You are on page 1of 13

Journal of Catalysis 292 (2012) 1325

Contents lists available at SciVerse ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Bimetallic PtPd/silicaalumina hydrotreating catalysts. Part II: Structureactivity


correlations in the hydrogenation of tetralin in the presence of dibenzothiophene
and quinoline
Yanzhe Yu a,1, Benjamin Fonf a,1, Andreas Jentys a, Gary L. Haller a, J.A. Rob van Veen b, Oliver Y. Gutirrez a,
Johannes A. Lercher a,
a
Department of Chemistry and Catalysis Research Center, Technische Universitt Mnchen, Lichtenbergstrasse 4, D-84747 Garching, Germany
b
Shell International Chemicals B.V., Badhuisweg 3, 1031 CM Amsterdam, The Netherlands

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

Article history: The catalytic hydrogenation of tetralin in the absence and presence of quinoline and dibenzothiophene
Received 12 December 2011 was studied on bimetallic PtPd catalysts supported on silica and amorphous silicaalumina (ASA).
Revised 19 March 2012 The proportion of Pt on the surface determined the activity given that the PtPd catalyst with the highest
Accepted 20 March 2012
proportion of surface Pt was the most active. In the absence of poisons, the electronegativity of the sup-
Available online 1 June 2012
port correlated with the hydrogenation. In the presence of quinoline, the activity of the catalysts
increased with the dispersion of the metal particles, whereas in the presence of dibenzothiophene, the
Keywords:
acidity of the support controlled the activity. The observed effects of the poisons indicated the presence
PtPd catalysts
Amorphous silicaalumina
of two kinds of adsorption sites, that is, metal particles and Brnsted acid sites at the perimeter. The key
Tetralin hydrogenation contribution of acid sites for hydrogenation was conrmed by removing the Brnsted acid sites of the
Sulfur poisoning support. Reference Pt catalysts were more active than the PtPd counterparts in poison-free feed. In
Nitrogen poisoning the presence of poisons, the Pt catalysts were also more active than the bimetallic formulations with
the remarkable exception of the bimetallic catalyst with the higher proportion of Pt on the surface.
The bimetallic catalysts are more resistant to sulfur and nitrogen poisoning as well as to sintering. The
poison resistance of bimetallic catalysts originates from the electron transference from Pt to Pd that
yields weak adsorption of poisons on electron decient Pt atoms. However, the surface coverage of Pt
must be maximized to compensate the low activity of Pd-enriched catalysts.
2012 Published by Elsevier Inc.

1. Introduction Two strategies are followed to improve the sulfur and nitrogen
tolerance of the catalyst, that is, using (i) acidic materials as
The growing need to convert heavy crude oils that are richer in supports [4,68] and (ii) noble metal alloys as active catalyst com-
heteroatoms than traditional crudes requires improved hydro- ponent [5,6,9,10]. The use of acidic carriers enhances the hydroge-
treating processes to meet the current and future fuel specica- nation activity of the Pt or Pd catalysts in addition to improving the
tions. In this context, aromatic molecules have to be saturated to sulfur resistance [1113]. The strong acidity of zeolite supports
increase the quality of diesel fuels. However, the presence of sul- leads, however, to excessive cracking and to rapid deactivation
fur- and nitrogen-containing compounds poisons the catalysts best by coke formation [14,15]. Amorphous silicaalumina (ASA) is, in
suited for this task [13]. Therefore, multistage hydrotreating pro- contrast, a more promising support, because it possesses interme-
cesses are typically utilized for upgrading of diesel fuels [4,5]. In diate acidity, which can be tuned further by varying the silica/alu-
the rst stage, the concentration of nitrogen- and sulfur-containing mina ratio [16,17]. Thus, ASA-supported Pt catalysts have been
species is lowered over transition metal sulde catalysts, while in systematically studied with the focus on the hydrogenation of tetr-
the second stage aromatic compounds are hydrogenated on noble alin using dibenzothiophene (DBT) and quinoline as model sub-
metal-based catalysts. However, the activity of the noble metal stances for S- and N-containing poisons [1820]. Two types of
catalysts is drastically reduced by the presence of residual S- and active sites, that is, metal sites alone and metal sites in combina-
N-containing compounds. tion with Brnsted acid sites at the perimeter of the metal clusters,
were identied to be active for the hydrogenation of tetralin on Pt/
ASA catalysts. Sulfur species, resulting from the hydrogenolysis of
Corresponding author. Fax: +49 89 28913544.
dibenzothiophene, selectively poison the metal sites, thus blocking
E-mail address: johannes.lercher@ch.tum.de (J.A. Lercher).
1
These authors contributed equally to this work. the main hydrogenation pathway occurring on metal particles.

0021-9517/$ - see front matter 2012 Published by Elsevier Inc.


http://dx.doi.org/10.1016/j.jcat.2012.03.018
14 Y. Yu et al. / Journal of Catalysis 292 (2012) 1325

However, the hydrogenation of tetralin persists with a lower rate here allowed us to describe the roles of the metal and acid func-
on an active site consisting of acid sites and metal atoms at the tionality as well as the effect of alloying.
perimeter of the metal particles. In this reaction route, the hydro-
carbon molecules adsorb on the acid sites and react with hydrogen 2. Experimental
dissociated on the (partially) sulfur covered metal surface. Quino-
line neutralizes the acid sites, largely reducing the positive effect of 2.1. Catalyst preparation
the Brnsted acid sites. Thus, using acid supports to enhance the
performance of Pt catalysts requires materials with relatively weak A series of PtPd-based catalysts (with a molar ratio of
acid sites, which allow competition between the reacting sub- Pd:Pt = 3) supported on ASA and SiO2 was prepared by the incipi-
strates and sulfur and nitrogen components. ent wetness impregnation method. The ASA materials are denoted
The strategy of alloying showed, for instance, that PtPd cata- as ASA(5/95), ASA(20/80), and ASA(55/45), the numbers in
lysts supported on ASA (Si/Al = 20) are more effective with regard brackets indicate the ratio of alumina/silica in weight percentage.
to increasing the cetane number of real diesel feedstock compared Additionally, a monometallic Pt catalyst (0.8 wt.% of Pt) was syn-
to an industrial reference catalyst [21]. Positive effects of other thesized on ASA with alumina to silica mass ratio of 38/62 to study
bimetallic catalysts for hydrogenation reactions in the presence the role of the Brnsted acidity. A portion of this catalyst was
of sulfur were observed for a variety of oxide supports [2225]. exchanged with Cs+ to fully eliminate its Brnsted acidity. The de-
The improved tolerance of bimetallic PtPd catalysts has been tailed preparation procedure of ASA materials and of the bimetallic
attributed to pronounced metalmetal and metalsupport interac- PtPd catalysts are described elsewhere [18,26].
tions. The electronic effects were claimed to originate from the for-
mation of electron decient Pt species in the bimetallic PtPd 2.2. Catalytic measurements
particles [17]. Alternatively, it has been suggested that Pd contrib-
utes to the hydrogenation activity, while Pt catalyzes hydrogenol- The hydrogenation of tetralin was carried out in a set of 4 par-
ysis [24]. allel trickle-bed reactors in continuous down-ow mode (Fig. 1).
The general agreement is that the sulfur tolerance of bimetallic The catalysts were pressed, ground, and sieved to obtain samples
catalysts is inuenced by the composition and structure of the Pd with particle size in the range 250500 lm. The catalysts samples
Pt clusters as well as by the interaction of the bimetallic particles were diluted with SiC (particle size <250 lm) before loading into
with the supports. However, the impact of sulfur and nitrogen the glass-coated reactors to ensure an isothermal environment
compounds on the hydrogenation activity on the metal and the during the reaction. The catalysts were dried at 275 C for 1 h
acid sites of PtPd catalysts is not understood on a molecular level. and then reduced in situ at 623 K in hydrogen for 2 h before the
In the rst part of this study, a series of PtPd catalysts sup- reaction. The activity of the catalysts was studied at 533593 K
ported on ASA with varying silica to alumina ratios were synthe- and 50 bar of hydrogen pressure. As the conversion of tetralin de-
sized and characterized in detail with respect to the bulk and creased in the presence of poisons, different weight hourly space
surface characteristics of the metal clusters, that is, metal distribu- velocities (WHSVs) were used to maintain tetralin conversion lev-
tion and electronic effects of alloying. The aim of this second part is els in a comparable range, that is, 539 h1 for the clean feed,
to show how the hydrogenation of tetralin in the absence and the 271 h1 for N-poisoned feed, 77 h1 for S-poisoned feed, and
presence of S- and N-containing poisons, that is, DBT and quino- 32 h1 for N- and S-poisoned feed. The WHSV was dened as the
line, is inuenced by the varying acid site and metal surface con- weight of feed per hour and per unit weight of catalyst. The change
centrations. Combining the structural information about the of the WHSV was achieved by varying the amount of catalyst and/
catalysts obtained in part one with the kinetic results presented or the ow rate keeping a constant molar ratio between H2 and

