higher kapp is attributable exclu-
reflect changes in the temperature dependence of rate constants sively to higher Aapp . As noted earlier, higher Aapp suggests a higher
(step 4) and equilibrium constants (steps 3 and 5)—a measure of density of active site pairs. Thus, this observation is consistent with
the inherent activity of the catalyst. With kapp normalized to the Carstens’ view of transient formation of new reduced Pd sites on
total number of surface Pd sites in the reactor, differences in Aapp the PdO surface at temperatures below the discontinuity, which, in
can reflect differences in fraction of surface Pd atoms that partici- turn, participate in new—and transient—PdO–Pd* active site pairs.
pate in PdO–Pd* pairs: larger Aapp means that a higher fraction of From the first through the second cycle of the experiment, kapp
surface Pd atoms participate in PdO–Pd* pairs. Conversely, lower approaches a constant value as the catalyst achieves its steady state
fractional participation of surface Pd atoms in site pairs means that morphology and surface chemistry. Activity is stable in the sec-
nS over counts the number of active sites, thereby suppressing Aapp . ond cycle, kapp (2up) ∼ kapp (2dn), but there are differences between
the lnAapp and Eapp parameters that contribute to kapp : both are
4.2. Fitting the microkinetic model to the TPRx data lower during the increasing temperature portion of the experiment.
This observation is likely a manifestation of reversible changes in
Figure 6 presents the results of the Arrhenius analysis of the first PdO–Pd’s surface chemistry/morphology that have been linked to a
two TPRx cycles of 1.0 Pd/Al(AA). As before, the first cycle is colour- widely reported temperature hysteresis in the methane oxidation
coded blue (1up and 1dn) and the second cycle red (2up and 2dn); activity of supported Pd and Pt catalysts [3,27,28].
solid lines represent the increasing temperature portion of the cycle We use the 2up portions of the TPRx experiments to com-
(1up and 2up) and dotted lines represent the decreasing tempera- pare the catalysts. This is a convenient choice because, having
ture portion of the cycle (2dn and 1dn). As expected from the raw not been preceded by an air treatment, the low-temperature fea-
results (Fig. 1), the plot for 1up is located at lower temperatures ture/discontinuity is not a factor in the analysis of Pd/Al reaction
(higher 1/T) than the others. data. For Pd/ZrCe catalysts, we pick a linear region between 5 and
The low temperature feature observed in the raw reaction 50% conversion, as we did for Pd/Al. We fit the reaction data from
data (Fig. 1) is easily identified as a discontinuity in the Arrhe- the less-active Pd/Ce catalysts from ∼5 to 30% conversion. Depar-
nius plot. This discontinuity has been observed by others [10,26]; tures from linearity occur at lower conversions for Pd/Ce than in the
Carstens et al. [10] provide perhaps the most complete context for other samples because higher temperatures are required to achieve
it. They suggest that during rapid heating, a transient, reduced Pd them. Table 3 compares lnAapp , Eapp and kapp (723 K) of all catalysts
forms on the surface of PdO crystallites and accelerates the rate of estimated from kinetic analysis of the 2up portions of the TPRx
low-temperature combustion. Then, as temperature increases, the experiments. The rate constants span several orders of magnitude.
reduced Pd is reoxidized and activity returns to ‘normal’ levels. They The samples fall into three groups; in order of decreasing activ-
make the related observation that PdO formed in an oxygen atmo- ity, they are 1.0 Pd/Al(AA) > ‘prepared’ Pd/Al ∼ Pd/ZrCe > PdCe. This
sphere (as in the air pre-treatment that precedes 1up and 3up) is a activity rank-order is consistent with results previously reported
mixture of amorphous and crystalline forms, while PdO formed in a by others [16,22,29].
reaction atmosphere (as in the other portions of our experiment) is Our samples display a range of Eapp , from 34 kcal/mol (Pd/Ce) to
largely crystalline. Our results link these observations, suggesting 44 kcal/mol (1.0 Pd/Al(AA)). Eapp of ∼40–44 kcal/mol (after correc-
that only the mixed form of PdO, the one that is formed in oxygen, tion for inhibition by H2 O) for methane oxidation over Pd/ZrO2 and
J.B. Miller, M. Malatpure / Applied Catalysis A: General 495 (2015) 54–62 59

