Você está na página 1de 14

Applied Catalysis A: General 242 (2003) 1730

Aromatics hydrogenation on silicaalumina


supported palladiumnickel catalysts
V.L. Barrio a, , P.L. Arias a , J.F. Cambra a , M.B. Gemez a,
B. Pawelec b , J.L.G. Fierro b,1
a

Department of Chemical and Environmental Engineering, School of Engineers, University of the Basque Country,
Alameda Urquijo s/n, 48013 Bilbao, Spain
b Institute of Catalysis and Petrochemistry, Spanish Council for Scientific Research (CSIC),
Campus de la UAM, Cantoblanco 28049 Madrid, Spain
Received 8 June 2002; received in revised form 29 August 2002; accepted 2 September 2002

Abstract
Nickel- and palladiumnickel catalysts supported on silicaalumina have been studied in the simultaneous hydrogenation of naphthalene and toluene. These catalysts were characterised by means of X-ray diffraction, CO chemisorption,
temperature-programmed reduction, NH3 temperature-programmed desorption, diffuse reflectance spectroscopy, Fourier
transform infrared spectroscopy of adsorbed CO and X-ray photoelectron spectroscopy techniques. All the catalysts studied
showed higher initial intrinsic activity in the hydrogenation of toluene than of naphthalene. Regarding the hydrogenation of
toluene, the 1Pd8Ni/SA sample displayed the strongest resistance to deactivation by coke precursors as compared with the
Ni-free 1Pd/SA catalyst. For the 1Pd8Ni catalyst, the characterisation data pointed not only to a high degree of reducibility
of nickel but also the greatest exposure of Pd species. Both findings appear to be related to the development of nickel hydrosilicate species at the support interface. Indeed, a better resistance towards deactivation was obtained by Pd incorporation
and by increasing Ni-loading.
2002 Elsevier Science B.V. All rights reserved.
Keywords: Aromatics hydrogenation; Nickel and palladium catalysts; Silicaalumina support

1. Introduction
Owing to environmental and clean-fuel legislation
[13], deep hydrodearomatisation of diesel fuel has
been the focus of many recent studies. Supported
noble metal catalysts are well known for their high
hydrogenation activity for deep hydrodearomatisation
Corresponding author. Tel.: +34-94-6017282;
fax: +34-94-6014179.
E-mail addresses: iapbacav@bi.ehu.es (V.L. Barrio),
jlgfierro@icp.csic.es (J.L.G. Fierro).
1 Tel.: +34-91-5854769; fax: +34-91-5854760.

at low reaction temperatures and moderate hydrogen pressures, although in general they show low
resistance to sulphur poisoning [4]. Due to the low
cost and acceptable resistance to sulphur poisoning,
Ni-based systems are good alternatives to noble metal
hydrogenation catalysts [5,6]. These catalysts are particularly suited to the production of middle distillates
through the hydroconversion of hydrocarbons derived
through the FischerTropsch process.
Most studies on Ni-based catalysts have been carried out using silica or alumina supports. The use of
silicaalumina as a support has received less attention. Recently, Guimon et al. [5] reported that the

0926-860X/02/$ see front matter 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 6 - 8 6 0 X ( 0 2 ) 0 0 4 8 9 - 1

18

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

addition of Al2 O3 to SiO2 has a positive effect on the


decrease in the coke formation. This was explained in
terms of the different interaction between the metal
and the support, as shown by X-ray photoelectron
spectroscopy and temperature-programmed reduction
studies. Among Ni-based catalysts, the PdNi formulation seems to be of special interest. One important
aspect of PdNi systems is the very mild deactivation
they undergo along the successive reaction experiments [810]. The use of palladium-promoted amorphous silicaalumina supported nickel catalysts for
the hydroisomerisation of alkanes has been patented
[7]. The lower deactivation of silica-supported PdNi
alloys as compared to unsupported ones has been reported for 1,3-butene hydrogenation [8,9]. In the case
of this latter reaction, a strong decrease in deactivation on PdNi/SiO2 catalysts was explained in terms
of a preferential migration of Pd atoms on low coordination sites occupied by Ni atoms [8]. Additionally,
Hermann et al. [10] correlated the increased activity
of the Pd layer on Ni with a 0.8 eV upward shift of Pd
3d core-levels. By contrast, limited modifications in
electronic structure, which have no influence on reactivity, were observed for the PdNi/SiO2 catalysts [8].
In keeping with the foregoing, it seemed to be of
interest to study the catalytic behaviour of bimetallic
PdNi supported on silicaalumina. Accordingly, in
this paper, nickel- and palladium-loaded nickel catalysts were studied in the simultaneous hydrogenation
of naphthalene and toluene. The catalysts in their oxidic, reduced and used forms were characterised by
different techniques with a view to explaining the relationship between activity and catalytic properties.

2. Experimental procedure
2.1. Catalyst preparation
A commercially available support, SiO2 Al2 O3
(SMR 5-473, Grace Davison Chemical, hereafter referred to as SA) containing 28 wt.% alumina was
calcined in air at 773 K for 3 h prior to catalyst preparation. Ni/SA and Pd/SA catalysts were prepared by
wet impregnation of the support with aqueous solutions of Ni(NO3 )2 4H2 O (Carlo Erba, 99%) and
Pd(NO3 )2 2H2 O (Fluka, >98%), respectively. Once
adsorption equilibrium had been reached, excess
water was removed in a rotary evaporator. Then,
the impregnates were dried at 383 K in air for 12 h
and calcined in air at 723 K for 2 h. The bimetallic
PdNi/SA catalysts were prepared by impregnation
of calcined Ni/SA catalysts with the Pd(NO3 )2 2H2 O
(Fluka, >98%) solution, followed by drying and calcination as above. The metal contents of the catalysts,
as determined by plasma analysis, are reported in
Table 1. For all the catalysts, the nominal palladium
composition was 1 wt.%. Monometallic and bimetallic systems are referred to xNi and 1PdxNi, respectively, where x denotes the nominal nickel percentage.
It can be observed in Table 1, that x can be 4 or 8,
and implies a Ni content of 4 or 8 wt.%.
2.2. Catalyst characterisation
The chemical compositions of the catalysts were determined by inductively coupled plasma atomic emission spectroscopy (ICP-AES) using a Perkin-Elmer