Pressure reducer
H2

Clean N S N+S
MFC Pump
MFC Pump
MFC Pump
MFC Pump
Check valve

He
GC Vent Vent Vent Vent Vent
Vent He He He He He
GC Vent
Vent He
GC Vent
Vent He TBR TBR TBR TBR

GC Vent
Vent
BPR BPR BPR BPR

Box at 323 K

Fig. 1. Schematic representation of the hydrogenation set-up; MFC = mass ow controller; TBR = trickle-bed reactor; BPR = back pressure regulator.
Y. Yu et al. / Journal of Catalysis 292 (2012) 1325 15

tetralin of 20. The poison-free feed consisted of 20 wt.% tetralin, [26]. In order to complement the discussion, the activity of the Pt
11.5 wt.% tetradecane as GC standard, and 68.5 wt.% hexadecane Pd/ASA catalysts was compared to that of the Pt/ASA counterparts.
as solvent. For the catalytic experiments applying quinoline and The characterization of the series of Pt/ASA catalysts was reported
DBT in separate, 400 ppm N and 100 ppm S were used (0.37 and in [20]. For a description of the effect of the Al2O3 and SiO2 compo-
0.058 wt.%, respectively), whereas in the experiments with simul- sition on the properties of the ASA material see Ref. [18].
taneous quinoline and DBT addition, 20 ppm N and 100 ppm S
were used (0.019 and 0.058 wt.%, respectively). The activity of 3.2. Catalytic tests
the catalysts is reported as mol of tetralin converted per hour per
1
gram of catalyst (moltetralin h g1Cat: ). Continuous deactivation was The conversion of tetralin as a function of temperature is pre-
observed in all reactions in the rst 20-h time on stream (TOS). sented in Table 1. Given the variation of WHSV, the conversion val-
Thus, all catalytic activities reported in this paper were measured ues cannot be directly compared. Hence, in the following sections,
at steady state after 24-h TOS. the activity is presented and analyzed as mol of tetralin converted
In order to verify the alloying effect on the sulfur resistance, the per hour and per gram of catalyst. The conversion of tetralin to dec-
hydrogenation of tetralin was carried out on Pt/ASA(20/80) and Pt alin proceeded with high selectivity given that side products, for
Pd/ASA(20/80) in the presence of 100 and 500 ppm S (0.058 and example, ring-opening or cracking products were not observed
0.29 wt.%, respectively). The reactions were carried out in the (the carbon balance in the steady state was 97100% at all
range of 533 to 593 K with a hydrogen pressure of 50 bar and a temperatures).
WHSV of 77.3 h1. All catalysts were deactivated rapidly during the rst 10-h TOS,
The hydrogenation of toluene and isomerization of n-heptane and afterward, the decrease in activity decelerated. After 20-h TOS,
were carried out with Pt/ASA(38/62) and CsPt/ASA(38/62) to the activity in all catalysts was sufciently stable to gain reliable
investigate the impact of the Cs+-exchange on the metal and acid information. As an example, Fig. 2 shows the activity of PtPd/
functions of the catalysts. For the former reaction, samples of SiO2 along with the TOS for all conditions studied, that is, 533
50 mg of the catalysts were loaded in the reactor and the 593 K in the presence and the absence of poisons. The deactivation
hydrogenation pressure was kept at 1 bar. The reaction tempera- within the rst 20-h TOS is attributed to coke formation as con-
ture was varied to obtain 40% conversion of toluene on the two cluded in Ref. [20] and indicated by the characterization of the
catalysts. The isomerization of n-heptane to iso-heptane was used catalyst (vide infra).
carried out at 30 bar H2, whereas the temperature was varied to
obtain 40% n-heptane conversion. 3.3. Hydrogenation of tetralin on PtPd catalysts in the absence of
In order to study the role of acid sites, the monometallic Pt/ poisons
ASA(38/62) and CsPt/ASA(38/62) catalysts were also tested in
the hydrogenation of tetralin in the absence and the presence of The catalytic hydrogenation of tetralin was studied between
quinoline and DBT at the same conditions used with the bimetallic 533 and 593 K. The catalytic activities at steady state after 24-h
catalysts. TOS for the different temperatures are compiled in Fig. 3a. The
PtPd/ASA(55/45) catalyst had the highest activity at all tempera-
2.3. Characterization of the catalysts after reaction tures followed by PtPd/ASA(5/95), PtPd/ASA(20/80) and PtPd/
SiO2. Accordingly, the apparent activation energies determined in
Transmission electron microscopy (TEM) was applied to inves- the hydrogenation of tetralin in the absence of poisons were 54,
tigate the metal particle size after reaction. The catalysts were 48, 46, and 36 kJ mol1 for the catalysts supported on SiO2,
washed in hexane, separated from the SiC by sieving, and dried ASA(5/95), ASA(20/80), and ASA(55/45), respectively. The ratios
at 353 K. The samples were then ground and ultrasonically dis- of cis- to trans-decalin observed with the bimetallic catalysts were
persed in ethanol. Drops of the dispersions were applied on copper below 0.4 (see Fig. 3b), which is signicantly lower compared to
grids with carbon lm. A JEM-2010 JEOL transmission electron the monometallic Pt catalysts (higher than 1.0) and indicates a
microscope operating at 120 kV was used. weaker adsorption of tetralin on the bimetallic particles.
Samples of selected catalysts after use were investigated by Ra-
man spectroscopy. The spectra were obtained using a Renishaw 3.4. Hydrogenation of tetralin on PtPd catalysts in the presence of
Raman Spectrometer (Type 1000), equipped with CCD detector quinoline
and a Leica microscope, equipped with a 514 nm Ar laser. Calibra-
tion was made with a Si(1 1 1) crystal prior to the measurements. The rates of hydrogenation of tetralin and the cis-/trans-decalin
The wavenumber accuracy was within 1 cm1. ratios in the presence of quinoline at steady state are shown in
Fig. 4. The PtPd catalysts showed a signicantly lower activity com-
3. Results pared to the reactions conducted in the absence of quinoline. PtPd/
ASA(55/45) had the highest activity at all temperatures followed by
3.1. Characterization of the catalysts the PtPd/SiO2, PtPd/ASA(5/95), and PtPd/ASA(20/80) catalysts.
The apparent activation energies determined in the hydrogenation
The synthesis and characterization of the PtPd/ASA catalysts of tetralin in the presence of quinoline were 94, 84, 74, and
here studied were described in detail in the rst part of this work 65 kJ mol1 for the catalysts supported on SiO2, ASA(5/95),

Table 1
Conversion of tetralin obtained as function of temperature over the bimetallic catalysts in the absence of poisons (clean feed, WHSV = 539 h1), in the presence of quinoline
(WHSV = 271 h1), DBT (WHSV = 77 h1) and quinoline and DBT (WHSV = 32 h1) after 24 h TOS.