Table 3 when supported on ZrO2 –CeO2 /Al2 O3 and CeO2 /Al2 O3 than when
Kinetic parameters, Eapp and lnAapp , and kapp (at 723 K) from the 2up portion of the
supported on Al2 O3 alone, confirming the role of ZrO2 –CeO2 and,
TPRx experiments for all catalysts.
in particular, CeO2 in stabilization of PdO.
Eapp (kcal/mol) ln(Aapp (s−1 )) kapp @723 K (s−1 ) Fujimoto et al. [6] also made a link between PdO stability and
1.0 Pd/Al(AA) 44 36 150 morphology, noting that small crystallites, which are more inti-
1.4 Pd(NO3 )/Al 43 33 19 mate contact with their ZrO2 support, can have more stable surface
1.4 Pd(TAPN)/Al 40 30 5.9 Pd O bonds (TPR maxima at higher temperature)—and display
1.1 Pd(NO3 )/ZrCe 38 28 5.3
lower TOF for methane oxidation. Roth et al. [9] made the same
0.5 Pd(TAPN)/ZrCe 38 28 4.8
0.5 Pd(NO3 )/Ce 35 22 0.092 connection between crystallite size and TOF for Pd/Al2 O3 cata-
1.0 Pd(NO3 )/Ce 34 20–21* 0.022–0.088* lysts. We observe a related trend—kapp (Table 3) decreases with
*
Pd dispersion was not measured for this sample. Ranges of A and k were cal- crystallite size (Table 1)—within both our prepared Pd/Al catalysts
culated using the minimum and maximum dispersions measured for the other and within our Pd/ZrCe catalysts, but not across supports. Pd/Ce
samples. presents a unique case; among our catalysts, it displayed the high-
est PdO stability and lowest activity—but the largest Pd crystallite
size. Satsuma et al. provide partial context for this result [15]. They
Pd/Al2 O3 catalysts have been reported by others [6,10,21,23,25,30];
observe that Pd supported by high surface area cerias can be eas-
our estimates for Pd/Al and Pd/ZrCe are consistent with those
ily and completely oxidized to highly dispersed PdO crystallites
reports. Within our sample set, activity differences are driven pri-
or other relatively inactive forms, such as Pd O Ce. On the other
marily by differences in lnAapp —kapp decreases with decreasing
hand, highly crystalline (like ours), low surface area, deeply reduced
lnAapp despite smaller Eapp . Because Aapp provides a relative mea-
cerias interact less strongly with Pd, allowing for growth of larger,
sure of the number of active site pairs, this pattern suggests that the
(partially) metallic structures that can be more active for hydro-
density of site pairs on the catalyst surface is a key determinant of
carbon oxidation. Understanding the relationships between ceria’s
activity. Furthermore, it indicates that 1.0 Pd/Al(AA), with highest
morphology and surface chemistry, and the activities of Pd/CeO2
Aapp , presents the most site pairs, followed by ‘prepared’ Pd/Al and
catalysts for hydrocarbon oxidation is an interesting topic that
Pd/ZrCe. Pd/Ce presents the fewest.
deserves additional investigation.
Our results underscore the role of the support in determin-
4.3. Correlation of PdO stability and catalyst activity ing PdO surface chemistry, and ultimately its activity for methane
oxidation. As we noted earlier, among our catalysts, only Pd/Al
TPR results support the view of site pair density as the primary displayed transient low-temperature activity when pre-treated in
determinant of activity. Figure 7 shows the relationship between air; Al2 O3 demonstrates a unique ability to stabilize the amor-
the location of the TPR maximum and lnAapp : as the reduction tem- phous/crystalline form of PdO that may be responsible for the
perature increases, lnAapp decreases. Pd/Al catalysts display the transient. At higher reaction temperatures, oxidation activity
lowest temperature TPR features, suggesting that creation of sur- depends on the stability of PdO on the catalyst surface. PdO is least
face Pd* from PdO to form the site pairs required for oxidation is stable on Al2 O3 , meaning that conversion of PdO to Pd* for creation
relatively facile. On the other hand, TPR features of the Pd/Ce cat- of PdO–Pd* site pairs is relatively facile—and that Pd/Al catalysts
alysts appear at temperatures more than 40 K higher than Pd/Al, are the most active. PdO is most stable on low-activity Pd/Ce.
consistent with a more stable surface PdO—and more difficult for-
mation of the Pd* partner of the site pair. Pd/ZrCe catalysts display
5. Conclusion
TPR maxima at intermediate temperatures. These results likely
reflect the contributions of lattice oxygen from easily reducible
Supported Pd catalysts display activities for total methane oxi-
supports, like CeO2 and ZrO2 –CeO2 , to PdO stability [31,32].
dation that depend on the composition of the support: among
Fujimoto et al. arrived at a similar conclusion about the relation-
our samples, the observed activity order is Pd/Al > Pd/ZrCe > Pd/Ce.
ship between high PdO stability and low density of PdO–Pd* pairs
Activity is related to the relative density of PdO–Pd* active site
in their study of Pd/ZrO2 catalysts [6]. Zhou et al. [12] and Colussi
pairs on the catalyst surface, which correlates with Aapp . The den-
et al. [14] reported that fraction of Pd present as an oxide is greater
sity of site pairs, in turn, depends on the stability of surface PdO:
if PdO is too stable, than it cannot be partially converted to Pd*
to form the site pairs required for reaction. Among the supports
examined, CeO2 delivers an especially stable PdO, which accounts
for the relatively low activity of the Pd/Ce catalyst.

