Sie sind auf Seite 1von 10

Applied Catalysis A: General 240 (2003) 151–160

Effect of Re loading on the structure, activity


and selectivity of Re/C catalysts in hydrodenitrogenation and
hydrodesulphurisation of gas oil
N. Escalona a , M. Yates b , P. Ávila b , A. López Agudo b ,
J.L. Garcı́a Fierro b , J. Ojeda a , F.J. Gil-Llambı́as a,∗
a Universidad de Santiago de Chile, Casilla 40, Correo 33, Santiago, Chile
b Instituto de Catálisis y Petroleoquı́mica, CSIC, Cantoblanco, 28049 Madrid, Spain
Received 26 April 2002; received in revised form 26 July 2002; accepted 27 July 2002

Abstract
A series of Re-containing catalysts supported on activated carbon, with Re loading between 0.74 and 11.44 wt.% Re2 O7 , was
prepared by wet impregnation and tested in the simultaneous hydrodesulphurisation (HDS) and hydrodenitrogenation (HDN)
of a commercial gas oil. Textural analysis, XRD, X-ray photoelectron spectroscopy (XPS) and surface acidity techniques
were used for physicochemical characterisation of the catalysts. Increase in the Re concentration resulted in a rise in the HDS
and HDN activity due to the formation of a monolayer structure of Re and the higher surface acidity. At Re concentrations
>2.47 wt.% Re2 O7 (0.076 Re atoms nm−2 ) the reduction in the catalytic activity was related to the loss in specific surface
area (BET) due to reduction in the microporosity of the carbon support. The magnitude of the catalytic effect was different
for HDS and HDN, and depended strongly on the Re content and reaction temperature. The apparent activation energies were
about 116–156 kJ mol−1 for HDS and 24–30 kJ mol−1 for HDN. This led to a marked increase in the HDN/HDS selectivity
with decreasing temperature (values >3 at 325 ◦ C), due to the large differences in the apparent activation energies of HDS and
HDN found for all catalysts. A gradual increase in the HDN/HDS selectivity with increased Re loading was also found and
related to the observed increase of catalyst acidity. The results are compared with those obtained for a series of Re/␥-Al2 O3
catalysts.
© 2002 Elsevier Science B.V. All rights reserved.
Keywords: Supported Re catalysts; Re sulphide; Hydrodenitrogenation (HDN); Hydrodesulphurisation (HDS)

1. Introduction catalysts, will be necessary to satisfy the demand


for cleaner transport fuels [1,2]. For this purpose,
In the future, a new generation of catalysts for considerable efforts have been made in recent years
deeper hydrodesulphurisation (HDS) and hydrodeni- to develop more active hydrotreating catalysts based
trogenation (HDN) and major reduction of aromatic either on new active phases or new and modified
content in diesel fuel, more efficient than the con- supports [2–4].
ventional alumina-supported Co (or Ni)-Mo (or W) Several systematic experimental studies of the cata-
lytic properties of transition metal sulphides demon-
∗ Corresponding author. Fax: +52-562-6812108. strated that unsupported or carbon-supported Re
E-mail address: fgil@lauca.usach.cl (F.J. Gil-Llambı́as). sulphide catalysts had a high activity for HDS [5–7]

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 3 0 - 1
152 N. Escalona et al. / Applied Catalysis A: General 240 (2003) 151–160

