You are on page 1of 11

Applied Catalysis A: General 324 (2007) 919

www.elsevier.com/locate/apcata

Hydrogenation and ring opening of naphthalene on bulk


and supported Mo2C catalysts
Shahrzad Jooya Ardakani 1, Xuebin Liu, Kevin J. Smith *
Department of Chemical and Biological Engineering, University of British Columbia, Vancouver, BC, Canada V6T1Z3
Received 26 October 2006; received in revised form 16 February 2007; accepted 25 February 2007
Available online 4 March 2007

Abstract
A series of Mo2C catalysts have been investigated for the hydrogenation and ring opening of naphthalene. At 573 K and 3 MPa H2, bulk Mo2C
showed no selectivity for ring-opening products (ROP) and, although treating the bulk catalysts in oxygen increased stability and naphthalene
conversion, ROP selectivity remained low. Supporting the Mo2C on HY zeolites significantly increased the hydrogenation of naphthalene and the
subsequent formation of ROP, as compared to the HY zeolites alone or the bulk materials. ROP yields were dependent upon the HY zeolite
SiO2:Al2O3 ratio and the Mo2C loading, the optimum loading being dependent upon the zeolite acidity. The maximum ROP yield of 33 wt.% was
obtained with 7.4 wt.% Mo2C supported on HY zeolite of moderate acidity (SiO2:Al2O3 ratio of 12) and the yield is comparable to yields reported
for noble metal catalysts on acidic supports. The Mo2C/HY zeolite bifunctional catalysts are most effective for ring opening of naphthalene when
the acidity is adjusted to an intermediate level that limits deactivation and the Mo2C loading provides sufficient hydrogenation capability to achieve
high conversion to the primary product tetralin.
# 2007 Elsevier B.V. All rights reserved.
Keywords: Ring opening; Catalysts; Bifunctional catalysts; Molybdenum carbide; Zeolite

1. Introduction
Selective ring opening (SRO) refers to the hydrogenolysis of
endocyclic CC bonds present in C5 and C6 cycloparaffins.
Interest in SRO stems in part from the need to reduce the aromatic
and cycloparaffinic content of bitumen-derived gas oils,
especially as bitumen-derived synthetic crude oil production
from the Canadian oilsands is expected to double in the next 10
years. Bitumen-derived heavy gas oil (HGO) has approximately
90% cycloparaffins plus aromatics, versus 60% in a typical
paraffinic HGO [1]. The aromatic content of bitumen-derived
HGO can be reduced through hydrogenation processes [2,3],
however, the cycloparaffinic content remains high, and these
compounds have low cetane numbers. Consequently, selective
ring opening of the cycloparaffins without a reduction in carbon
number of the product molecule, is desired.

* Corresponding author. Tel.: +1 604 822 3601; fax: +1 604 822 6003.
E-mail address: kjs@interchange.ubc.ca (K.J. Smith).
1
Department of Chemistry, Sharif University of Technology, PO Box 113659516, Tehran, Iran.
0926-860X/$ see front matter # 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2007.02.048

SRO of cycloparaffins can occur via thermal free radical


reactions, via acid catalysed carbocation chemistry or by metal
catalysed hydrogenolysis [4]. The thermal route is not of
interest because of poor selectivity, while the acid catalysed
route generally yields low SRO selectivity because the rate of
b-scission of endocyclic CC bonds is much lower than that of
exocyclic carbon. Consequently, metal catalysts or metal/alloy
catalysts are preferred for SRO and a number of noble metals
have been reported to have good activity and selectivity,
including Rh, Ru, Ir, Pt, Pd/Pt and Pt/Ir [410]. However, these
metals are costly and have relatively low resistance to sulphur
poisoning, an important consideration since bitumen-derived
HGO will contain high levels of S (800 ppm).
Although opening of the ring can be accomplished on either
acid or metal catalysts, the combination of the two functions is
much more effective than each of them alone [1]. Dehydrogenation, cracking, isomerization and dealkylation reactions
can take place on the acidic sites, while the metal sites catalyse
hydrogenation and hydrogenolysis. Mild acidity is necessary to
avoid exocyclic carbon cracking, but is also thought to be
necessary for improved sulphur resistance (although recently
the role of the acid support in this regard has been questioned