Table 1
Chemical compositiona , textural propertiesb and metal dispersionc of Pd/SA, Ni/SA and NiPd/SA catalysts
Catalyst

SA
1Pd
4Ni
8Ni
1Pd4Ni
1Pd8Ni
a

Metal contenta
Ni (wt.%)

Pd (wt.%)

4.0
8.3
3.9
8.2

0.9

0.8
0.8

Dc (%)

BETb (m2 /g)

dp b (nm)

VN2 b (cm3 (CN)g1 )

26.6
14.6
21.7
16.1

394
372
389
375
324
326

7.5
6.9
6.6
5.5
6.9
5.7

118.4
102
107
103
89
89

As determined by chemical analysis.


BET specific area, average pore diameter (dp ) and volume of nitrogen adsorbed at P/P 0 = 0.2 as measured by N2 adsorptiondesorption
isotherms at 77 K.
c As measured by CO chemisorption.
b

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

Optima 3300DV instrument. Samples were solubilised in a mixture of HF, HCl and HNO3 at 363 K
and homogenised in a microwave oven. Diffuse reflectance UV-Vis spectroscopy was performed on
the calcined samples using a Shimadzu-UV-2100
spectrophotometer. The calcined samples were also
characterised by X-ray diffractometry according to
the step-scanning procedure (step size 0.02 ), using a
computerised Seifert 3000 diffractometer, with Cu K
( = 0.15406 nm) radiation, and a PW 2200 Bragg
Brentano /2 goniometer equipped with a bent
graphite monochromator and an automatic slit. Scanning 2 angles ranged from 10 to 70 . Line-broadening
measurements were carried out using the NiO peak
(43.35 in 2). The full-width at half-maximum
(FWHM) of these peaks was corrected for instrumental broadening (b). Diffuse reflectance spectra (DRS)
of the finely ground calcined samples were recorded
in the 240800 nm range with a Shimadzu UV-2100
spectrophotometer, using BaSO4 as a reference and
converted to the SchusterKubelkaMunk function.
The textural properties of the calcined catalysts
were evaluated from the N2 (air liquide, 99.994%)
adsorptiondesorption isotherms obtained at 77 K
over the whole range of relative pressures, using a
Micromeritics ASAP-2000 apparatus, for samples
previously outgassed at 413 K for 18 h. A value of
0.1620 nm2 was taken for the cross-section of the
physically adsorbed N2 molecule. BET specific areas
were computed from these isotherms by applying
the BET method over the 0.0050.25 P/P0 range. In
all cases, correlation coefficients above 0.999 were
obtained.
Volumetric CO chemisorption isotherms at 303 K
were obtained in order to estimate metal dispersion.
Before measurements, the sample was reduced under
H2 at 673 K for 1 h and then outgassed at 105 mbar.
In order to avoid sintering of the Pd0 particles, the
1Pd sample was reduced at 573 K. After cooling the
sample to room temperature, CO was admitted and
the adsorption isotherm was measured. The monolayer
adsorption capacity was obtained by extrapolation to
zero-pressure the linear part of the isotherm. In the
calculation of the dispersion, the stoichiometry of CO
to either Pd or Ni was assumed to be unity.
Temperature-programmed reduction profiles were
obtained on a semiautomatic Micromeritics TPD/TPR
2900 apparatus interfaced with a computer. Since PdO

19

is an easily reducible oxide, even at room temperature,


all temperature-programmed reduction experiments
were carried out after sample (0.05 g) conditioning (air
treatment at 473 K for 0.5 h) and further exposure to
the 10% H2 /Ar mixture (air liquide, 99.996% purity;
flow 60 ml/min) at temperatures from 263 to 1173 K.
Temperature was increased at a rate of 15 K/min and
the amount of H2 consumed was determined with a
thermal conductivity detector (TCD). A cooling trap
placed between the sample and the detector retained
the water formed during the reduction process.
Temperature-programmed desorption of ammonia
measurements were carried out with the same apparatus described for temperature-programmed reduction.
The sample (0.050 g) was pretreated in an He (air liquide) stream at 383 K for 1 h and then reduced in a 10%
H2 /Ar flow at 673 K for 1 h. After cooling down to
473 K, the reduced sample was ammonia-saturated in
a stream of 5% NH3 /He (air liquide) flow (50 ml/min)
for 1 h. Thereafter, after catalyst equilibration in a helium flow at 400 K for 0.5 h, the ammonia was desorbed using a linear heating rate of 10 K/min from 400
to 873 K. The water evolved was trapped in a KOH
trap placed just before the TCD. In order to determine
the total acidity of the catalyst from its NH3 desorption profile, the area under the curve was integrated.
A semiquantitative comparison of the strength distribution was achieved by Gaussian deconvolution of the
peaks. Medium and strong acidities were defined as
the areas under the peaks at the lowest and highest
desorption temperatures, respectively.
The infrared spectra of chemisorbed CO were
recorded with a Nicolet 5ZDX Fourier Transform
spectrophotometer, working with a resolution of
4 cm1 over the entire spectral range and averaged
over 100 scans. The samples, in the form of selfsupporting wafers (thickness ca. 10 mg/cm2 ), were
reduced in flowing hydrogen at 673 K for 1 h, and
then outgassed under vacuum at the same temperature for 1 h. After admission of CO at room temperature (30 mbar), the fraction of physically adsorbed
molecules was removed by outgassing at room temperature for 15 min. Net infrared spectra of chemisorbed
CO were obtained after subtraction of the background
spectrum of the solid.
The photoelectron spectra of the fresh, reduced and
used catalysts were recorded on a VG Escalab 200R
electron spectrometer equipped with a hemispherical