and HDN reactions [8,9]. However, relatively few de- catalysts, it has previously been reported that the thio-
tailed studies on the catalytic HDS or HDN activities phene HDS activity increased in the order SiO2 <
of supported Re-based catalysts have been reported ␥-Al2 O3 < C and the Re catalysts were 2–20 times
[10–14]. In a previous study [15], we found that the more active than molybdenum ones [10]. Thus, cata-
Re/␥-Al2 O3 system exhibited a high selectivity for lysts based on Re/C could be an interesting alternative
HDN reactions relative to HDS. This behaviour is for hydrotreating processes, especially for HDN.
unusual on conventional supported molybdenum or In the present study, the performance of a series of
tungsten sulphide catalysts, which have generally sulphided Re catalysts supported on activated carbon
higher HDS activity relative to HDN activity. Excep- was examined in the simultaneous HDS and HDN of
tional properties for HDN have been reported only a commercial gas oil. Catalysts were characterised by
with catalysts of unconventional composition, such as several physicochemical techniques.
unsupported Fe-Mo or Fe-W sulphide catalysts [16],
certain noble metal sulphides supported on active car- 2. Experimental
bon [17] and Mo-Ir/␥-Al2 O3 sulphided catalysts [18].
However, most of these catalysts gave HDN/HDS ra- 2.1. Catalyst preparation
tios generally <1, while on Re/␥-Al2 O3 catalysts the
HDN/HDS ratios were between 1.5 and 2.5 at 325 ◦ C, Re-based catalysts at various metal loadings (0.74–
depending on Re content, and <1 at higher reaction 11.44 wt.% Re2 O7 ) were prepared by wet impregna-
temperatures [15]. This exceptional HDS selectivity tion of the activated carbon support (SBET 817 m2 g−1 ,
of Re/␥-Al2 O3 catalysts is a very promising result total pore volume 0.637 cm3 g−1 , particle size
for improvement of catalysts for HDN, and deserves 20–16 mesh) with aqueous solutions of the appro-
further research on other supports more inert than alu- priate concentrations of NH4 ReO4 (Aldrich, p.a.)
mina, such as carbon. Metal sulphides supported on in a rotary evaporator. After impregnation, the sam-
carbon are known to have a higher intrinsic activity in ples were dried at 110 ◦ C for 12 h. The Re loading
HDS, HDN and hydrogenation (HYD) reactions than was expressed as atoms nm−2 of the support. The
the corresponding ␥-Al2 O3 -supported ones [6,19]. Re content was determined by inductively coupled
The high activity of carbon-supported Co-Mo cata- plasma-atomic emission spectroscopy (ICP-AES), in
lysts has been usually explained in terms of a weak a Perkin-Elmer model Optima 3300 DV spectrometer
catalyst–carbon interaction, which leads to a higher using the 221.426 nm Re emission line. The catalysts
degree of sulphidation and formation of the more ac- listed in Table 1 are denoted by the number of metal
tive Co-Mo-S type II structure [19], but the role of car- atoms nm−2 of initial support area, e.g. Re(0.076)/C
bon is still not clear. Recently, Chianelli and Pecoraro contains 0.076 atoms of Re nm−2 (2.47 wt.% Re2 O7 ).
have claimed that carbon stabilises MoS2 particles,
keeping crystallites smaller and less stacked, leading 2.2. Catalyst characterisation
to a better dispersion on a carbon support [20]. In that
study, it was also suggested that the active surface The specific surface areas, micro and mesopore vol-
in the stabilised structure of the molybdenum sul- umes and pore size distributions of the samples were
phide phase was carbided. In the case of Re sulphide determined from analysis of N2 adsorption–desorption

Table 1
Composition and physical characteristics of oxidic catalysts
Catalyst Re loading Re loading SBET Sext Total pore volume Micro pore
(wt.% Re2 O7 ) (atoms nm−2 ) (m2 g−1 ) (m2 g−1 ) (cm3 g−1 ) diameter (nm)
Re(0.00)/C – – 817 44 0.637 0.865
Re(0.024)/C 0.74 0.024 795 51 0.739 0.847
Re(0.076)/C 2.47 0.076 819 51 0.776 0.941
Re(0.135)/C 4.29 0.135 704 49 0.684 0.954
Re(0.380)/C 11.44 0.380 584 47 0.593 0.954
N. Escalona et al. / Applied Catalysis A: General 240 (2003) 151–160 153