Clean feed Quinoline DBT Quinoline and DBT


Temperature (K) 533 553 573 593 533 553 573 593 533 553 573 593 533 553 573 593
PtPd/ASA(55/45) 10.6 15.4 18.8 19.8 3.3 6.3 9.3 16.2 7.0 12.9 29.1 53.7 0.9 1.8 5.3 25.1
PtPd/ASA(20/80) 5.9 10.0 12.1 14.1 0.8 1.4 2.6 4.1 1.9 3.6 7.6 20.6 0.9 1.6 3.1 13.4
PtPd/ASA(5/95) 8.2 13.9 17.3 18.5 0.7 1.3 2.5 4.3 1.5 2.6 5.6 13.6 0.3 0.5 1.0 4.0
PtPd/SiO2 3.4 6.1 8.1 9.0 1.5 3.4 3.7 9.9 2.5 3.4 5.0 7.2 1.3 2.2 4.1 8.4
16 Y. Yu et al. / Journal of Catalysis 292 (2012) 1325

300 300
533 K

225 225 553 K

573 K
150 150 593 K
Activity [moltetralinh-1gCat.-1] 103

75 75

(a) (b)
0 0
0 10 20 30 40 50 0 10 20 30 40 50

20 20

15 15

10 10

5 5

(c) (d)
0 0
0 10 20 30 40 50 0 10 20 30 40 50
TOS [h] TOS [h]

Fig. 2. Activity of PtPd/SiO2 for tetralin hydrogenation with time on stream (TOS) in the temperature range 533593 K in the absence of poisons (a), in the presence of
quinoline (b), dibenzotiophene (c) and in the simultaneous presence of quinoline and dibenzothiophene (d).

300 0.5
Activity [moltetralinh-1gCat.-1] 103

Pt-Pd/ASA(55/45)
Pt-Pd/ASA(20/80)
Cis/trans decalin ratio

Pt-Pd/ASA(5/95)
200 Pt-Pd/SiO2 0.4

100 0.3

(a) (b)
0 0.2
533 553 573 593 533 553 573 593
Temperature [K] Temperature [K]

Fig. 3. Activities of PtPd catalysts (a) and cis-/trans-decalin ratios (b) observed at steady state for the hydrogenation of tetralin in the temperature range 533593 K (20 wt.%
tetralin, 50 bar, 24-h TOS, WHSV = 539 h1).

ASA(20/80), and ASA(55/45), respectively. The cis-/trans-decalin poisons, leading to a remarkable increase in the reaction rate be-
ratios decreased with all catalysts with increasing temperature. tween 573 and 593 K is (see Fig. 5a). The cis-/trans-decalin ratio
(see Fig. 5b) decreased with temperature as in the case of the
reactions performed in the presence of quinoline. Furthermore,
3.5. Hydrogenation of tetralin in the presence of DBT
the cis-/trans-decalin ratios observed with bimetallic catalysts,
were signicantly lower than those obtained with monometallic
The activities for the tetralin hydrogenation and the cis-/trans-
Pt samples in previous studies [20].
decalin ratios in the presence of DBT at steady state are summa-
rized in Fig. 5. In these conditions, the catalytic activities were
drastically lower than in the absence of DBT. PtPd/ASA(55/45) 3.6. Hydrogenation of tetralin in the presence of quinoline and DBT
showed the highest activity. The apparent activation energies
determined in the hydrogenation of tetralin in the presence of The reaction rates of tetralin hydrogenation and cis-/trans-dec-
DBT were 70, 81, 86, and 89 kJ mol1 for the catalysts supported alin ratios in the presence of N- and S-poisons are shown in Fig. 6.
on SiO2, ASA(5/95), ASA(20/80), and ASA(55/45), respectively. The The catalytic activity drastically decreased compared to the reac-
presence of sulfur induced a stronger dependence of the hydroge- tions with poison-free, quinoline-, or DBT-containing feeds. The
nation reaction rates on the temperature compared to the reac- apparent activation energies determined in the hydrogenation of
tions performed in the presence of quinoline or in the absence of tetralin in the presence of DBT and quinoline were 75, 83, 81,
Y. Yu et al. / Journal of Catalysis 292 (2012) 1325 17

80 0.5

Activity [moltetralinh-1gCat.-1] 103


Pt-Pd/ASA(55/45)
Pt-Pd/ASA(20/80)

Cis/trans decalin ratio


60 Pt-Pd/ASA(5/95)
Pt-Pd/SiO2 0.4

40

0.3
20

(a) (b)
0 0.2
533 553 573 593 533 553 573 593
Temperature [K] Temperature [K]

Fig. 4. Activities of PtPd catalysts (a) and cis-/trans-decalin ratios (b) observed at steady state for the hydrogenation of tetralin in the presence of quinoline in the
temperature range 533593 K (20 wt.% tetralin, 0.37 wt.% quinoline, 50 bar, 24-h TOS, WHSV = 271 h1).

80 0.6
Activity [moltetralinh-1gCat.-1] 103

Pt-Pd/ASA(55/45)
Pt-Pd/ASA(20/80)

Cis/trans decalin ratio


60 Pt-Pd/ASA(5/95) 0.5
Pt-Pd/SiO2

40 0.4

20 0.3

(a) (b)
0 0.2
533 553 573 593 533 553 573 593
Temperature [K] Temperature [K]

Fig. 5. Activities of PtPd catalysts (a) and cis-/trans-decalin ratios (b) observed at steady state for the hydrogenation of tetralin in the presence of DBT in the temperature
range 533593 K (20 wt.% tetralin, 0.058 wt.% DBT, 50 bar, 24-h TOS, WHSV = 77 h1).

40 0.5
Activity [moltetralinh-1gCat.-1] 103

Pt-Pd/ASA(55/45)
Pt-Pd/ASA(20/80)
Cis/trans decalin ratio

30 Pt-Pd/ASA(5/95)
Pt-Pd/SiO2 0.4

20

0.3
10

(a) (b)
0 0.2
533 553 573 593 533 553 573 593
Temperature [K] Temperature [K]

Fig. 6. Activities of PtPd catalysts (a) and cis-/trans-decalin ratios (b) observed at steady state for the hydrogenation of tetralin in the presence of quinoline and DBT in the
temperature range 533593 K (20 wt.% tetralin, 0.019 wt.% quinoline, 0.058 wt.% DBT, 50 bar, 24-h TOS, WHSV = 32 h1).

and 83 kJ mol1 for the catalysts supported on SiO2, ASA(5/95), PtPd/ASA(20/80) catalyst was compared to evaluate the relative
ASA(20/80), and ASA(55/45), respectively. The observed cis-/ sulfur resistance of Pt and PtPd catalysts. Fig. 7 shows that the
trans-decalin ratios were generally below 0.4 and decreased with monometallic Pt catalyst has a higher hydrogenation activity than
temperature over all catalysts. the bimetallic PtPd/ASA(20/80) catalyst regardless of the sulfur
concentration in the feed. However, it is remarkable that the
3.7. Impact of alloying with Pd on the resistance toward S-poisoning increase in sulfur concentration inuenced the activity of the
and characterization after activity tests bimetallic sample only slightly, while it led to a strong activity
reduction for the Pt catalyst.
The hydrogenation of tetralin in the presence of 100 and Samples of fresh and used PtPd/ASA(5/95) were analyzed by
500 ppm S on the monometallic Pt/ASA(20/80) and bimetallic transmission electron microscopy to explore sintering of metal
18 Y. Yu et al. / Journal of Catalysis 292 (2012) 1325