10

S.J. Ardakani et al. / Applied Catalysis A: General 324 (2007) 919

[10]). The acid function plays another important rolenamely


as a catalyst for isomerization reactions, especially C6 ! C5
ring shrinkage since it is well established that hydrogenolysis of
C5 cycloparaffins is easier than C6 cycloparaffins [1].
Metal carbides and nitrides are known to have good
hydrogenolysis activities and a number of studies have reported
on their activity for various alkane hydrogenolysis reactions.
Neylon et al. [11] reported activities of Mo2N for n-butane
hydrogenolysis similar to that of Pt-Sn/Al2O3. Activities of
Mo2C were lower. Other studies have reported on metal carbide
hydrogenolysis activities [12] and have shown that exposing
these catalysts to oxygen to produce an oxycarbide, results in a
reduction in catalyst hydrogenolysis activity in favor of
isomerization [13,14]. These experiments were conducted
using cyclohexane and cyclopentane. Furthermore, ring
enlargement reactions (the reverse of that desired in SRO)
have been shown to be catalysed by metal oxycarbides [13].
Metal oxycarbides have also been shown to have good
selectivity for alkane reforming [15,16] via acid catalysis.
All of these reactions are of potential importance to selective
ring opening of cycloparaffins, yet none of these studies have
investigated the isomerization and/or hydrogenolysis of C5 and
C6 cycloparaffins. The metal carbides have been shown to be
resistant to sulphur and have been proposed as HDN and HDS
catalysts. At the lower temperature of SRO, it is expected that
these materials could operate in the presence of sulphur, but this
needs to be proven experimentally.
In the present study, both bulk and supported Mo2C catalysts
have been investigated as catalysts for the vapor-phase
hydrogenation and SRO of naphthalene at 3.0 MPa H2.
Naphthalene was chosen as a model reactant since it can
undergo hydrogenation, isomerization and ring-opening reactions.
2. Experimental
2.1. Catalyst preparation
Bulk Mo2C was synthesized by temperature-programmed
reaction. Approximately 2 g of MoO3 (MoO3, +99.5%, Sigma
Aldrich) was placed in a quartz tubular reactor and heated at a
rate of 5 K/min from room temperature to 973 K in
330 ml (STP)/min of H2/CH4 (1:1). The final temperature
was held for 3 h before rapidly cooling the sample to room
temperature in a flow of H2 (150 ml (STP)/min). The catalyst
was subsequently passivated by exposing it to a mixture of 2%
O2 in He at a flowrate of 100 ml (STP)/min for a period of 17 h.
Oxidized samples were also prepared by passivating the
catalyst at 623 K for 1 h (Mo2C-O623) and at 648 K for 1 h
(Mo2C-O648).
Mo2C catalysts supported on HY zeolite, with Mo loadings
of 2, 5, 7 and 10 wt.% were prepared by wet impregnation of the
zeolite using an ammonium heptamolybdate tetrahydrate
(MoO3 81.083.0%, Sigma) aqueous solution containing the
required amount of Mo. Three commercial zeolite Y samples
(CBV720 (Y1), CBV712 (Y2), CBV400 (Y3), Zeolyst
International), with SiO2/Al2O3 ratios of 30, 12 and 5.1,