20

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

electron analyser, using a Mg K (h = 1253.6 eV,


1 eV = 1.603 1019 J) 120 W X-ray source. The
ECLIPSE software was used to record and analyse spectra. The reduction of calcined samples was
achieved in flowing hydrogen at 673 K for 1 h. The
samples used in the reaction were embedded in
i-octane in order to avoid exposure to air. All catalysts
were outgassed at 105 mbar, and then transferred
to the ion-pumped analysis chamber, where residual
pressure was kept below 7 109 mbar during data
acquisition. Pd 3d5/2 , Ni 2p3/2 , Si 2p and Al 2p
core-level spectra were recorded and the corresponding binding energy (BE) referenced to the C 1s line at
284.9 eV (accuracy within 0.1 eV). Peak intensities
were estimated by calculating the integral of each
peak after subtraction of an S-shaped background
and fitting the experimental peak to a combination of
Lorentzian/Gaussian lines (20% L/G).
2.3. Activity measurements
The catalysts were tested in the hydrogenation
of aromatics. The feed composition was 20 wt.%
of naphthalene and 2 wt.% of toluene dissolved in
n-hexadecane. The reaction was performed in a
bench-scale unit equipped with a stainless steel fixed
bed catalytic reactor (1.15 cm i.d. and 30 cm length).
In order to avoid hot spots, 0.5 g of catalyst was
diluted with 0.5 g of SiC (both in the 0.420.5 mm
particle size range) at a mass ratio of 1:1. Before
the reaction, the catalyst was activated in situ by reduction at atmospheric pressure under 166.6 ml/min
of H2 /N2 mixture (9:1 v/v) at 673 K for 3 h. After
catalyst activation, the reactor was cooled to the reaction temperature and the system was pressurised. The
reaction conditions were P = 5 MPa; T = 548 K;
WHSV = 0.48 h1 ; and the hydrogen-to-liquid feed
ratio was 550840 Nm3 /m3 . The reaction was kept
running until steady-state was reached. Liquid samples were collected for 1 h and analysed by GC
(HP5890) equipped with a FID detector and for product identification a mass selective detector (HPMSD
5973) and an atomic emission detector (HPG2350A)
were also used. Besides the high conversions observed for all catalysts, the reaction was free of mass
transport effects since, according to the Weisz parameter [11], the conversion was always proportional to
the mass of the catalyst or to the inverse of the flow.

3. Results
3.1. Chemical and textural properties of oxidic
catalysts
Although under hydrogenation conditions the catalysts are in reduced state, it is relevant to study their
oxidic precursors because there is a direct relationship
between the properties of calcined catalysts and the
corresponding reduced samples.
The chemical composition and textural properties of
the NiPd/SA catalysts as revealed by chemical analysis and nitrogen adsorptiondesorption isotherms, respectively, are given in Table 1. For all samples, the
shape of the N2 adsorptiondesorption isotherms (not
shown) and their hysteresis loops were similar to the
bare SA. As expected, the sorption capacity of the
SA decreased after Ni and/or Pd incorporation. The
drop in the adsorbed volumes of N2 at a relative pressure of P/P 0 = 0.2, and consequently in the BET
area and porosity, indicate that some of the micropores and/or the smaller mesopores become occupied
or blocked by nickel and/or palladium species. Noticeably, the 8Ni/SA catalyst shows a similar BET area to
that of the 1Pd/SA sample. The similar BET specific
area of the 1Pd4Ni and 1Pd8Ni catalysts shows that
the decrease in the BET specific area is mainly governed by the second impregnation step with the Pd
salt. According to the X-ray profiles (not shown), all
the calcined catalysts exhibited only one peak coming from the support (2 = 24.16), but none from the
Pd or Ni phases. Thus, it was inferred either the formation of well-dispersed amorphous PdO/NiO phases
or solid state reaction between NiO and support to
form an ill-crystallized surface compound (Ni aluminate/silicate) [12].
3.2. CO chemisorption
Data on the metal dispersion of catalysts obtained
by the chemisorption method are given in Table 1.
Metal dispersion follows the trend: 4Ni > 1Pd4Ni >
1Pd8Ni > 8Ni. As the metal dispersion of a given
metal usually increases in consecutive impregnations
[13], for 1Pd8Ni sample some re-dispersion of the
Ni species after impregnation of 8Ni catalyst with
the palladium solution could be expected. However,
for the1Pd4Ni sample, in which Ni-loading is much

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

lower and Ni species interact strongly with the support surface, re-dispersion of the nickel phase during
consecutive impregnation with the palladium salt did
not occur.
3.3. UV-Vis diffuse reflectance analysis
In order to confirm the identity of the nickel
species and the formation of nickel hydrosilicate, the
UV-Vis diffuse reflectance spectra of Ni/SA catalysts are obtained. Fig. 1 shows the UV-Vis diffuse
reflectance spectroscopy profiles for the representative 4Ni and 8Ni catalysts in the visible range. Both
catalysts showed two distinctive major bands at 370
and 650 nm, and a minor one at around 530 nm.
Since a similar band was observed for NiO obtained
by decomposition of nickel nitrate [12], the band at
370 nm was attributed to the octahedral coordinated
Ni2+ ions. Considering the diffuse reflectance spectroscopy study of Houalla and Delmon [12] on the
Ni/silicaalumina catalyst with several alumina contents, we ascribed a band at 650 nm as being due to
Ni2+ in a tetrahedral environment, and the band at
530 nm as being due to Ni3+ (the excess of oxygen

Fig. 1. UV-Vis diffuse reflectance spectra of 4Ni/SA and 8Ni/SA


catalysts.