The detection of carbon concentrations between 1.5 and 8 wt.%


Activity [moltetralinh-1gCat.-1] 103

100
Pt/ASA(20/80) (depending on the feed) indicated the formation of coke. Accord-
Pt-Pd/ASA(20/80) ingly, the Raman spectra of the PtPd/ASA(20/80) catalyst (not
75 0.058 wt.% S shown) used in the reaction in the presence of quinoline and
0.29 wt.% S
DBT exhibited a signal at 1628 cm1 assigned to benzenoid units
in polyaromatic structures [30,31]. No other signals were detected
50
due to the intense uorescence caused by the silica support.
Interestingly, after reactions in the presence of quinoline the
25 carbon content was four times higher than in the absence of quin-
oline. As an example, after the reaction in poison-free feed, in Pt
Pd/ASA(20/80) the content of carbon and hydrogen was 1.5 and
0 1.4 wt.%, respectively. After the reaction in the presence of quino-
533 553 573 593 line and DBT, in PtPd/ASA(20/80) the content of carbon, hydrogen,
Temperature [K] nitrogen and sulfur was 7.3, 1.9, 0.3, and 0.1 wt.%, respectively.
Hence, the strong interaction of quinoline with the catalysts led
Fig. 7. Comparison of different DBT poisoning levels on Pt/ASA(20/80) and PtPd/
to the accumulation of N-containing coke. In the catalysts used
ASA(20/80) for tetralin hydrogenation (20 wt.% tetralin, 50 bar, 24-h TOS,
WHSV = 77.3 h1). in the presence of DBT, the S/(Pt + Pd) molar ratio was approxi-
mately 0.5. This low sulfur to noble metal ratio allowed us to dis-
count suldation of the metal phase during the reaction. It is likely
particles during the reaction. Representative images of the catalyst
that the sulfur found in the used catalysts was strongly adsorbed
used for the hydrogenation of the poison-free, quinoline-, DBT as
on the surface of the metal phase and, to a minor extent, contained
well as the quinoline and DBT-containing feed are compiled in
in coke molecules [20].
Fig. 8. The number average particle size for all used catalysts was
1.6 nm, which is the same number average size determined for
the as-synthesized catalyst. Note that the high stability of the 3.8. Role of Brnsted acid sites
bimetallic catalyst contrasts with the lower stability of monome-
tallic Pt catalysts, where the particle size increased from 0.8 to The role of Brnsted acid sites in the hydrogenation of aromatic
1.31.6 nm during the reaction [20]. This is in line with the high compounds was studied by comparing the activity of monometal-
stability found for bimetallic PdPt particles supported on Y zeo- lic Pt/ASA(38/62) and CsPt/ASA(38/62) catalysts. In order to verify
lites [27]. Consequently, we can conclude that the catalyst deacti- that the Pt particles were not affected by the ion exchange with
vation does not originate from sintering of the metal particles, as Cs+, the hydrogenation of toluene, a metal-catalyzed reaction,
proposed elsewhere [28,29]. and n-heptane isomerization, which takes place on acid sites and

No poison
No poison N poison
N poison

S poison
S poison N+S poison
N+S poison

Fig. 8. TEM micrographs of PtPd/ASA(5/95) before and after the hydrogenation of tetralin in the absence and presence of quinoline and/or DBT.
Y. Yu et al. / Journal of Catalysis 292 (2012) 1325 19

800 400
Pt/ASA(38/62) (a) Pt/ASA(38/62) (b)
Cs-Pt/ASA(38/62) 300 Cs-Pt/ASA(38/62)
600

200
Activity [moltetralinh-1gCat.-1] 103

400
100

200 0
533 553 573 593 533 553 573 593

100 60
Pt/ASA(38/62) (c) Pt/ASA(38/62) (d)
75 Cs-Pt/ASA(38/62) 45 Cs-Pt/ASA(38/62)

50 30

25 15

0 0
533 553 573 593 533 553 573 593
Temperature [K] Temperature [K]

Fig. 9. Comparison of the hydrogenation activities of CsPt/ASA(38/62) and Pt/ASA(38/62): (a) in absence of quinoline or DBT, (b) in the presence of quinoline (0.37 wt.%), (c)
in the presence of DBT (0.058 wt.%) and (d) in the presence of quinoline (0.019 wt.%) + DBT (0.058 wt.%) in the temperature range 533593 K (20 wt.% tetralin, 50 bar, 24-h
TOS, WHSV = 539, 271, 77, 32 h1 for clean feed, quinoline-, DBT-, and (DBT + quinoline)-containing feed).

on the metal, were carried out on Pt/ASA(38/62) and on CsPt/ 62) and CsPt/ASA(38/62) in order to emphasize the importance
ASA(38/62). For the hydrogenation of toluene, the temperature re- of Brnsted acid sites for the hydrogenation of tetralin. The differ-
quired for reaching 40% conversion was almost the same, that is, ences observed in the catalytic behavior of both materials are
337 K on Pt/ASA(38/62) and 332 K on CsPt/ASA(38/62). In con- attributed to the replacement of Brnsted acid sites by Cs+ cations.
trast, the temperature required to reach 40% conversion in the This attribution assumes that the metal function was not affected
isomerization of n-heptane had to be increased from 626 K on Pt/ by the Cs-exchange as concluded from the results of hydrogenation
ASA(38/62) to 692 K on CsPt/ASA(38/62). Therefore, it can be con- of toluene and isomerization of n-heptane. For this discussion, the
cluded that the Pt particles were not affected by the ion exchange activities of Pt/ASA(38/62) and CsPt/ASA(38/62) were normalized
with Cs+, while the acid sites were effectively removed and the to the activity of Pt/ASA(38/62) at each temperature. In Fig. 10, the
activity remaining can be solely related to Pt. gray area correspond to the activity of CsPt/ASA(38/62), whereas
The activity of CsPt/ASA(38/62) and Pt/ASA(38/62) for hydro- the white areas represent the difference in activity between Pt/
genation of tetralin in the absence and presence of quinoline ASA(38/62) and CsPt/ASA(38/62).
and/or DBT is shown in Fig. 9. In all cases, the activity of the Cs- In the absence of poisons, CsPt/ASA(38/62) was less active
modied catalyst was lower than that of the unmodied material than Pt/ASA(38/62) increasing the difference from 20% to 33% with
and the difference became larger at increasing temperatures. In increasing temperature in the range 533593 K. The higher activity
the absence of poisons (Fig. 9a), the activity of CsPt/ASA(38/62) of Pt/ASA(38/62) is attributed to the participation of Brnsted acid
was much lower than that observed for Pt/ASA(38/62). In the pres- sites in the reaction. The comparison of the rates shows that up to
ence of quinoline (Fig. 9b), the activity decreased considerably one third of the overall hydrogenation activity is related to the
compared to the poison-free feed, while the activity of the CsPt/ molecules adsorbed on Brnsted sites at the perimeter of the Pt
ASA(38/62) catalyst differed only little from the activity observed clusters.
on Pt/ASA(38/62) at 533 and 553 K. In the presence of DBT In the presence of quinoline, the activity of both catalysts de-
(Fig. 9c), the activity of both catalysts decreased strongly and the creases. However, the difference in activity between Pt/ASA(38/
differences in activity of both catalysts increased largely with tem- 62) and CsPt/ASA(38/62) is marginal at 533 and 553 K, that is,
perature. In the presence of both poisons (Fig. 9d), the activity of only 510%. This nearly equal activity allows us to conclude that
the catalysts is negligible at 533 K and 553 K, and at higher tem- quinoline neutralizes the Brnsted acid sites in Pt/ASA(38/62),
peratures, the activity of the Cs-exchanged catalyst remains much while it affects the metal particles in both catalysts through the
lower than that of Pt/ASA(38/62). same mechanism.
In contrast, in the presence of DBT, Pt/ASA(38/62) was clearly
4. Discussion more active than CsPt/ASA(38/62) showing a difference of around
70% over the whole temperature range. Thus, the absence of
4.1. On the role of Brnsted acid sites in the hydrogenation of tetralin Brnsted acid sites in CsPt/ASA(38/62) led to an activity which
was only a third of that of Pt/ASA(38/62).
Before commenting on the results of the bimetallic PtPd cata- In the mixed-poison experiment, quinoline (20 ppm N) was
lysts, we would like to discuss the results obtained on Pt/ASA(38/ added to the sulfur-containing feed. This small concentration of
20 Y. Yu et al. / Journal of Catalysis 292 (2012) 1325

100 100

75 75

50 50
Cs-Pt/ASA (38/62) and Pt/ASA (38/62) [%]
Normalized hydrogenation activities of

25 25

(a) (b)
0 0
533 553 573 593 533 553 573 593

100 100

75 75

50 50

25 25

(c) (d)
0 0
533 553 573 593 533 553 573 593
Temperature [K] Temperature [K]

Fig. 10. Comparison of the hydrogenation activity of Pt/ASA(38/62) and Cs-Pt/ASA(38/62). The activity of Pt/ASA(38/62) is normalized to 100%. The gray areas correspond to
the relative activity of Cs-Pt/ASA(38/62). The reactions were performed in absence of quinoline and DBT (a); in the presence of quinoline (0.37 wt.%) (b); in the presence of
DBT (0.058 wt.%) (c); and in the presence of quinoline (0.019 wt.%) and DBT (0.058 wt.%) (d).