isotherms at −196 ◦ C using a Carlo Erba 1800 Sorp- The X-ray diffraction (XRD) was carried out on a
tomatic apparatus. The samples were previously out Siemens D5000 diffractometer using the Cu K␣ radi-
gassed at 110 ◦ C to a final vacuum of <0.05 Pa. The ation (λ = 1.540598 Å) operating at 40 kV and 40 mA
BET method [21] was employed to determine the spe- and scanning 2θ angles in the range from 10 to 50◦ .
cific surface areas, taking the area of the N2 molecule The surface acidity of oxided catalysts was mea-
as 0.162 nm2 and the thickness of the monolayer as sured potentiometrically by titration with n-butylamine
0.354 nm. As all of the samples were microporous, the in acetonitrile using an Ag/AgCl electrode [25]. Acid
linear region of the BET equation was taken in the strength was estimated by the initial electrode poten-
range of relative pressures between 0.02 and 0.12 p/p◦ . tial, Ei .
The adsorption isotherms were also used to calculate
the micropore volume and external surface area by a 2.3. Activity measurements
t-plot analysis [22] using the equation of Halsey [23]
to determine the thickness of the adsorbed layer at As in a previous study [15], catalytic measure-
each relative pressure value. The micropore size dis- ments for simultaneous HDS and HDN of gas oil
tributions were determined by the method developed were carried out in a high-pressure continuous-flow
by Mikhail et al. applied to the corresponding t-plots micro-reactor. The catalyst bed consisted of 1 g of
[24]. The mesoporosity was determined from the dif- catalyst diluted 1:1 (v/v) with SiC particles to op-
ference between the microporosity calculated from the timise hydrodynamics. The remaining space in the
t-plot and the volume adsorbed at a relative pressure reactor was filled with SiC particles. Prior to reaction,
of 0.96 on the desorption branch of the correspond- the catalysts were sulphided with a 7 vol.% CS2 /gas
ing isotherms, equivalent to a pore diameter of 50 nm. oil mixture at 350 ◦ C and 2 MPa total pressure for
Mercury intrusion porosimetry (MIP) analyses were 4 h. The feed for HDS and HDN was a commercial
determined on samples previously dried overnight at gas oil, containing 470 ppm S and 190 ppm N. The
110 ◦ C in a Fisons Pascal 140/240 apparatus. Starting performance of the catalysts was determined in the
from vacuum and raising the pressure to 200 MPa this temperature range 325–375 ◦ C under standard con-
technique covers the pore diameters of ∼300 ␮m down ditions: 3 MPa total pressure 9 h−1 LHSV, 3600 h−1
to 7.5 nm applying the Washburn equation. Summa- GHSV, and H2 /feed ratio of 400. Under these reaction
tion of the micro and mesopore volumes obtained from conditions the catalysts were stable and the reaction
N2 isotherms with the pore volumes in pores >50 nm was not controlled by mass transfer phenomena [26].
determined by MIP lead to the total pore volume of In all the experiments a stabilisation period of at least
the sample. 2 h was allowed before the first sample was collected.
The chemical state and surface composition of Total sulphur in the effluents was determined by
the sulphided catalysts were studied by X-ray pho- iodometric titration of SO2 using a LECO analyser,
toelectron spectroscopy (XPS). The XP spectra were and total nitrogen was analysed on a Antek 703C in-
recorded in a VG Escalab 200R electron spectrome- strument by chemiluminescence detection. HDS and
ter equipped with a hemispherical electron analyser HDN conversions were defined as percent of total
and a Mg K␣ (1253.6 eV) photon source. Energy sulphur and nitrogen, respectively, removed from the
corrections were performed employing the C line of initial gas oil.
the carbon support at 284.9 eV as internal reference.
The catalyst samples for XPS were pre-sulphided ex
situ with a mixture of 10% H2 S/H2 at 350 ◦ C for 3. Results and discussion
4 h. The samples were then cooled to room temper-
ature, flushed with He and transferred into flasks 3.1. Textural properties
containing iso-octane. The intensities of the peaks
were estimated by calculating the integral of each The BET areas results of dried Re(x)/C catalysts are
peak after subtracting an S-shaped background and presented in Table 1. The BET area was almost
fitting the experimental curve to a combination of unchanged with Re incorporation of up to 0.076
Gaussian/Lorentzian lines. atoms nm−2 , but then fell severely with higher
154 N. Escalona et al. / Applied Catalysis A: General 240 (2003) 151–160

Fig. 1. Variation of the pore volume as a function of the Re content for Re(x)/C catalysts.