respectively, were used as the supports. Prior to use, Y2, which


was in the ammonia form, was calcined at 773 K for 3 h to
convert it to the hydrogen form. Y1 and Y3 were used as
supplied. After impregnation, the catalysts were dried at 373 K
for at least 20 h and then calcined at 773 K for 5 h. The calcined
catalysts were placed in a quartz tubular reactor and heated at a
rate of 5 K/min from room temperature to 973 K in
330 ml (STP)/min of H2/CH4 (1:1). The final temperature
was held for 3 h before rapidly cooling the sample to room
temperature in a flow of H2 (150 ml (STP)/min). The catalyst
was subsequently passivated by exposing it to a mixture of 2%
O2 in He at a flowrate of 100 ml (STP)/min for a period of 1 h.
These catalysts are designated Mo2C(i)/Yj, where i is the
nominal Mo content (wt.%) and j is the zeolite.
2.2. Catalyst characterization
The passivated catalysts were characterized by powder Xray diffraction (XRD), X-ray photoelectron spectroscopy
(XPS), temperature-programmed desorption (TPD) of npropylamine, CO pulse chemisorption and BET surface area
analysis. XRD analyses were performed on the passivated
catalysts using a Siemens D500 Cu Ka X-ray source of
. The analysis was performed using a scan
wavelength, 1.54 A
range of 3708 with a step size of 0.048 and step time of 2 s. The
phase identification was carried out after subtraction of the
background using standard software. Crystallite size (dc)
estimates were made using the Scherrer equation, dc = Kl/
b cos u where the constant K was taken to be 0.9, l is the
wavelength of radiation, b is the peak width in radians and u is
the angle of diffraction.
XPS analysis was done using a Leybold Max 200
spectrometer using Al Ka radiation as the photon source,
generated at 15 kV and 20 mA. The pass energy was set at
192 eV for the survey scan and at 48 eV for the narrow scan.
The catalysts were analysed after the passivation step without
further treatment and all XPS spectra were corrected to the C 1s
peak at 285.0 eV.
Transmission electron microscope (TEM) images were
generated using a Fei Tecnai 20 scanning transmission electron
microscope operating at 200 kV. Catalyst samples were ground
to a fine powder, dispersed in ethanol and sonicated for 2 h. A
drop of the catalyst suspension was placed on a 200 mesh
copper grid coated with formvar and carbon, and left to dry
before analysis.
The catalyst Brnsted acid sites were titrated by n-propyl
amine (n-PA) temperature-programmed desorption. The TPD
was conducted in a stainless steel reactor (i.d. = 9 mm)
containing 0.5 g of catalyst that was pre-treated at 773 K for
1 h in 10% H2 in Ar at a flowrate of 60 ml (STP)/min. The
sample was cooled to 383 K. A flow of He (30 ml/min)
saturated at room temperature with n-propyl amine (Aldrich,
99.8%) passed through heated gas lines to the reactor
containing the pre-treated catalyst. After a 2 h adsorption at
383 K, the reactor was flushed in pure He at a flowrate of
30 ml (STP)/min for 1 h to ensure that physically adsorbed nPA was removed. The chemisorbed n-PA was then desorbed by