21

in NiO generates the latter). Considering the support


composition (28% of Al2 O3 only) and calcination
temperature (723 K), it is more likely that the nickel
hydrosilicate would be formed [14], although the
formation of the nickel aluminate cannot be excluded.
3.4. Temperature-programmed reduction
The temperature-programmed reduction profiles
of the Ni/SA and PdNi/SA catalysts are shown
in Fig. 2A and B, respectively. The temperatureprogrammed reduction profile of the bare SA, also

Fig. 2. Temperature-programmed reduction profiles of Ni/SA (A)


and PdNi/SA (B) catalysts. For the sake of comparison, the
temperature-programmed reduction profiles of the Pd/SA and bare
silicaalumina support are given in figures (A) and (B), respectively.

22

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

included for comparison, exhibits two overlapping


peaks of about 947 and 1087 K, associated with the reduction of oxidic impurities. Temperature-programmed
reduction profiles of the 4Ni and 8Ni catalysts show
a broad peak with a maximum at around 800 K with
a shoulder around 900 K. Considering the UV-Vis
diffuse reflectance spectra (Fig. 1), the peak at 800 K
and its shoulder at around 900 K can tentatively be
assigned to the reduction of Ni2+ species in the octahedral and tetrahedral coordination, respectively,
the latter interacting with the support more strongly
than the former. As can be seen, the increase in
Ni content leads to an increase in the intensity of
the peaks around 800 K and 900 K. Thus, the extent of formation of a nickel hydrosilicate/aluminate
increases upon raising the Ni content, and this increase is larger for binary samples. For the 1Pd8Ni
sample, it is suggested that NiO particles placed
on the nickel hydrosilicate/aluminate phase are
dominant.
From temperature-programme reduction data, the
formation of PdNi alloy is excluded since after Pd
incorporation onto Ni catalysts no lowering in the NiO
reduction temperature did occur. This can be expected
considering that palladium has a lower surface tension
and a larger atomic radius than Ni as well as the Ni
is in large excess compared to Pd. As a consequence,
Pd tends to be expelled out of the Ni matrix leading
to a strong Pd segregation [9].
Temperature-programmed reduction profiles of the
Pd-containing catalysts show an H2 -consumption
peak at 327 K and a negative one at 355 K. The former comes from the reduction of the Pd oxide species
[15] while the latter is usually ascribed to the desorption of hydrogen from the decomposition of a bulk
palladium hydride formed through H-diffusion into
the Pd crystallites [16]. For the binary systems both
peaks are much more intense than in the 1Pd sample,
suggesting that the structure and/or location of the
PdO phase changes in the presence of nickel. For the
1Pd4Ni catalyst, the maximum of this peak appears
at almost the same temperature, indicating a similar
particle size of PdO particles. For the 1Pd8Ni catalyst, the peak of Pd reduction is also larger than in
its 1Pd4Ni counterpart. Moreover, the shift in the
temperature maximum (350 versus 327 K) points to a
stronger interaction between Pd species and the support. In addition, the second peak of the bimetallic

catalysts (typical of NiO) shows a broader reduction


peak, which is consistent with a more heterogeneous
distribution of oxidic nickel species.
3.5. Temperature-programmed desorption of
ammonia
Ammonia temperature-programmed desorption is a
commonly used technique for the titration of surface
acid sites. It is generally accepted that ammonia is an
excellent probe molecule for testing the acidic properties of solid catalysts since its strong basicity and
small molecular size allow the detection of acidic sites
located even in very narrow pores [17]. The strength
of an acid site can be related to the corresponding
desorption temperature, while the total amount of
ammonia desorption after saturation coverage permits
quantification of the number of acid sites at the surface.
The temperature-programmed desorption profiles are
given in Fig. 3. Experimental curves were analysed by
a mathematical fitting program and deconvoluted into
Gaussian functions. The area of the components, expressed as moles of ammonia per gram of catalyst, is
given in Table 2. Thus, temperature-programmed desorption peaks at two temperature ranges of 529640
and 700900 K can be ascribed to ammonia adsorbed
on middle- and high-strength acid sites, respectively.
Middle-temperature acid sites were split into two
groups of acid sites: the first from 529 to 576 K and
the second ones from 601 to 640 K.
In comparison with the SA support, the total
acidity of the catalysts after metal incorporation
diminishes and strong acidity almost completely
disappears, due to the exchange of H+ by Ni2+
Table 2
Total acidity and distribution of the acidic sites for Pd/SA, Ni/SA
and PdNi/SA catalystsa
Catalyst

Bare SA
1Pd
4Ni
8Ni
1Pd4Ni
1Pd8Ni
a

Acid amount (mmol NH3 /gcat )


T536576 K

T601650 K

T702711 K

5.3
3.8
3.1
2.9
1.0
0.7

3.4
4.3
1.0
1.5

5.8

As measured by the temperature-programmed desorption of


ammonia.

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

23

Fig. 3. Normalised NH3 temperature-programmed desorption profiles of reduced 1Pd/SA, Ni/ASA, 1PdNi/SA catalysts and bare support.
For the sake of comparison, profiles have been deconvoluted.