Acknowledgement

This work was supported by Carnegie Mellon University through


a Berkman Faculty Development grant; by the College of Engineer-
ing through a Dean’s Infrastructure Grant; and by the Department
of Chemical Engineering.
The authors thank George Huber and members of the Huber
group at the University of Wisconsin-Madison for assistance with
CO chemisorption characterization; and Meghan Swanson and
Mark Zanella at Mine Safety Appliances Company for many helpful
discussions.
Fig. 7. The pre-exponent of the oxidation rate constant decreases (the catalyst
becomes less active) as the temperature of PdO reduction in the TPR experiment Appendix A.
increases (PdO becomes more stable). Green triangles are results for 1.0 Pd(NO3 )/Ce
calculated using the minimum and maximum Pd dispersions of the other catalysts.
(Color version available online.) See Figs. A1–A9.
60 J.B. Miller, M. Malatpure / Applied Catalysis A: General 495 (2015) 54–62

Fig. A1. XRD patterns of Al2 O3 and Pd/Al catalysts. All spectra are typical of ␥-Al2 O3 , with no evidence of either Pd or PdO.

Fig. A2. XRD patterns of ZrO2 –CeO2 and Pd/ZrCe catalysts. All patterns reflect a mixture of tetragonal and monoclinic phases of ZrO2 , with no evidence of CeO2 , PdO or Pd.

Fig. A3. XRD patterns of CeO2 and Pd/Ce catalysts. All patterns display the features of CeO2 , with no evidence of either Pd or PdO.
J.B. Miller, M. Malatpure / Applied Catalysis A: General 495 (2015) 54–62 61

Fig. A4. TPRx of 1.4 Pd (NO3 )/Al, 0.1004 g sample size. The catalyst is most active Fig. A7. TPRx of 1.1 Pd (NO3 )/ZrCe, 0.1966 g sample size. Catalyst activity increases
during 1up. A ‘low-temperature feature’ is observed during the 1up and 3up portions throughout the experiment. No ‘low-temperature features’ are observed. (Color
of the experiment. (Color version available online.) version available online.)

Fig. A8. TPRx of 0.5 Pd (NO3 )/Ce, 0.2023 g sample size. The catalyst never achieves
full conversion; activity declines throughout the experiment. No ‘low-temperature
features’ are observed. (Color version available online.)

Fig. A5. TPRx of 1.4 Pd (TAPN)/Al, 0.0982 g sample size. Catalyst activity is
approximately constant throughout the experiment. A ‘low-temperature feature’
is observed during 1up and 3up portions of the experiment. (Color version available
online.)

Fig. A9. TPRx of 1.0 Pd (NO3 )/Ce, 0.0996 g sample size. The catalyst never achieves
full conversion; activity declines throughout the experiment. No ‘low-temperature
features’ are observed. (Color version available online.)