concentrations. This behaviour was repeated with micropores, causing the observed shift in the average
both the micro and total pore volumes, presented micropore size to wider diameters.
in Fig. 1. The mesopore volumes in Fig. 1 and ex-
ternal areas in Table 1, were largely unchanged in 3.2. X-ray diffraction
the studied concentration range. On the contrary, the
macropore volume determined by mercury porosime- The powder XRD patterns of dried Re(x)/C cata-
try rose to a maximum for the sample with 0.076 lysts were recorded. Diffraction lines different from
Re atoms nm−2 , then fell. The average micropore those of activated carbon were observed only in the
diameter from the MP method displayed a shift to XRD pattern of the Re(0.380)/C sample (not shown
wider pores as the Re content was increased, as can here), which displayed clearly two diffraction lines of
be seen in Fig. 2 and Table 1. This last trend was low intensity at approximately 16.9 and 25.9◦ , corre-
in close agreement with BET area and pore volumes sponding closely to the two most intense lines of bulk
results. For alumina-supported Re catalysts the BET NH4 ReO4 . The position and broadening of these XRD
area did not change significantly with the Re contents lines suggests the presence of distorted NH4 ReO4 ag-
up to 0.97 atoms Re nm−2 because the Re is located gregates in the Re(0.380)/C catalyst. These results in-
only in meso and macropores since ␥-Al2 O3 has no dicate that Re species are highly dispersed forming
microporosity [15]. either amorphous or crystalline phase <4 nm at Re
These textural characterisation results demonstrate loadings ≤0.135 atoms nm−2 .
that for the incorporation of quantities of Re lower
than about 0.076 atoms nm−2 the surface area and 3.3. Catalyst acidity
micropore volume were not effected but the macropore
volumes increased, suggesting that most of the Re was Fig. 3 shows the total acidity of calcined Re(x)/C
located outside of the micropores. However, at higher catalysts as a function of the Re content. It is evident
concentrations the reductions of the specific area and that both the number of acid sites and their strength
micropore volume suggest that the Re filled the narrow increased gradually on increasing the Re content. The
N. Escalona et al. / Applied Catalysis A: General 240 (2003) 151–160 155

Fig. 3. Effect of Re content on total acidity and strength of the


acid sites of dried Re(x)/C catalysts. Acid strength estimated by
the initial electrode potential, Ei .

tra revealed two partially overlapping doublets, both


Fig. 2. Pore size distribution for Re(x)/C catalysts.
containing the Re 4f7/2 and 4f5/2 peaks. Table 2 sum-
marises the binding energies (BE) of the most intense
Re 4f7/2 component of each doublet, their relative pro-
acidity results displayed a nearly linear increase in portion and the surface Re/C atomic ratios. The Re
acid strength with Re incorporation up to 0.135 atoms 4f7/2 component of the most intense doublet remained
nm−2 and then a slower rise. This suggested that up constant, at about 41.7 eV over the whole range of
to this concentration the dispersion of the Re salt re- Re loading considered, and corresponds closely to the
mained similar in all of the samples, but above this value reported for ReS2 [6,27,28]. The Re 4f7/2 peak
value formation of aggregates or the accessibility to of the less intense doublet, at about 45.3 eV can be
Re was reduced, due to it entering the narrow micro- assigned to Re(VI) and Re(IV) oxidation states [27].
pores. However, in the case of Re/␥-Al2 O3 catalysts, These observations indicate that the sulphidation of the
both acid strength and total acidity increased with Re supported Re species was slightly incomplete, under
loading similarly to the Re/C catalysts but over the
whole range of Re content, indicating that in that sys- Table 2
tem the fraction of the ␥-Al2 O3 surface covered by Re XPS binding energies (eV), surface atomic ratios and degree of
gradually increased without formation of aggregates sulphidation of sulphided catalysts
[15]. Catalyst Re 4f7/2 (Re/C) × 102 Resulf /
atomic ratio Retotal
3.4. X-ray photoelectron spectra Re(0.024)/C 41.4 (87), 45.5 (12 ) 0.117 0.77
Re(0.076)/C 41.9 (91), 45.2 (9) 0.233 0.84
The XPS of the Re 4f region of sulphided Re(x)/C Re(0.135)/C 41.9 (74), 45.2 (26) 0.434 0.80
Re(0.380)/C 41.8 (83), 45.5 (17) 1.018 0.89
catalysts are shown in Fig. 4. Curve fitting of the spec-
156 N. Escalona et al. / Applied Catalysis A: General 240 (2003) 151–160

occurs as this last support has no micropores but has


a much higher external surface area (194 m2 g−1 )
and, therefore, more Re loading can be incorporated
before multilayer or small three-dimensional particle
formation.