24

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

during the impregnation step or to some interaction


between the metals and H+ along the calcination
step. It was noted that this decrease did not correlate
with the amount of metal incorporated. As a general
trend, the proportion of weak middle-temperature
acid sites decreases with increasing metal contents
whereas the strong middle-temperature acidity follows the opposite trend. According to these results, the
middle-temperature acidity in the 536676 K range
might well be associated to acid sites on the support
and the middle-strength acidity in the 601650 K
range could be ascribed to Brnsted-type acidity on
Ni [18]. Compared with the monometallic catalysts,
the bimetallic ones show the peak maximum of Ni2+
acidity shifted toward a higher temperature, and this
shift increases with the increase in Ni contents. This
indicates a higher strength of this type of acidic site
in the bimetallic systems.
3.6. Fourier transform infrared spectroscopy
of the CO probe
The infrared spectra of adsorbed CO on the reduced
(673 K, 1 h) 1Pd and 1Pd8Ni samples are shown
in Fig. 4. At saturation coverage, the spectra of the
1Pd sample displays an intense absorption band at
2107 cm1 and a less intense one at 1959 cm1 , corresponding to carbon monoxide bonded to the surface
of palladium atoms in linear and bridged structures,

Fig. 4. Fourier Transform infrared spectra of adsorbed CO for


1Pd/SA and 1Pd8Ni/SA catalysts.

respectively [19]. The bridged form is characteristic


of the carbonyl (Pd2 CO) species on the (1 0 0) planes
of the metal crystal [20]. In comparison with the 1Pd
sample, the IR spectra of the 1Pd8Ni catalyst show
the less intense band shifted from 2107 to 2097 cm1 .
Weakly adsorbed CO, giving a band in the region
20702090 cm1 , has been reported for Ni/Al2 O3 and
attributed to CO held by isolated Ni0 atoms or at Ni
sites adjacent to oxygen atoms or oxide support [21].
Considering TPR data, the most likely explanation is
that 1Pd8Ni catalyst contains NiO crystallites, which
are stabilized against reduction in hydrogen. Therefore, its band at 2097 cm1 may be attributed to CO
adsorbed at nickel hydrosilicate species.
For 1Pd and 1Pd8Ni samples, the most apparent
difference in region 20631700 cm1 is the general
increase in the intensity of bands assigned to bridging/
multisite adsorption. The bands at 1987 and 1880 cm1
could be indicative of the two- and three-fold, respectively, bridging sites of the CO bonded on the Ni0
[22]. Similarly, the strong band at 1842 cm1 and
much weaker band at 1731 cm1 may be assigned to
CO multisite adsorbed on Pd0 or Ni0 phases.
3.7. X-ray photoelectron spectroscopy
Photoelectron spectra of Si 2p, Al 2p, Ni 2p and Pd
3d core-levels were recorded from reduced and used
samples. The binding energies and peak percentages
are summarised in Table 3. The binding energies of
Si 2p and Al 2p were observed in all the samples
at 74.5 and 102.5 eV, respectively. For the reduced
sample, the Ni 2p3/2 profile exhibits three peaks: one
at 852.8 eV, due to metallic Ni0 species; the second
one at 855.4 eV, ascribed to unreduced Ni2+ species,
and the third oneat ca. 860 eVis the satellite of the
second one. This satellite has also been considered for
quantification [23]. Additionally, the Pd 3d5/2 profile
displays a single peak at 335.0 eV, characteristic of
Pd0 [24]. Noticeably, for used Pd-containing catalysts
this peak is shifted toward higher binding energies
(from 335.0 to 335.8 eV). By contrast, in comparison
with the 1Pd4Ni sample, the peak indicative of Ni0
species in the 1Pd8Ni sample is shifted toward lower
binding energies (from 852.7 to 851.5 eV), indicating
a higher electron density in the reduced Ni species. It
can also be noted that, with palladium incorporation,
the reducibility of the nickel species increases from

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

25

Table 3
Binding energies (eV) of core electrons and surface atomic ratios of reduced (fresh) and used Pd/SA, Ni/SA and PdNi/SA catalystsa
Catalyst

Ni 2p3/2

Pd 3d5/2

Ni/(Al + Si) atomic ratio

Pd/(Al + Si) atomic ratio

335.0
335.8

0.010
0.005

1Pd

Fresh
Used

4Ni

Fresh

852.7
855.3
852.8
855.3

(18)
(82)
(40)
(60)

0.029

852.7
855.3
852.8
855.3

(25)
(75)
(37)
(63)

0.040

852.6
855.1
852.7
855.2

(31)
(69)
(54)
(46)

335.0

0.029

0.015

335.8

0.016

0.008

852.6
855.2
851.5
854.6

(30)
(70)
(41)
(59)

335.2

0.049

0.023

335.8

0.048

0.013

Used
8Ni

Fresh
Used

1Pd4Ni

Fresh
Used

1Pd8Ni

Fresh
Used

0.032

0.039

From X-ray photoelectron spectroscopy.

1825 to 3031% (see peak percentages in parentheses


in Table 3).
For quantification, the Ni/(Al + Si) and Pd/(Si + Al)
atomic ratios are also offered in Table 3. The
Ni/(Al + Si) surface ratio did not change after
on-stream operation. The only exception is the
1Pd4Ni sample, in which this ratio decreases strongly
in the used sample. Since an intense carbon signal (C
1s peak) was observed in this sample, its Ni/(Al + Si)
ratio should be taken with caution. Thus, the nickel
phase remains virtually unchanged upon on-stream reaction. The opposite occurs with palladium. Looking
at the Pd/(Al + Si) ratios in Table 3, it is evident that
Pd undergoes some sintering in all the used samples.
3.8. Simultaneous toluene and naphthalene
hydrogenation
Simultaneous hydrogenations of toluene and naphthalene were chosen as a test reaction. For the hydrogenation of toluene and naphthalene, the intrinsic
activity at 548 K (expressed as gram of toluene or
naphthalene per gram of catalyst and per hour) plotted against the run time are shown in Figs. 5 and 6,
respectively. As expected, the intrinsic activity of

Fig. 5. Effect of the reaction time on intrinsic activity in the


hydrogenation of toluene. Reaction conditions were: T = 548 K;
P = 50 bar; WHSV = 0.48 h1 .