quinoline poisoned a certain fraction of Brnsted acid sites in Pt/ cules on the metals lead to different selectivities and cause the Pt-
ASA(38/62) decreasing the difference in activity between both cat- based catalysts to be more active than the Pd containing catalysts
alysts to 50%. for the hydrogenation of benzene, toluene, and naphthalene [32
In general, the activity difference between Pt/ASA(38/62) and 36]. All bimetallic PtPd catalysts show intermediate activity and
CsPt/ASA(38/62) increased with increasing temperature (the selectivity, which is usually rationalized as dilution of the more ac-
activity of the former catalyst was always higher). This tendency tive Pt by Pd [37]. Interestingly, under reaction conditions in the
is attributed to lower coverage with sulfur compounds on metal presence of sulfur-containing compounds and with strongly acidic
particles as well as a lower concentration of basic nitrogen-con- supports, the bimetallic PtPd catalysts maintain higher activities
taining compounds on the Brnsted acid sites of Pt/ASA(38/62) as than the corresponding monometallic catalyst [38,39]. However,
the temperature increased. as we show in this work, alloying of Pt with Pd does not necessarily
lead to catalysts with better catalytic performance.
4.2. Hydrogenation of tetralin in the absence of poisons Let us recall at this point the characteristics of the bimetallic
catalysts as described in the rst part of this work. A summary of
The reaction temperature used in the catalytic tests is below the the main properties of the bimetallic catalysts reported in Ref.
temperature of that used during the preparation steps. Therefore, [19,26] is compiled in Table 2. The ASA-supported PtPd catalysts
in the following discussion, we assume that the surface composi- contain PtPd and Pd nanoparticles with number average diameter
tion of the catalysts does not change under reaction conditions in varying between 1.4 and 1.8 nm as determined by TEM. EXAFS
such a way as to vitiate the conclusions. This statement is sup- analysis demonstrates that the PtPd particles contain a Pt-en-
ported by the facts that (i) the number average particle size of riched core and a Pd-enriched shell and have Pd/Pt molar ratios
the metal clusters is conserved after reaction (vide supra) and (ii) from 1.47 to 1.27. The presence of monometallic Pd particles is de-
the deactivation within the rst 20-h TOS can be attributed to coke rived from the process of modeling the bimetallic clusters. Despite
formation as indicated by the elemental analysis and Raman the Pd-rich surface, all bimetallic particles have a variable fraction
characterization. of exposed Pt atoms available on the surface, decreasing in the or-
Turnover frequencies (TOF) were calculated from the conver- der PtPd/ASA(55/45) > PtPd/SiO2 > PtPd/ASA(5/95) > PtPd/
sion and the dispersion of the catalysts for all reactions [26]. The ASA(20/80). The materials have only weak Brnsted acidity. The
respective graphs are shown in the Supplementary information. concentration of Brnsted acid sites increases in the sequence
With only one notable exception (vida infra), the trends are pre- PtPd/SiO2 < PtPd/ASA(55/45) < PtPd/ASA(5/95) < PtPd/
served in all cases; therefore, in the following we discuss the activ- ASA(20/80). The concentration of Lewis acid sites (weak and
ity as it is reported in the Results section, that is, on a weight basis. strong) increases in the order PtPd/SiO2 < PtPd/ASA(5/95) < Pt
The catalytic properties of monometallic Pt and Pd catalysts dif- Pd/ASA(20/80) < PtPd/ASA(55/45). In the following, we discuss
fer markedly. The different adsorption strengths of aromatic mole- rst the effect of alloying PtPd on the hydrogenation activity of
Y. Yu et al. / Journal of Catalysis 292 (2012) 1325 21

Table 2
Characterization of the supported PtPd catalysts: chemical composition, Brnsted (BAS) and Lewis (LAS) acid sites concentration, size and proportion of surface Pt for the
supported metal particles and intermediate Sanderson electronegativity of the supports (Sint) [19,26].

Catalyst Composition (wt.%) Acidity (mmol g1) Particle size (nm) Surface Pt (%) Sint
Pt Pd SiO2 Al2O3 BAS LAS
PtPd/ASA(55/45) 0.30 0.55 55.2 44.0 22 192 1.4 16.9 2.86
PtPd/ASA(20/80 0.30 0.52 20.2 79.0 58 183 1.8 13.9 2.98
PtPd/ASA(5/95) 0.29 0.50 5.4 93.8 45 156 1.6 14.0 3.04
PtPd/SiO2 0.30 0.52 0 99.2 0 0 1.4 14.5 3.05

the catalysts in the poison-free feed to differentiate the decisive electronegativity of the support determines the activity of the cat-
parameters for the activity in the presence of DBT and quinoline. alyst. The adsorption of reactants on Pt becomes stronger with
The experimental activation energies (Ea,Exp) determined for the increasing electronegativity (intermediate Sanderson electronega-
series of bimetallic catalysts in the hydrogenation of tetralin in the tivity) of the carrier. Thus, the activity of the catalyst increases
absence of poisons are 54, 48, 46, and 36 kJ mol1 for the catalysts with increasing SiO2 content in the ASA support. However, the cat-
supported on SiO2, ASA(5/95), ASA(20/80), and ASA(55/45), respec- alyst supported on SiO2 is the least active, because the absence of
tively. The values determined for the same reaction carried out LAS blocks any signicant effect of the support electronegativity on
with monometallic Pt catalysts are 57, 35, 40, and 47 kJ mol1 for the metal particles.
Pt/SiO2, Pt/ASA(5/95), Pt/ASA(20/80), and Pt/ASA(55/45) [19]. The important role of the support electronegativity in the
Clearly, the Ea,Exp values are in the same range; the trend, however, hydrogenation of tetralin is conrmed by the TOF values (support-
is different. While in the case of monometallic particles increasing ing information). The intrinsic activity of PtPd/ASA(5/95) is quite
concentrations of SiO2 in the mixed oxide leads to an increasing similar to that of PtPd/ASA(55/45) in the absence of poisons (see
activity of the platinum particles (the exception is the catalyst sup- Fig. S-1 in the supporting information). Therefore, the different Pt
ported on SiO2 as explained below), the more complex variation of coverage for PtPd/ASA(55/45) and PtPd/ASA(5/95) (16.9% and
the properties for the PtPd catalysts used in this work suggests 14% respectively [26]) does not exert an important inuence in
more than one inuencing factor. The differences in the activation the intrinsic activity of the metal clusters in the absence of poisons.
energies must be determined on the one side by the varying con- Fig. 11 shows the stepwise hydrogenation of adsorbed tetralin
centrations of Pd, which binds the aromatic molecules relatively to cis- and trans-decalin. The adsorption of tetralin on Pt is stronger
weakly, and the increasing strength of bonding by Pt (being inu- leading to high cis-decalin selectivity, whereas the weaker adsorp-
enced through alloying and through the support). tion of tetralin on Pd causes the trans-decalin isomer to be pre-
PtPd/ASA(55/45) was the most active bimetallic catalyst, ferred [41]. Weak interaction of the intermediates allows the
which is attributed to the high fraction of platinum on its surface. desorption and re-adsorption of D1,9octalin and formation of
This is in line with the fact that the hydrogenation activity of Pt is trans-decalin [42]. An alternative interpretation is the rollover of
higher than that of Pd in the absence of poisons [40,41]. Accord- the intermediate without desorption possible if weak interactions
ingly, the experimental activation energy increases from are involved [43]. In the poison-free experiments, the cis-/trans-
36 kJ mol1 for PtPd/ASA(55/45) to 46 and 48 kJ mol1 for Pt decalin ratio on all the bimetallic catalysts was below 0.4, as ex-
Pd/ASA(20/80) and PtPd/ASA(5/95), respectively, that is, by pected for a catalyst with abundant surface Pd. One would argue
around 25%. This proportion corresponds to the decrease in Pt cov- that because the concentration of surface Pt is the factor determin-
erage from 16.9% to 13.9% comparing PtPd/ASA(55/45) with the ing the rate of tetralin hydrogenation, the cis-/trans-decalin ratio
other bimetallic catalysts supported on ASA (28% of decrease). has to be close to that of pure Pt catalysts. However, the interme-
Hence, accessible Pt surface atoms are the most important feature diate D9,10-octalin formed on Pt may isomerize promptly to D1,9-
of the bimetallic catalysts for reaching high hydrogenation activity. octalin on a neighbor Pd site thus enhancing the formation of
The relation between Ea,Exp and the true activation energy (Ea) is trans-decalin.
given by the Temkin Eq. (1) where ni is the reaction order in the The cis- to trans-decalin ratio decreases steadily with increasing
reactant i and DHi is the corresponding enthalpy of adsorption. temperature on PtPd/ASA(55/45) (with relatively high proportion
In agreement with ref. [19], we assume that the variation of Ea,Exp
is related to changes in the heat of adsorption of tetralin.