3.5. Activity and selectivity of catalysts

Figs. 6 and 7 show the effect of the Re loading


on the activity for simultaneous HDS and HDN reac-
tions, respectively, of gas oil over Re(x)/C catalysts,
expressed as a function of conversion versus Re con-
tent (atoms nm−2 ) and the reaction temperature. The
variation of the activity with Re loading was simi-
lar for both HDS and HDN reactions, although the
magnitudes were different for both reactions. For both
HDS and HDN, the activity increased with increas-
ing Re loading up to about 0.076 Re atoms nm−2 ,
and then slightly decreased. Similar activity trends
for thiophene HDS [10] and for gas oil HDS and
HDN [15] were previously reported over Re/␥-Al2 O3
catalysts.
These results for HDS and HDN activities could be
compared and contrasted with those obtained from the
textural characterisation of the catalysts, the acidity of
Fig. 4. X-ray photoelectron spectra of the Re 4f region of sulphided the materials and the Re/C ratio by XPS. Acidity re-
Re(x)/C catalysts. sults displayed a nearly linear increase of acidity with
Re incorporation up to 0.135 atoms nm−2 and then a
slower rise. This behaviour indicated that up to 0.135
the sulphidation conditions used. A slight increase in atoms nm−2 the dispersion of the Re salt remained
the degree of Re sulphidation with increasing Re con- similar in all of the samples, but above this value there
tent for the Re/C catalysts (Table 2) reflects changes was a probable formation of aggregates or the acces-
in Re dispersion. sibility to Re was reduced due to it entering the nar-
The variation of the surface XPS Re/C atomic ra- row micropores. The ratio of Re/C by XPS showed
tio as a function of the nominal Re content in the a similar trend of an almost linear rise up to 0.135
catalysts is shown in Fig. 5. The sulphided Re phase atoms nm−2 and then a slower increase. On the con-
appears to be monolayer-like dispersed up to about trary, the activity results for both HDN and HDS at
0.135 Re atoms nm−2 . The observed deviation from the three studied temperatures displayed a maximum
linearity above 0.135 Re atoms nm−2 indicates a for the catalyst with 0.076 Re atoms nm−2 . Therefore,
change in Re dispersion, the formation of multilayers this maximum in activity may be related to the tex-
or small three-dimensional ReS2 particles, in agree- tural characteristics of this series of catalysts, where
ment with the change also observed for textural and it was noted that up 0.076 atoms nm−2 the specific
acidity results of catalysts in the oxidic state. This area, micropore and total pore volumes were at their
view is supported by the XRD results which show highest. With further Re content, although the cata-
presence of NH4 ReO4 crystallites at 0.380 Re atoms. lysts had higher acidities and Re still well dispersed
In Re/␥-Al2 O3 catalysts the deviation from linearity (XPS results), the reduction in activity seems to be
occurs >0.5 Re atoms nm−2 . [15] These differences associated with the fall in the pore volume and area
between carbon- and ␥-Al2 O3 -supported catalysts being more significant. At contents up to 0.076 Re
N. Escalona et al. / Applied Catalysis A: General 240 (2003) 151–160 157

Fig. 5. Relationship between the XPS Re/Al atom ratio and the nominal surface density of Re for Re/␥-Al2 O3 catalysts; the dashed straight
line is the best-fit linear correlation at low Re loading.

atoms nm−2 , as the concentration is very low, the Re this content more Re enters the micropores and re-
located inside the micropores is highly dispersed and duces their accessibility or blocks the narrower ones,
accessible to the reactant molecules, leading then to sterically inhibiting the diffusion of the reactants to
a linear increase in HDS and HDN activities. Above the active sites due to the large size of the gas oil