26

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

Fig. 6. Effect of the reaction time on intrinsic activity in the


hydrogenation of naphthalene. Reaction conditions as in Fig. 5.

catalysts decreased with time on stream. For the hydrogenation of toluene, the 8Ni/SA and 1Pd8Ni/SA
samples show higher activity as compared to the
1Pd/SA sample during the first 7 h on-stream. Considering the run time without deactivation, the
1Pd8Ni/SA sample shows an approximately two-fold
higher resistance to deactivation than its Pd-free
8Ni/SA catalyst. Similar behaviour has previously
been reported for PdNi alloys supported on silica;
these underwent less deactivation than the Pd-free Ni
catalyst [8]. According to the steady-state data shown
in Fig. 5, only the 1Pd/SA sample is active in the
hydrogenation of toluene whereas the activity of the
other catalysts resembles that of bare SA. Contrary to
toluene hydrogenation, the steady-states of the naphthalene hydrogenation activities of all the catalysts
are much more active than bare SA. The activities of
both 1Pd/SA and 1Pd8Ni/SA catalysts are highest
and are very close each other. Interestingly, unlike
the hydrogenation of toluene, these two catalysts
do not undergo deactivation in the hydrogenation of
naphthalene.

Fig. 7. Effect of the reaction time on selectivity in the hydrogenation of naphthalene (tetralin and trans-decalin formation) and toluene
conversion on the 1Pd4Ni/SA catalyst. Reaction conditions as in Fig. 5.

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

27

Fig. 8. Effect of the reaction time on selectivity in the hydrogenation of naphthalene (tetralin and trans-decalin formation) and toluene
conversion on the 1Pd8Ni/SA catalyst. Reaction conditions as in Fig. 5.

For the hydrogenation of naphthalene at 548 K, the


selectivities toward tetralin and trans-decalin formation for the most active 1Pd4Ni/SA, 1Pd8NiSA and
1Pd/SA catalysts are plotted in Figs. 79, respectively.

As seen in Fig. 7, the main product in the hydrogenation of naphthalene over the 1Pd4Ni/SA catalyst is tetralin. A similar type of behaviour is observed
for the 4Ni/SA catalyst (data not shown). Under the

Fig. 9. Effect of the reaction time on selectivity in the hydrogenation of naphthalene (tetralin and trans-decalin formation) and toluene
conversion on the 1Pd/SA catalyst. Reaction conditions as in Fig. 5.

28

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

conditions used in this work, the hydrogenation of


naphthalene to tetralin can be considered as an irreversible reaction because at T = 548 K and a hydrogen pressure of 50 bar the equilibrium concentration of
naphthalene is negligible [2]. Contrary to the 4Ni/SA
and 1Pd4Ni/SA catalysts, the main reaction product
for the 1Pd/SA, 8Ni/SA and 1Pd8Ni/SA catalysts
is decalin. In particular, the 1Pd/SA catalyst shows
the highest selectivity toward decalin formation (over
90%) (Fig. 9). For the 1Pd4Ni/SA and 1Pd8Ni/SA
catalysts, tetralin and toluene conversion are compared
in Figs. 7 and 8, respectively. The decrease in toluene
conversion in both samples is accompanied by an opposite increase in selectivity toward tetralin formation
in the hydrogenation of naphthalene. A similar type
of behaviour is seen for the other catalysts.
All the catalysts studied show a preferential
trans-decalin formation during the whole run time.
The 1Pd/SA catalyst shows the greatest selectivity
toward trans-decalin formation (about 70%) (Fig. 9).
The preferential trans-decalin formation contrasts
with the general behaviour of Group VIII metals
[25,26], but is similar to the trend observed with metal
sulphide [26]. For metal sulphides, the edgewise adsorption of the naphthalene, which allows H-addition
from opposite sites, has been proposed [26]. Such
addition is possible because hydrogen chemisorption
is heterolytic [27]. Interestingly, selectivity toward
cis-decalin formation increases with run time over
Ni/SA and PdNi/SA catalysts, but not on the 1Pd/SA
one. Since the 1Pd/SA catalyst shows the greatest
stability during on-stream operation, it seems that
coke formation on the active sites might change the
adsorption mode of naphthalene, thus influencing the
steroselectivity of decalin formation.

4. Discussion
In order to avoid the influence of deactivation and
to establish some catalyst structureactivity correlations, Fig. 10(a) and (b) compare the initial activity (calculated at zero-time conversion) expressed in
gToluene/Naphthalene /(gcat h) and the turnover frequencies (TOFs), respectively. Considering the latter, the
major feature is that, irrespectively of the reacting
molecule, the 1Pd sample showed larger TOFt=0 than
other Ni catalysts. This is because Ni has a weaker

Fig. 10. Comparison of the intrinsic activity (a) and turnover frequencies (TOFs) at zero-time conversion (s1 ) (b) for simultaneous toluene (Tol) and naphthalene (NP) hydrogenation over Pd/SA,
Ni/SA and PdNi/SA catalysts. Reaction conditions as in Fig. 5.

hydrogenation activity compared to palladium and the


synergy between Pd and Ni is hardly expected [8]. The
other important point is the larger TOFt=0 of all catalysts for naphthalene than for toluene. This is in line
with previous findings [28,29] and could be related to
a decrease in the resonance energy per aromatics ring
as well as to differences in the -electron density in the
aromatics ring as a result of the inductive effect of the
methyl group [30]. As expected, for both toluene and
naphthalene, there is no correlation between the metal
dispersion determined by CO chemisorption (Table 1)
and TOFs values. These confirm that hydrogenation
of aromatics is a structure insensitive reaction, i.e. specific activity does not depend on the details of the surface structure of the metal crystallites.