Ea Ea;Exp  Rni DHi 1

As the reaction is rst order in tetralin, the lowest Ea,Exp for the Pt
Pd/ASA(55/45) catalyst suggests the strongest adsorption of tetra-
lin, that is, for the catalyst with the highest fraction of accessible
Pt. Moreover, the Ea,Exp for the catalysts supported on ASA(20/80)
and ASA(5/95) are all lower than those of the SiO2-supported cata-
lyst, suggesting that the presence of Lewis and Brnsted acid sites
induces an important effect on the adsorption strength of tetralin
[19].
The activity of the catalysts with low concentration of surface Pt
in poison-free feed increases in the sequence PtPd/SiO2 < PtPd/
ASA(20/80) < PtPd/ASA(5/95). The same dependence of the activ-
ity on the composition of the support was observed in the hydro-
genation of tetralin on ASA-supported monometallic Pt catalysts
[19]. Given that the proportion of surface Pt in those three bimetal-
lic catalysts is rather similar (around 14%), it is proposed that the Fig. 11. Stepwise hydrogenation of adsorbed tetralin to cis- and trans-decalin.
22 Y. Yu et al. / Journal of Catalysis 292 (2012) 1325

of surface Pt), whereas on the rest of the catalysts, the ratio sons and the presence of quinoline (except for the catalyst
decreased only to a very small degree and remained constant at supported on silica). The high conversion of DBT suggests that
higher temperatures (Fig. 3b). Note that conceptually the cis-/ DBT and products of its surface reactions adsorb on the stronger
trans-ratio should decrease with temperature, as higher tempera- adsorbing metal sites (Pt for instance) permitting binding of tetra-
tures should lead to looser bound substrates, intermediates and lin only on relatively weaker sites. Considering that the OH
transition states. The observations indicate, however, that only in groups of the support must have the weakest adsorption strength,
the presence of a higher concentration of Pt high temperature sen- it is highly probable that tetralin interacts with the OH groups at
sitivity exists. It is unclear at the present why the selectivity does the perimeter of the metal cluster during hydrogenation. With
not depend on the temperature at high reaction temperature (see the metal surface interacting mostly with DBT, the adsorption of
the equal values of all catalysts in Fig 3b). However, overall the tetralin is limited to the Brnsted sites; hence, the support acidity
observation is in agreement with the literature [41], reporting that becomes the determining factor for tetralin hydrogenation. We
on a monometallic Pd catalyst, the cis- to trans-decalin ratio is con- cannot rigorously exclude that the strong adsorption of DBT on
stant in a wide conversion range, whereas on a comparable Pt cat- the metal surface drastically reduces the adsorption enthalpy of
alyst the ratio decreases steadily with conversion. tetralin, thus increasing the Ea,Exp that leads to the low reaction
rates of tetralin hydrogenation.
4.3. Hydrogenation of tetralin in the presence of poisons In the simultaneous presence of DBT and quinoline, the hydro-
genation activity is strongly reduced. The stronger adsorption of
The low cis- to trans-decalin ratio conrms the dilution of Pt in poisons on metal (DBT and reaction products of DBT) and acid sites
Pd because a value higher than 1 is expected for tetralin hydroge- (quinoline and reaction products of quinoline) hinders the hydro-
nation on Pt. Also the increase in Ea,Exp seems to be related to a genation of tetralin. Indeed, in all experiments the conversion rates
decreasing proportion of surface Pt. Hence, at this point, the results of the DBT and quinoline are much higher than that of tetralin. The
imply that in the absence of poisons the bimetallic catalysts be- trends of the Ea,Exp in the presence of sulfur (with and without
have as a Pt catalyst diluted with Pd. In contrast, at reaction condi- quinoline) do not correlate with the parameters that are signicant
tions inducing poisoning (DBT and quinoline present in the feed), in other conditions, that is, particle size, electronegativity of the
the activity depends not only on the morphology of the bimetallic support, or proportion of surface Pt.
clusters and their interaction with the support, but also on the size While in the presence of poisons the overall value for cis-/trans-
of the metal particles, the concentration of acid sites, and the feed decalin ratio decreases with temperature as to be conceptually
composition. expected, neither the overall values nor the changes induced by tem-
As in the case of the poison-free feed, PtPd/ASA(55/45) shows peratures strictly reect the surface concentration in Pt. Thus, the
the highest activity in the presence of DBT, quinoline and in the question arises as to what extent the variations of the cis-/trans-dec-
presence of both poisons. This conrms that the proportion of sur- alin ratio is inuencing the surface concentration and to what extent
face Pt is the parameter with the largest inuence on the hydroge- this concentration changes as a function of the temperature. Unfor-
nation activity. Let us focus now on the effect of quinoline and DBT tunately, these questions cannot be addressed with the already quite
on the activity of the other catalysts. In the presence of quinoline extensive physicochemical data reported in Part I of this study and
(Fig. 4), the activity follows the trend PtPd/ASA(20/80) < PtPd/ are subject to further studies with model catalysts.
ASA(5/95) < PtPd/SiO2, which correlates with the dispersion of The specic effects of the poisons are consistent with the model
the metal particles. The EXAFS and IR characterization of the bime- presented in ref. [20], suggesting that the adsorption of tetralin oc-
tallic catalysts supported on ASA(20/80), ASA(5/95) and SiO2 curs on two different sites, that is, on metal particles and on
showed that the proportion of metal atoms on the surface is simi- Brnsted and Lewis acid sites at the perimeter of the metal parti-
lar among these catalysts. Thus, the activity increases with the de- cles. Quinoline strongly adsorbs on the acid sites and therefore re-
crease in particle size, that is, increasing dispersion of the metal. tards the hydrogenation pathway that involves adsorption at the
Special attention must be paid in the case of the catalysts PtPd/ perimeter of the metal particles. The metal-catalyzed hydrogena-
ASA(20/80) and PtPd/ASA(5/95). The activity of the former is only tion route, however, remains active for the hydrogenation of tetr-
slightly higher than that of the latter because the particle size is alin, and consequently, the fraction of accessible metal atoms is
also slightly larger in PtPd/ASA(20/80). Given that both catalysts the decisive factor for the activity under these conditions. When
have almost identical proportion of surface Pt, their intrinsic activ- DBT is added to the feed, the adsorption of substrate on the metal
ities are nearly the same as shown in Fig. S-2 in the supporting surface is blocked by strongly adsorbed sulfur-containing mole-
information. cules or sulfur on the metal (illustrated in Reaction (1)) and the
The Ea,Exp values observed in the presence of quinoline are much activity is strongly reduced. The presence of DBT has a weaker
higher than those calculated in the series of reactions using clean inuence than that of quinoline on the pathway involving acid
feed. This is tentatively attributed to the competitive adsorption sites. Thus, in the presence of DBT tetralin adsorbs on Brnsted acid
of quinoline (molecule of basic character and more electron rich sites at the perimeter of the metal cluster and reacts with hydro-
than tetralin) decreasing the strength of adsorption of tetralin on gen, which has been dissociated on the S-covered metal surface.
the metal clusters [44]. The Ea,Exp decreases in the same order as Therefore, in the presence of quinoline the particle size is the
in the absence of poisons, that is, PtPd/SiO2 > PtPd/ASA(20/ determining factor for the activity, whereas in the presence of
80) > PtPd/ASA(5/95) > PtPd/ASA(55/45), suggesting that the DBT, the acidity determines the rate of hydrogenation. The simul-
hydrogenation activity depends largely on the adsorption of tetra- taneous addition of 20 ppm nitrogen and 100 ppm sulfur as quin-
lin on the metal phase (in turn inuenced by the metal surface pro- oline and DBT, respectively, to the feed results in very low
portion and electronegativity of the support). activity, because on both the metal particle and the Brnsted acid
In the presence of DBT, the activity of the catalysts with low sites strongly adsorbing poisons compete with the substrate for
proportion of surface Pt follows a different trend than in the pres- adsorption sites. Thus, we conclude that the reaction pathway
ence of quinoline or of poison-free feed. The hydrogenation activity involving the Brnsted acid sites and metal particles is valid also
increases in the order PtPd/SiO2 < PtPd/ASA(5/95) < PtPd/ for PtPd-based catalysts.
ASA(20/80), which is in line with the increasing concentration of With respect to the desulfurization of DBT, we would like to
Brnsted acid sites on the catalysts. The measured activation en- note that direct desulfurization dominates, that is, biphenyl and
ergy in the presence of sulfur is higher than in the absence of poi- sulfur adsorbed on the metal as shown in Reaction (1) are formed
Y. Yu et al. / Journal of Catalysis 292 (2012) 1325 23