Fig. 6. Gas oil HDS activity of Re(x)/C catalysts as a function of Re loading and reaction temperature.
158 N. Escalona et al. / Applied Catalysis A: General 240 (2003) 151–160

Fig. 7. Gas oil HDN activity of Re(x)/C catalysts as a function of Re loading and reaction temperature.

molecules. This change at 0.076 Re atoms in the Re catalysts, since both HDS and HDN reactions were
dispersion into the micropores was, however, appar- measured simultaneously, the low apparent activation
ently not reflected either by the XPS results because energy for HDN cannot be attributed to pore-diffusion
this technique is only sensitive to a surface layer, or limitations but was probably caused by differences in
in the acidity results because the n-butylamine can not sorption capacities of the nitrogen and sulphur com-
enter the narrow micropores. With ␥-Al2 O3 supported pounds on the catalysts.
Re catalysts [15], the decrease in HDS and HDN ac- In Figs. 6 and 7 and Table 3, it was observed
tivity occurred at higher Re concentrations than in the that the Re(0.076)/C catalyst had the highest activ-
carbon supported ones because the ␥-Al2 O3 is not mi- ity in both HDS and HDN reactions and also the
croporous and, consequently, in this support the loss highest apparent activation energy. This apparently
in activity is essentially due to the reduction in the Re paradoxical behaviour can be readily explained from
dispersion at much higher loadings. the compensation relationship [29]. For Re/C cat-
The apparent activation energies (Eap ) calculated alysts the calculated crude isokinetic temperatures
from Arrhenius plots are given in Table 3. The Eap (Tiso ) are 300 ± 8 ◦ C and 295 ± 14 ◦ C for HDS and
were in the range 116–156 kJ mol−1 for HDS and HDN reactions, respectively. Thus, considering that
24–30 kJ mol−1 for HDN, and similar to the corre- conversions were measured at reaction temperatures
sponding 138–158 and 25–33 kJ mol−1 values found higher than both Tiso , the most active catalyst must
over Re/␥-Al2 O3 catalysts [15]. As with Re/␥-Al2 O3 have the highest Eap . This trend has been reported
for HDS in Co-Mo/␥-Al2 O3 , Ni-Mo/␥-Al2 O3 [29]
and W/␥-Al2 O3 catalysts [30], and recently in Ni-Re
Table 3
Apparent activation energy (Eap ) in the HDS and HDN of gas oil
catalysts [31].
The fact that the apparent activation energies for
Catalyst HDS (kJ mol−1 ) HDN (kJ mol−1 ) HDS over Re(x)/C catalysts differ markedly from
Re(0.024)/C 137 ± 5 27 ± 8 those for HDN indicates that the two reactions involve
Re(0.076)/C 156 ± 10 30 ± 7 different types of catalytic sites, as has generally been
Re(0.135)/C 123 ± 6 28 ± 2 established in the literature [32–34]. Higher activa-
Re(0.380)/C 116 ± 1 24 ± 2
tion energies for HDS compared to those for HDN
N. Escalona et al. / Applied Catalysis A: General 240 (2003) 151–160 159

Fig. 8. HDN/HDS selectivity of Re(x)/C catalysts as a function of Re loading and reaction temperature.