S.J. Ardakani et al. / Applied Catalysis A: General 324 (2007) 919

ramping the reactor temperature from 383 to 983 K at a rate of


5 K/min and a thermal conductivity detector was used to
quantify the amount of n-PA desorbed. The system was
calibrated using zeolite samples of known acidity.
The CO uptake of the bulk catalysts was estimated by pulsed
chemisorption. The passivated catalysts were pre-treated in
60 ml (STP)/min of 10% H2 in Ar while heating from 313 to
723 K at a rate of 2 K/min, and maintaining the final
temperature for 1 h, to remove the passivation layer. The
reactor was cooled to 298 K in the flow of H2 and flushed in He
at 30 ml (STP)/min for 30 min, before 1 ml pulses of CO were
injected into the flow of He (30 ml/min) that passed through the
reactor. CO pulses were repeatedly injected until the response
from the thermal conductivity detector showed no further CO
uptake after consecutive injections.
BET surface areas were measured by a single-point N2
adsorption at 196 8C using a Flowsorb 2300 (Micromeritics)
analyzer. A 30% N2/70% He mixture, fed at 15 ml (STP)/min
was used for single-point surface area measurements. Samples
were degassed at 250 8C for 2 h prior to the measurements.
2.3. Catalyst activity
Naphthalene hydrogenation was chosen as the model reactant
for the hydrogenation and ring opening of cycloparaffins. The
reaction was carried out at temperatures of 553613 K in H2 at a
total pressure of 3.0 MPa. Approximately 1 g of passivated
catalyst (dp < 0.7 mm) was placed in the isothermal zone of the
fixed-bed reactor (i.d. = 9 mm). The passivated catalyst was
activated in 200 ml (STP)/min of H2 at 723 K for 1 h. The reactor
temperature was then adjusted to the desired reaction
temperature before a 10 wt.% solution of naphthalene in nheptane was fed to the reactor at a rate of 0.16 ml/min using a
Gilson Model 0154E metering pump. Prior to entering the
reactor, the liquid was evaporated into a stream of flowing H2
(58 ml (STP)/min). Each catalyst test was performed for a period
of at least 5 h and the liquid products of the reaction were
collected periodically (every hour) and analysed using a 3400 GC
Varian Star Gas Chromatograph equipped with a flame ionization
detector (FID). Component separation was achieved using a
capillary column (CP-Sil 19 CB, 25 m length and 0.53 mm i.d.).
Component identification was confirmed using the same column
and a GC-MS (Agilent 6890/5973N). To facilitate the discussion,
reaction products were grouped as follows(a) LP: light
hydrocarbons, mainly alkylbutane and alkylpentane; (b) THN:
tetralin or 1,2,3,4-tetrahydonaphthalene; (c) DHN: decalin or
decahydronaphthalene (cis and trans isomers); (d) ROP: ringopening products, mainly alkylcyclohexanes, alkylbenzenes and
alkylindenes; (e) HP: aromatics and naphthenes with more than
10 carbon atoms, mainly alkyltetralins.
A wide range of reaction conditions was investigated
(temperature and contact time) over one of the catalysts.
Furthermore, since all potential real feedstocks contain small
amounts of sulphur that can poison the catalyst, the thioresistance of one of the Mo2C catalysts was also investigated by
adding dibenzothiophene (DBT) to the naphthalene/n-heptane
feed mixture.

11

3. Results
3.1. Catalyst characterization
X-ray diffractograms of the passivated bulk Mo2C, and the
Mo2C treated in O2/He at elevated temperature are shown in
Fig. 1. The peak positions are consistent with the presence of
crystalline b-Mo2C. There was no evidence of crystalline
molybdenum oxide in the bulk catalyst passivated at room
temperature for 17 h. The catalysts passivated in the same gas
mixture for 1 h at 623 and 648 K show low intensity features at
2u = 278 and these can be attributed to the formation of
molybdenum oxide. Passivation at 673 K resulted in complete
conversion of the carbide to the oxide.
The crystallite size of the bulk catalysts, estimated via
XRD line broadening, and their corresponding BET surface
areas, are given in Table 1. The data show that the TPR
yielded low surface area Mo2C, and that upon exposure to O2
at elevated temperature (623 and 648 K) the crystallite size of
the Mo2C increased, whereas the BET area decreased.
Although the Mo2C was still present after the passivation/
oxidation procedure, the increasing crystallite size and
decreasing BET area suggest some inclusion of oxygen into
the Mo2C and this may be as a consequence of the formation
of molybdenum oxycarbide. The measured physico-chemical

Fig. 1. XRD diffractograms of bulk Mo2C catalysts treated in oxygen (see text
for conditions).

12

S.J. Ardakani et al. / Applied Catalysis A: General 324 (2007) 919

Table 1
Physico-chemical properties of the bulk molybdenum carbide catalysts
Catalyst

Mo2C
Mo2C-O623
Mo2C-O648
a

Passivation
Time (h)

Temperature (K)

17
1
1

300
623
648

SBET (m2/g)

Crystallite size (nm)

Total aciditya (mmol/g)

CO uptake (mmol/g)

8.5
3.9
1.9

9.3
15.2
14.4

0.020
0.002
0.002

1
4
2

Total acidity measured by TPD of n-propylamine.