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

Contrary to TOFt=0 values, all catalysts showed


greater specific reaction rates in the hydrogenation of
toluene than naphthalene (Fig. 10(a)). For toluene, the
initial intrinsic activity ranking is: 1Pd8Ni/SA >
8Ni/SA 1Pd/SA
4Ni/SA > 1Pd4Ni/SA

SA, whereas for naphthalene the reactivity of all the


catalysts is very similar. On comparing the activities
of both the 1Pd/SA and 8Ni/SA catalysts with the binary 1Pd8Ni/SA one, the absence of a synergy effect is evident. A similar absence of a synergy effect
was observed for the silica-supported PdNi systems
tested for the hydrogenation of 1,3-butadiene [8].
Considering Ni-containing catalysts, several explanations can be advanced to explain the greater reaction
rate of the 1Pd8Ni/SA catalyst in the hydrogenation
of toluene at steady-state: (i) electronic modification
of the metal particles [10]; (ii), the absence of sintering of the metal phases; (iii) additional activity from
acid sites [3134]; and (iv), increasing resistance to
poisoning of the metal sites of the catalysts by coke
precursors [810].
Considering point (i), the IR spectra of adsorbed
CO and X-ray photoelectron spectroscopy data of
the 1Pd8Ni/SA catalyst rule out the participation of
electronic modification of Pd atoms by surrounding
Ni atoms. The limited modification of the electronic
structure of the PdNi/SiO2 catalyst, which have no
effect on the reactivity, were also observed by Renouprez et al. [8]. Moreover, the similar Ni/(Al + Si)
atomic ratio for the reduced and used 1Pd8Ni/SA
catalyst precludes its deactivation due to the sintering
of Ni particles. Since the temperature-programmed
reduction measurements point to the presence of
nickel hydrosilicate phases in all the samples, their
role could be similar to that of a NiAl2 O4 phase
[35]. This phase is considered to act as a support for
Ni phases [35], providing a stabilising effect while
maintains a very low reactivity [6,36]. Since for the
reduced and used catalysts, the Ni surface exposure
measured correlates well with hydrogenation activity,
the initial formation of nickel hydrosilicate/aluminate
could inhibit the sintering of nickel species. The only
exception is the 1Pd4Ni catalyst. This is because of
the longer period needed for its stabilisation during
on-stream operation (100 h; the rest of the catalysts
only required 24 h). As a consequence, this sample
simultaneously presents a large amount of coke on
the surface.

29

Considering additional activity from acid sites


[3,3134], the differences in activity expressed as the
initial reaction rates and in TOFt=0 suggest that both
metal and acid sites of silicaalumina may be involved in the rate-determining step of the hydrogenation of toluene [37]. Indeed, contrary to naphthalene,
the bare SA was active in hydrogenation of toluene.
However, in this study no exists correlation between
the ammonia temperature-programmed desorption
data (Table 2) and the trend in catalytic activity
(Fig. 10(a)). Moreover, the reduction temperature of
the Ni species after Pd incorporation did not change
(Fig. 2). Thus, the additional activity from hydrogen
spilled over from metal to the acid sites where toluene
can be adsorbed is hardly possible. This is due to the
formation of large amount of a nickel hydrosilicate
phase, which modifies the environment of the acid
sites to a considerable extent.
Finally, considering the greater resistance to deactivation of the 1Pd8Ni/SA catalyst, it can tentatively be
assumed that as a consequence of the stronger metallic character of Ni atoms the hydrogenation of adsorbed coke precursors may be favoured. Despite this,
the stronger metallic character of Ni0 is produced during on-stream operation, and this was not seen in the
reduced samples. Since 1Pd8Ni/SA catalyst shows
resistance to deactivation only for 7 h, and after this
run time a dramatic inhibition in activity occurs, some
changes in the structure of the catalyst after coke saturation must occur. This could be explained in terms
of the notion that a drastic decay in the conversion of
toluene would be accompanied by a parallel decrease
in selectivity toward decalin formation. Noticeably,
this decrease in hydrogenation capability was accompanied by parallel increase in the metallic character
of Ni0 . Since the X-ray photoelectron spectroscopy
data confirmed a decrease in the exposure of Pd atoms
during the reaction, in agreement with the findings of
Renouprez et al. [8], we can tentatively propose that
during on-stream reaction the Pd atoms migrate to
the sites where coke is preferentially adsorbed. Since
coke formation starts on low coordination metal sites
[8] and their surroundings and then continues on the
support [38], the strong resistance to the deactivation
would be related to Pd incorporation and nickel hydrosilicate formation, which might inhibit the growth
of coke precursors at the support interface. Since Pd
exhibits a greater hydrogenation ability than Ni, once

30

V.L. Barrio et al. / Applied Catalysis A: General 242 (2003) 1730

palladium becomes deactivated a drastic activity decay is to be expected.