obviously means that the former reaction was rela- content range studied and followed a similar trend to
tively more favoured at higher reaction temperatures that observed for the catalyst acidity, both the number
than the latter reaction and reflected in the change of acid sites and their strength. Therefore, it was con-
of HDN/HDS selectivity. Accordingly, Fig. 8 shows sidered that the observed increase in the HDN/HDS
that the HDN/HDS selectivity increased with de- selectivity was more related to the increase in catalyst
creasing reaction temperature, being <0.8 at 375 ◦ C acidity rather than to changes in the size and morphol-
and >3 at 325 ◦ C for all the catalysts. This may also ogy of the ReS2 slabs. It is well known that the acid
explain the apparently contradictory results on the sites promote the C–N bond scission reaction [32,33].
HDS and HDN activities of unsupported Re catalysts
previously reported in the literature [9,35], since they 3.6. Comparison of selectivity/selectivity of carbon
were obtained under different reaction conditions, and alumina-supported catalysts
particularly temperature.
Another interesting result shown in Fig. 8 is that It is interesting to compare the activities and selec-
the HDN/HDS selectivity increased gradually with in- tivities of these catalysts on carbon with those from a
creasing Re content. Since the Re content does not previous study on ␥-Al2 O3 [15]. At surface Re load-
modify the nature of the active sites for either reaction ings below about 0.2 atoms nm−2 , Re(x)/C catalysts
(catalyst had Eap almost constant for both reactions), it had higher HDS activity per gram of catalyst and
suggests that Re content changes the relative concen- also per metal atom than equivalent Re(x)/␥-Al2 O3
trations of the sites for HDS and HDN reactions, in- catalysts, and on carbon the intrinsic HDS activity
creasing relatively more those for HDN. This relative per metal atom decreased abruptly with loading while
increase in the overall HDN activity does not seem to on alumina increased smoothly not shown here. For
be related to a change in the size and the stacking of the the same range of metal loading, similar trends were
ReS2 slabs since according to the XPS results (Fig. 3) previously reported by Arnoldy et al. [10] for the
Re dispersion only changed at very high Re loading thiophene HDS over carbon- and alumina-supported
and then only moderately. However, the increase of the Re catalysts. The behaviour for intrinsic HDN activ-
HDN/HDS selectivity was gradual throughout the Re ity showed similar trends as those observed for HDS
160 N. Escalona et al. / Applied Catalysis A: General 240 (2003) 151–160