100 It has also been shown that the adsorption strength of tetralin is
weaker on PtPd/ASA than on Pt/ASA (deduced from the low cis-/
trans-decalin ratio on PtPd/ASA). Thus, the relatively low activity
DBT conversion [%]

of Pd overcompensates for the benecial inuence of the electronic


80 interaction of Pt and Pd evidenced by the XANES analysis in [26]
hindering any improvement of the hydrogenation activity. In the
presence of DBT, the PtPd catalysts still showed lower activities
Pt-Pd/ASA(55/45) compared to the corresponding Pt samples except for the notable
60
Pt-Pd/ASA(20/80) case of the PtPd/ASA(55/45) (see Fig. 14). The low activity of most
Pt-Pd/ASA(5/95) of the bimetallic catalysts here studied contradicts the reports of
Pt-Pd/SiO2 high activity of bimetallic PtPd catalyst in the presence of poisons
40 [38,39]. The low activity of the bimetallic catalysts investigated
533 553 573 593 here is attributed to the low concentration of Pt on the surface.
Temperature [K] The benecial effect of alloying with regard to sulfur tolerance
accompanied by lower activity of Pd is not completely compen-
Fig. 12. DBT conversion during the hydrogenation of tetralin in the presence of DBT
sated in the case of the PtPd/ASA(55/45) catalyst that results from
(0.058 wt.%) and quinoline (0.019 wt.%) in the temperature range 533593 K over
PtPd catalysts (20 wt.% tetralin, 50 bar, 24-h TOS, WHSV = 32 h1). the higher proportion of surface Pt in the bimetallic clusters of this
catalyst compared to the other bimetallic catalysts.
in the most abundant reaction [45,46]. Subsequently, sulfur is re- The activity of PtPd/ASA(55/45) in the presence of sulfur is
moved by H2 yielding H2S and a free metal atom, via reaction (3) higher than the activity of Pt/ASA(55/45). Considering that the con-
[47]. Note that reactions (1) and (3) only illustrate the overall sul- centration of exposed Pt in the bimetallic catalyst must be lower
fur poisoning and surface regeneration of the metal and do not in- than in the monometallic catalyst, the higher activity must be
tend to be mechanistically correct. attributed to an enhanced activity of Pt upon alloying with Pd. Spe-
cically, the electron transfer from Pt to Pd (exceeding the transfer
Me  C12 H8 S H2 ! Me  S C12 H10 2
to the support driven by electronegativity) reduces the strength of
the metalsulfur bond thereby accelerating the removal of surface
Me  S H2 H2 S Me 3
sulfur via reaction (3) and shifting the equilibrium toward a more
The increase in activity with temperature is more evident in the accessible metal surface [48]. While the electronic effect of alloying
presence of DBT, suggesting that the extent of sulfur poisoning is similar in all bimetallic catalyst, the proportion of surface Pt in
strongly depends on the temperature. This is to be related to the PtPd/ASA(55/45) must be signicant to allow the benecial effect
decreasing concentration of adsorbed DBT with temperature (see of alloying to overturn the negative effect of Pt dilution with Pd.
the plot of DBT conversion versus temperature in Fig. 12) as well This discussion does not imply that Pd lacks catalytic activity itself.
as to the overall lower concentration of sulfur. At 573 K or above, However, for the catalysts target of this study, the activity of Pd
the DBT conversion is well above 80% with all the catalysts indicat- must be signicantly lower than that of Pt, because Pd enrichment
ing that the further increase would not increase the rate of sulfur in the surface leads to decreasing activities under all conditions.
formation on the metal surface, while the rate of sulfur removal Therefore, we rule out that Pd atoms are the reason for the high
will further increase. The net result is that the sulfur coverage of activity of PtPd/ASA(55/45) in the presence of DBT.
the surface decreases with temperature. Direct evidence of the enhanced sulfur resistance of Pt alloyed
with Pd is shown in Fig. 7. The activity of the Pt/ASA(20/80) cata-
4.4. Comparison of monometallic Pt and bimetallic PtPd catalysts lyst decreases by 55%, whereas the activity of PtPd/ASA(20/80) re-
mains almost unaffected after increasing the sulfur content from
The efciency of bimetallic and monometallic Pt catalysts has to 100 to 500 ppm. Let us emphasize that the main effect of alloying
be compared to demonstrate the importance of optimizing the Pt seems to be the weakening of the sulfurmetal bond so that sulfur
surface concentration. Hence, control experiments were performed is easily removed from the metal sites or has a less favorable equi-
on the Pt catalysts studied in Ref. [20]. All bimetallic PtPd cata- librium as shown in Reaction (3). Moreover, some DFT calculations
lysts showed a signicantly reduced activity compared to their suggest that the enhanced sulfur tolerance of PdPt alloys is re-
monometallic Pt counterparts in the absence of poisons (Fig. 13). lated to an enhanced H2 adsorption compared to that of sulfur

800
Pt-Pd/ASA(55/45)
- ] 103

Pt-Pd/ASA(20/80)
600
Activity [moltetralinh-1gCat.-1

Pt-Pd/ASA(5/95)
Pt-Pd/SiO2

400

Pt/ASA(55/45)
200 Pt/ASA(20/80)
Pt/ASA(5/95)
Pt/SiO2
0
533 553 573 593 533 553 573 593
Temperature [K] Temperature [K]

Fig. 13. Activities of Pt and PtPd catalysts at steady state for the hydrogenation of tetralin in the temperature range 533593 K (20 wt.% tetralin, 50 bar, 24-h TOS,
WHSV = 539 h1).
24 Y. Yu et al. / Journal of Catalysis 292 (2012) 1325

100
Activity [moltetralinh-1gCat.--1] 103 Pt/ASA(55/45) Pt-Pd/ASA(55/45)
-
Pt/ASA(20/80) Pt-Pd/ASA(20/80)
-
75 Pt-Pd/ASA(5/95)
-
Pt/ASA(5/95)
Pt/SiO2 Pt-Pd/SiO
- 2

50

25

0
533 553 573 593 533 553 573 593
Temperature [K] Temperature [K]

Fig. 14. Activities of Pt and PtPd catalysts at steady state for the hydrogenation of tetralin in the presence of DBT (0.058 wt.%) in the temperature range 533593 K (20 wt.%
tetralin, 50 bar, 24-h TOS, WHSV = 77 h1).