5. Conclusions
The surface characteristics of SA-supported Ni
and PdNi catalysts and their catalytic performance
for aromatics hydrogenation have been studied. The
following conclusions were obtained: (i) an increase
in activity in the hydrogenation of toluene might
be attained by Pd incorporation and by increasing
Ni-loading. However, for the binary (PdNi) catalysts,
no synergy effect between both Pd and Ni metals was
observed; (ii) the greater activity of the 1Pd8Ni/SA
catalyst can be associated with its greater resistance to
deactivation due to coke formation because the larger
amount of metal phase incorporated could inhibit the
growth of coke precursors at the support interface.
As a general conclusion, the incorporation of 1% Pd
and a larger amount of Ni (8 wt.%) are beneficial for
aromatics hydrogenation at 548 K over silicaalumina
supported catalysts. These catalysts could be used
in the final stages of refinery processes, where high
quality and very low environmental impact diesel fuel
must be produced.
Acknowledgements
The University of the Basque Country, Repsol-YPF
Foundation, and grants-in-aid from CICYT (Spain)
and the EU supported this research. One of us (B.P.)
acknowledges financial support from the Spanish Ministry of Science and Technology through the Ramon
y Cajal program.
References
[1] A. Avidan, B. Klein, R. Ragsdale, Hydroc. Process 80 (2)
(2001) 47.
[2] A. Stanislaus, B.H. Cooper, Catal. Rev. Sci. Eng. 36 (1)
(1994) 75.
[3] B. Pawelec, R. Mariscal, R.M. Navarro, S. van Bokhorst, S.
Rojas, J.L.G. Fierro, Appl. Catal. A: Gen. 225 (2002) 223.
[4] C.H. Bartholomew, P.K. Agrawal, J.R. Katzer, Adv. Catal. 31
(1982) 135.
[5] C. Guimon, N. El Horr, E. Romero, A. Monzon, Stud. Surf.
Sci. Catal. 130D (2000) 3345.

[6] J.A. Pea, J. Herguido, C. Guimon, A. Monzon, J.


Santamaria, J. Catal. 159 (1996) 313.
[7] C.M. van Ballegoy, W.H.J. Stork, J.P. Gilson, Eur. Patent
Appl. EP 587245 A1, 16 Mar 1994, p. 7.
[8] A. Renouprez, J.F. Faudon, J. Massardier, J.L. Rousset, P.
Delichre, G. Bergeret, J. Catal. 170 (1997) 181.
[9] P. Miegge, J.L. Rousset, B. Tardy, J.C. Bertolini, J. Catal.
149 (1994) 404.
[10] P. Hermann, B. Tardy, D. Simon, J.M. Guignier, B. Bigot,
J.C. Bertolini, Surf. Sci. 307 (1994) 422.
[11] C.N. Satterfield, Heterogeneous Catalysis in Industrial
Practice, 2nd ed., Mc Graw-Hill, New York, 1980.
[12] M. Houalla, B. Delmon, J. Phys. Chem. 84 (1980) 2194.
[13] C.V. Cceres, J.L.G. Fierro, M.N. Blanco, H.J. Thomas, Appl.
Catal. A: Gen. 10 (1984) 333.
[14] J. van de Loosdrecht, A.M. van der Kraan, A.J. van Dillen,
J.W. Geus, J. Catal. 170 (1997) 217.
[15] T.C. Chang, J.J. Chebn, V.T.J. Yeh, J. Catal. 96 (1985) 51.
[16] M.A.Vannice, P. Chou, in: Proceedings of the 8th International
Congress on Catalysis, Tokyo, 1984, p. V-99.
[17] M.C. Kung, H.H. Kung, Catal. Rev. Sci. Eng. 27 (3) (1985)
425.
[18] S. Benbenek, E. Fedorynska, P. Winiarek, React. Kinet. Catal.
Lett. 51 (1) (1993) 189.
[19] J.K. Lee, H.K. Rhee, J. Catal. 177 (1998) 208.
[20] P. Marcot, A. Akhachane, J. Barbier, Catal. Lett. 36 (1996)
37.
[21] J.B. Peri, J. Catal. 86 (1984) 84.
[22] B. Pawelec, L. Daza, J.L.G. Fierro, J.A. Anderson, Appl.
Catal. A: Gen. 145 (1996) 307.
[23] W. Dai, M. Qiao, J. Deng, Appl. Surf. Sci. 120 (1997) 119.
[24] Y. Yazawa, H. Yoshida, N. Takagi, S. Komai, A. Satsuma, T.
Hattori, J. Catal. 187 (1999) 15.
[25] A.W. Weitkamp, Adv. Catal. 18 (1968) 1.
[26] J. Shabtai, N.K. Nag, F.E. Massoth, in: M.J. Phillips, M.
Ternan (Eds), Proceedings of the 9th International Congress
on Catalysis, Calgary, vol. 1, Chem. Institute of Canada,
Ottawa, 1988, p. 1.
[27] R. Redey, W.K. Hall, J. Catal. 119 (1989) 534.
[28] M.V. Rahaman, M.A. Vannice, J. Catal. 127 (1991) 251.
[29] R.M. Navarro, B. Pawelec, J.M. Trejo, R. Mariscal, J.L.G.
Fierro, J. Catal. 189 (2000) 184.
[30] C. Moreau, P. Geneste, in: J.B. Moffat (Ed.), Theoretical
Aspects of Heterogeneous Catalysis, Van Nostrand Reinhold,
New York, 1990, p. 256.
[31] M.V. Rahaman, M.A. Vannice, J. Catal. 127 (1991) 251.
[32] P. Chou, M.A. Vannice, J. Catal. 107 (1987) 129.
[33] S.D. Lin, M.A. Vannice, J. Catal. 143 (1993) 539, 554, 563.
[34] H. Liu, G.D. Lei, W.M.H. Sachtler, Appl. Catal. A: Gen. 137
(1996) 167.
[35] A. Al-Ubaid, E.E. Wolf, Appl. Catal. 40 (1988) 73.
[36] R. Lamber, G. Schulz-Ekloff, J. Catal. 146 (1994) 601.
[37] S. Siffert, J.-L. Schmitt, J. Sommer, F. Garin, J. Catal. 184
(1999) 19.
[38] C.L. Pieck, R.J. Verderone, E.J. Jablonski, J.M. Parera, Appl.
Catal. 55 (1989) 1.

Você também pode gostar