activity but with more pronounced differences and up [4] M. Breysse, J.L. Portefaix, M. Vrinat, Catal. Today 10 (1991)
to Re loadings of about 0.4 atoms nm−2 . 689.
[5] T. Pecoraro, R.R. Chianelli, J. Catal. 67 (1981) 430.
Comparison of the HDN/HDS selectivties revealed [6] J.P.R. Vissers, C.K. Groot, E.M. van Oers, V.H.J. de Beer,
that the Re(x)/C catalysts were about twice as selec- R. Prins, Bull. Soc. Chim. Belg. 93 (1984) 813.
tive for HDN than the Re(x)/␥-Al2 O3 catalysts over [7] M.J. Ledoux, O. Michaux, G. Agostini, J. Catal. 102 (1986)
the whole range of metal loading and reaction tem- 275.
[8] S. Eijsbouts, V.H.J. de Beer, R. Prins, J. Catal. 109 (1988)
perature studied. A similar effect was found for Mo 217.
supported on carbon in comparison to the conven- [9] S. Eijsbouts, V.H.J. de Beer, R. Prins, J. Catal. 127 (1991)
tional Ni-Mo/␥-Al2 O3 catalysts [36]. The increased 619.
HDN selectivity of Re sulphide deposited on carbon [10] P. Arnoldy, E.M. van Oers, V.H.J. de Beer, J.A. Moulijn, R.
Prins, Appl. Catal. 48 (1989) 241.
compared to on alumina suggests that the carbon [11] J. Räty, T.A. Pakkanen, Catal. Lett. 65 (2000) 175.
support causes a small additional effect on selectivity [12] J. Quartararo, S. Mignard, S. Kasztelan, J. Catal. 192 (2000)
towards the HDN reaction. In principle, the support 307.
might introduce structural and textural modifications [13] E.W. Stern, J. Catal. 57 (1979) 390.
[14] J. Shabtai, Q. Guohe, K. Balusami, N.K. Nag, F.E. Massoth,
of the active phase [20] which were more favourable J. Catal. 113 (1998) 206.
to HDN than HDS. Thus, recent results show that the [15] N. Escalona, J. Ojeda, R. Cid, G. Alvez, A. López Agudo,
morphology of MoS2 supported on carbon, due to its J.L.G. Fierro, F.J. Gil Llambı́as, Appl. Catal., in press.
[16] T.C. Ho, A.J. Jacobson, R.R. Chianelli, C.R.F. Lund, J. Catal.
high dispersion and high fraction of corner sites, leads
138 (1992) 353.
to high HYD activities [37]. This higher HYD activ- [17] Z. Vı́t, M. Zdrazil, J. Catal. 119 (1989) 1.
ity of carbon-supported catalysts compared with the [18] J. Cinibulk, Z. Vı́t, Appl. Catal. 204 (2000) 107.
corresponding alumina supported ones could explain [19] H. Topsøe, B.S. Clausen, F.E. Massoth, Hydrotreating cata-
lysts, in: J.R. Anderson, M. Boudart (Eds.), Catalysis Science
the increased HDN/HDS selectivity. The carbon could
and Technology, vol. 11, Springer, Berlin, 1996.
also take part directly in some of the reaction steps, [20] R.R. Chianelli, T.A. Pecoraro, J. Catal. 198 (2001) 9.
for instance, its surface oxygenated acidic groups [38] [21] S. Brunauer, P.H. Emmett, E. Teller, J. Am. Chem. Soc. 60
participating in the cleavage of the C–N bond [32,33]. (1938) 309.
[22] B.C. Lippens, B.G. Linsen, J. H de Boer, J. Catal. 3 (1964)
Another possibility is that the increased HDN selec- 32.
tivity on carbon-supported catalysts could stem from [23] J.H. Singleton, G.D. Halsey, Can. J. Chem. 33 (1955) 184.
the formation of a surface Re-carbide layer, similar to [24] R.S. Mikhail, S. Brunauer, E.E. Bodor, J. Colloid Interf. Sci.
that recently reported for Mo sulphide catalysts [20], 26 (1968) 45.
[25] R. Cid, G. Pechi, Appl. Catal. 14 (1985) 15.
since metal carbides have shown to be highly active [26] F.J. Gil Llambı́as, A. López Agudo, Anal. Quim. Suppl. 1
for HDN reactions [39]. (1978) 1.
[27] P. Clark, B. Dhandapani, S.T. Oyama, Appl. Catal. A 184
(1999) L175.
Acknowledgements [28] A.N. Startsev, V.N. Rodin, V.I. Zaikovskii, A.V. Kalinkin,
V.V. Kriventsov, D.L. Kochubei, Kinet. Catal. 38 (1997) 548.
[29] L. Boussieres, C. Flores, J. Poblete, F.J. Gil Llambı́as, Appl.
Financial support from Projects 1990496-6 and Catal. 23 (1986) 271.
1020043 FONDECYT (Chile), and from the Program [30] F.J. Gil Llambı́as, J. Salvatierra, L. Boussieres, M. Escudey,
CYTED, Subprogram V (Spain) is kindly acknowl- Appl. Catal. 59 (1990) 185.
[31] P. Baeza, J. Ojeda, N. Escalona, G. Alvez, R. Garcı́a, R. Cid,
edged. N. Escalona gratefully acknowledges the fel- F.J. Gil Llambı́as, Bol. Soc. Chil. Quim. 46 (2001) 415.
lowship from CONICYT (Chile). Empresa Nacional [32] S.H. Yang, C.N. Satterfield, J. Catal. 81 (1983) 168.
de Petróleo Chile (ENAP). [33] G. Perot, Catal. Today 10 (1991) 447.
[34] B. Delmon, Bull. Soc. Chim. Belg. 104 (1995) 173.
[35] C.J.H. Jacobsen, E. Törnqvist, H. Topsøe, Catal. Lett. 63
References (1999) 179.
[36] Z. Vı́t, Catal. Lett. 13 (1992) 131.
[37] E.J.M. Hensen, P.J. Kooyman, Y. van der Meer, A.M. van der
[1] K.G. Knudsen, B.H. Cooper, H. Topsøe, Appl. Catal. A 189 Kraan, V.H.J. de Beer, J.A.R. van Veen, R.A. van Santen, J.
(1999) 205. Catal. 199 (2001) 224.
[2] P. Grange, X. Vanhaeren, Catal. Today 36 (1997) 375. [38] G.J. McDougall, R.D. Hancock, Min. Sci. Eng. 12 (1980) 85.
[3] F. Luck, Bull. Soc. Chim. Belg. 100 (1991) 781. [39] R. Prins, Adv. Catal. 46 (2001) 399.

Das könnte Ihnen auch gefallen