compounds [10]. Thus, alloying increases the activation and the of Pt must be maximized in order to compensate for the dilution
transfer rate of hydrogen to tetralin on the acid sites at the perim- effect (Pt in Pd) that tends to decrease the hydrogenation activity.
eter of the metal clusters.
As a nal remark, we would like to emphasize that better PtPd Acknowledgments
catalysts, that is, ones with a higher surface Pt concentration at a
xed Pt:Pd ratio, should exist and might be obtainable via different The authors are grateful to the staff of the beamline X1 at Hasy-
preparation procedures or tuning the properties of the support. The lab DESY, Hamburg, Germany for their kind help and continuous
study of the role of preparation steps on the properties of PtPd support during the experiments. Helpful discussions and editing
catalysts is going to be reported in the near future and will contrib- of the manuscript by Prof. Breitkopf, TU Dresden, is thankfully
ute to the explanation of the differences in the reported properties acknowledged.
of PdPt catalysts for hydrotreating.
Appendix A. Supplementary material
5. Conclusions
Supplementary data associated with this article can be found, in
The catalytic hydrogenation of tetralin in the absence and pres- the online version, at http://dx.doi.org/10.1016/j.jcat.2012.03.018.
ence of quinoline and dibenzothiophene was studied on bimetallic
PtPd (0.8 wt.% metal content, molar ratio Pd/Pt = 3) catalysts sup- References
ported on silica and amorphous silicaalumina (ASA). A complex
interaction between the properties of the metal clusters and those [1] G. Perot, Catal. Today 10 (1991) 447.
[2] T. Isoda, S. Nagao, X. Ma, Y. Korai, I. Michida, Appl. Catal. A 150 (1997) 1.
of SiO2 or ASA (Al2O3/SiO2 wt.% ratio of 55/45, 20/80, and 5/95) [3] M. Egorova, R. Prins, J. Catal. 224 (2004) 278.
determined the performance of the catalysts. The proportion of [4] A. Stanislaus, B.H. Cooper, Catal. Rev. Sci. Eng. 36 (1994) 75.
Pt on the surface was the most important parameter, because the [5] B.H. Cooper, B.B.L. Donnis, Appl. Catal. A 137 (1996) 203.
[6] H. Yosuda, Y. Yoshimura, Catal. Lett. 46 (1997) 43.
PtPd catalyst supported on ASA(55/45) with the highest propor- [7] H. Yasuda, T. Sato, Y. Yoshimura, Am. Chem. Soc. Div. Petrol. Chem. Prepr. 42
tion of surface Pt (16.9%) was the most active bimetallic formula- (1997) 580.
tion under all tested conditions. In the absence of poisons, the [8] H. Yasuda, T. Sato, Y. Yoshimura, Catal. Today 50 (1999) 63.
[9] T.B. Lin, C.A. Jan, J.R. Chang, Ind. Eng. Chem. Res. 34 (1995) 4284.
electronegativity of the support correlated with the hydrogenation
[10] H. Jiang, H. Yang, R. Hawkins, Z. Ring, Catal. Today 125 (2007) 282.
activity provided that it exhibits Lewis acidity. In the presence of [11] S.D. Lin, M.A. Vannice, J. Catal. 143 (1993) 563.
quinoline, the activity of the catalysts increased with the disper- [12] S.D. Lin, M.A. Vannice, J. Catal. 143 (1993) 539.
sion of the metal particles, whereas in the presence of dibenzothio- [13] D. Poondi, M.A. Vannice, J. Catal. 161 (1996) 742.
[14] V. Fouche, P. Magnoux, M. Guisnet, Appl. Catal. 58 (1990) 189.
phene, the acidity of the support compensated for the sulfur [15] O. Cairon, K. Thomas, A. Chambellan, T. Chevreau, Appl. Catal. A 238 (2003)
poisoning of the metal phase. 167.
The observed effects of the poisons indicated the presence of [16] T. Fujikawa, K. Idei, T. Ebihara, H. Mizuguchi, K. Usui, Appl. Catal. A 192 (2000)
253.
two kinds of adsorption sites for binding the reacting substrates, [17] R.M. Navarro, B. Pawelec, J.M. Trejo, R. Mariscal, J.L.G. Fierro, J. Catal. 189
that is metal surface, and Brnsted acid sites at the perimeter of (2000) 184.
the metal particles. The key contribution of acid sites in the hydro- [18] M.F. Williams, B. Fonf, C. Sievers, A. Abraham, J.A. van Bokhoven, A. Jentys,
J.A.R. van Veen, J.A. Lercher, J. Catal. 251 (2007) 485.
genation reaction was conrmed by exchanging the Brnsted acid [19] M.F. Williams, B. Fonf, C. Woltz, A. Jentys, J.A.R. van Veen, J.A. Lercher, J. Catal.
sites of the support by Cs+. 251 (2007) 497.
For comparison purposes, ASA-supported monometallic Pt cat- [20] M.F. Williams, B. Fonf, A. Jentys, C. Breitkopf, J.A.R. van Veen, J.A. Lercher, J.
Phys. Chem. C 114 (2010) 14532.
alysts (0.8 wt.%) were tested under the same reaction conditions. [21] M. Jacquin, D.J. Jones, J. Roziere, A.J. Lopez, E. Rodriguez-Castellon, J.M.T.
All the Pt catalysts were more active than the PtPd counterparts Menayo, M. Lenarda, L. Storaro, A. Vaccari, S. Albertazzi, J. Catal. 228 (2004)
in poison-free feed. In the presence of poisons, the Pt catalysts 447.
[22] Y. Yoshimura, M. Toba, T. Matsui, M. Harada, Y. Ichihashi, K.K. Bando, H.
were also more active than the bimetallic formulations with the
Yasuda, H. Ishihara, Y. Morita, T. Kameoka, Appl. Catal. A 322 (2007) 152.
remarkable exception of PtPd/ASA(55/45). The bimetallic cata- [23] S. Albertazzi, G. Busca, E. Finocchio, R. Glockler, A. Vaccari, J. Catal. 223 (2004)
lysts, however, are more resistant to poisons. It is concluded that 372.
the poison resistance of bimetallic catalysts originates from the [24] S. Albertazzi, E. Rodriguez-Castellon, M. Livi, A. Jimenez-Lopez, A. Vaccari, J.
Catal. 228 (2004) 218.
electron transfer from Pt to Pd that yields weak adsorption of poi- [25] B. Pawelec, R. Mariscal, R.M. Navarro, S. van Bokhorst, S. Rojas, J.L.G. Fierro,
sons on electron decient Pt atoms. However, the surface coverage Appl. Catal. A 225 (2002) 223.
Y. Yu et al. / Journal of Catalysis 292 (2012) 1325 25

[26] Y. Yu, B. Fonf, A. Jentys, G.L. Haller, J.A. R van Veen, O.Y. Gutirrez, J.A. Lercher, [39] S. Jongpatiwat, Z.R. Li, D.E. Resasco, W.E. Alvarez, E.L. Sughrue, G.W. Dodwell,
J. Catal. 292 (2012) 1. Appl. Catal. A 262 (2004) 241.
[27] C.C. Costa Augusto, J.L. Zotin, A. Da Costa Faro, Catal. Lett. 75 (2001) 37. [40] H. Greeneld, N.Y. Ann, Acad. Sci. 214 (1973) 233.
[28] T. Fujikawa, K. Idei, K. Ohki, H. Mizuguchi, K. Usui, Appl. Catal. A 205 (2001) 71. [41] S. Dokjampa, T. Rirksomboon, S. Osuwan, S. Jongpatiwut, D.E. Resasco, Catal.
[29] A. Infantes-Molina, J. Merida-Robles, E. Rodriguez-Castellon, J.L.G. Fierro, A. Today 123 (2007) 218.
Jimenez-Lopez, Appl. Catal. B 73 (2007) 180. [42] A.W. Weitkamp, Adv. Catal. 18 (1968) 1.
[30] Y.T. Chua, P.C. Stair, J. Catal. 213 (2003) 39. [43] K. Schrage, R.L. Burwell, J. Am. Chem. Soc. 88 (1966) 4555.
[31] G.M. do Nascimento, T.B. Silva, P. Corio, M.S. Dresselhaus, J. Raman Spectrosc. [44] J.J. Rooney, J. Mol. Catal. 31 (1985) 147.
41 (2010) 1587. [45] V.G. Baldovino-Medrano, S. Giraldo, A. Centeno, J. Mol. Catal. A 301 (2009) 127.
[32] C. Song, A.D. Schmitz, Energy Fuels 11 (1997) 656. [46] V.G. Baldovino-Medrano, P. Eloy, E.M. Gaigneaux, S.A. Giraldo, A. Centeno,
[33] J. Chupin, N.S. Gnep, S. Lacombe, M. Guisnet, Appl. Catal. A: Gen. 206 (2001) 43. Catal. Today 150 (2010) 186.
[34] T.T. Phuong, J. Massardier, P. Gallezot, J. Catal. 102 (1986) 456. [47] M. Guenin, M. Breysse, R. Frety, K. Tifouti, P. Marecot, J. Barbier, J. Catal. 105
[35] J.M. Orozco, G. Webb, Appl. Catal. 6 (1983) 67. (1987) 144.
[36] S.D. Lin, C. Song, Catal. Today 31 (1996) 93104. [48] T. Matsui, M. Harada, K.K. Bando, M. Toba, Y. Yoshimura, Appl. Catal. A 290
[37] K. Thomas, C. Binet, T. Chevreau, D. Cornet, J.P. Gilson, J. Catal. 212 (2002) 63. (2005) 73.
[38] A.D. Schmitz, G. Bowers, C. Song, Catal. Today 31 (1996) 45.

You might also like