References

[1] Y.-H. Chin, D.E. Resasco, in: J.J. Spivey (Ed.), Catalysis, Royal Scoiety of Chem-
istry, Cambridge, 1999, pp. 1–39.
[2] G. Centi, J. Mol. Catal. A: Chem. 173 (2001) 287–312.
[3] D. Ciuparu, M.R. Lyubovsky, E. Altman, L.D. Pfefferle, A.K. Datye, Catal. Rev. 44
Fig. A6. TPRx of 0.5 Pd (TAPN)/ZrCe, 0.1949 g sample size. Catalyst activity increases (2002) 593–649.
throughout the experiment. No ‘low-temperature features’ are observed. (Color [4] T.V. Choudhary, S. Banerjee, V.R. Choudhary, Appl. Catal. A: Gen. 234 (2002)
version available online.) 1–23.
62 J.B. Miller, M. Malatpure / Applied Catalysis A: General 495 (2015) 54–62

[5] P. Gelin, M. Primet, Appl. Catal. B 39 (2002) 1–37. [19] P. Castellazzi, G. Groppi, P. Forzatti, A. Baylet, P. Marecot, D. Duprez, Catal. Today
[6] K. Fujimoto, F.H. Ribeiro, M. Avalos-Borja, E. Iglesia, J. Catal. 179 (1998) 155 (2010) 18–26.
431–442. [20] G. Groppi, C. Cristiani, L. Lietti, C. Ramella, M. Valentini, P. Forzatti, Catal. Today
[7] S.K. Matam, G.L. Chiarello, Y. Lu, A. Weidenkaff, D. Ferri, Top. Catal. 56 (2013) 50 (1999) 399–412.
239–242. [21] F.H. Ribeiro, M. Chow, R.A. Dalla Betta, J. Catal. 146 (1994) 537–544.
[8] L.M.T. Simplicio, S.T. Brandao, E.A. Sales, L. Lietti, F. Bozon-Verduraz, Appl. Catal. [22] M. Schmal, M.V.M. Souza, V.A. Alegre, M.A.P. da Silva, D.V. Cesar, C.A.C. Perez,
B: Environ. 63 (2006) 9–14. Catal. Today 118 (2006) 392–401.
[9] D. Roth, P. Gelin, A. Kaddouri, E. Garbowski, M. Primet, E. Tena, Catal. Today 112 [23] C.F. Cullis, B.M. Willatt, J. Catal. 83 (1983) 267–285.
(2006) 134–138. [24] S.C. Su, J.N. Carstens, A.T. Bell, J. Catal. 176 (1998) 125–135.
[10] J.N. Carstens, S.C. Su, A.T. Bell, J. Catal. 176 (1998) 136–142. [25] J. Au-Yeung, K. Chen, A.T. Bell, E. Iglesia, J. Catal. 188 (1999) 132–139.
[11] N. van Vegten, M. Maciejewski, F. Krumeich, A. Baiker, Appl. Catal. B 93 (2009) [26] J.G. McCarty, Catal. Today 26 (1995) 283–293.
38–49. [27] R.J. Farrauto, M.C. Hobson, T. Kennelly, E.M. Waterman, Appl. Catal. A 81 (1992)
[12] R. Zhou, B. Zhao, B. Yue, Appl. Surf. Sci. 254 (2008) 4701–4707. 227–237.
[13] S. Colussi, A. Trovarelli, C. Cristiani, L. Lietti, G. Groppi, Catal. Today 180 (2012) [28] A. Amin, A. Abdeli, R. Hayes, M. Votsmeier, W. Epling, Appl. Catal. A 478 (2014)
124–130. 91–97.
[14] S. Colussi, A. Trovarelli, E. Vesselli, A. Baraldi, G. Comelli, G. Groppi, et al., Appl. [29] P.O. Thevenin, A. Alcalde, L.J. Pettersson, S.J. Jaras, J.L.G. Fierro, J. Catal. 215
Catal. A 390 (2010) 1–10. (2003) 78–86.
[15] A. Satsuma, R. Sato, K. Osaki, K. Shimizu, Catal. Today (2012) 61–65. [30] P. Araya, S. Guerrero, F. Robertson, F.J. Gracia, Appl. Catal. A 283 (2005)
[16] B. Yue, R. Zhou, Y. Wang, X. Zheng, Appl. Catal. A: Gen. 295 (2005) 31–39. 225–233.
[17] W.R. Schwartz, L.D. Pfefferle, J. Phys. Chem. C 116 (2012) 8571–8578. [31] C. Bozo, N. Guilhaume, J.-M. Herrmann, J. Catal. 203 (2001) 393–406.
[18] D. Ciuparu, F. Bozon-Verduraz, L.D. Pfefferle, J. Phys. Chem. B 106 (2002) [32] L. Kai, W. Xuezhong, Z. Zexing, W. Xiaodong, W. Duan, J. Rare Earth 25 (2007)
3434–3442. 6–10.

Potrebbero piacerti anche