properties of the bulk Mo2C are also shown in Table 1. Both


the CO chemisorption and total acidity as determined by n-PA
TPD are expectedly low on all the bulk catalysts. The CO
adsorption data of Table 1 suggests that about 0.02% of the
total Mo is available at the surface of the bulk Mo2C.
Treatment in oxygen at elevated temperature decreased the
acidity. Nonetheless for all bulk catalysts the acidity was very
low.
X-ray diffractograms for the HY zeolites and the Mo2C
supported on the zeolites are presented in Fig. 2. These data
show that the zeolite structure was preserved during the
carburization process and there was no discernable difference
between the diffractograms obtained for the zeolites and the
Mo2C supported on the zeolite. Furthermore, the presence of
molybdenum carbide was not apparent from the XRD data. The
measured physico-chemical properties of the Mo2C(i)/Yj
catalysts are shown in Table 2. The BET areas of the
Mo2C(i)/Yj catalysts all decreased with increased Mo loading.
The total acidity of the zeolites followed the expected trend of
increased acidity (Y1 < Y2 < Y3) with decreasing SiO2/Al2O3
ratio (Y1 = 30; Y2 = 12; Y3 = 5.1), and upon addition of the
Mo2C, a further decrease in acidity occurred. In all cases the
acidity of the zeolite-supported Mo2C catalyst was much
greater than the bulk catalysts. For any particular zeolite

support, the acidity remained relatively constant as the Mo


loading increased.
Table 2 also reports the surface composition of the zeolitesupported Mo2C catalysts, as determined by XPS. Note that as
the Mo loading increased the surface Mo concentration also
increased. Furthermore, the Si/Al ratio varied according to the
nominal composition of the zeolite support, but there was no
consistent trend in how the ratio changed as the Mo loading
increased, and this observation is consistent with the acidity
data measured by n-PA TPD. The Mo 3d region of the XPS
spectra are plotted in Figs. 3 and 4. In all cases, peaks at
228.7, 232.2232.5 and 235.9236.4 eV were evident. The
low BE peak at 228.7 eV was assigned to the Mo 3d5/2 BE
associated with Mo2C, in agreement with [17], whereas the
higher BE peaks correspond to Mo 3d5/2 and 3d3/2 BE for
MoO3 [18,19]. The XPS data provide evidence for the
presence of Mo2C on the supported catalyst, and given that
this phase was not apparent from the XRD diffractograms of
the same catalysts, we conclude that the Mo2C was well
dispersed on the zeolite.
The TEM data of Fig. 5 also show the presence of welldispersed Mo2C on the support. The estimated particle size of
the Mo2C is 10 nm, although a relatively wide particle size
distribution is apparent.

Fig. 2. XRD diffractograms of Mo2C catalysts supported on HY zeolites.

S.J. Ardakani et al. / Applied Catalysis A: General 324 (2007) 919

13

Table 2
Physico-chemical properties of the zeolite-supported molybdenum carbide catalysts
Sample

Zeolite Y1
Mo2C(2)/Y1
Mo2C(5)/Y1
Mo2C(7)/Y1
Zeolite Y2
Mo2C(2)/Y2
Mo2C(5)/Y2
Mo2C(7)/Y2
Mo2C(10)/Y2
Zeolite Y3
Mo2C(5)/Y3
a

Mo2C loading nominal (wt.%)

2.1
5.3
7.4

2.1
5.3
7.4
10.6

5.3

SBET (m2/g)

711
658
574
483
726
562
500
443
344
604
497

Total aciditya (mmol/g)

0.84

0.75
0.39
1.48
0.69
0.78
0.71
0.62
1.87
0.78

XPS
Mo/(Si + Al) (%)

Si/Al

1.0
2.5
8.1

1.2
1.7
6.7

2.8

20.3
13.4
18.0

3.7
5.6
4.0

1.7

Total acidity measured by TPD of n-propylamine.

3.2. Catalyst activity


Fig. 6 shows the naphthalene conversion as a function of
time-on-stream for the bulk Mo2C catalysts measured at 573 K
and 3 MPa H2 with 10 wt.% naphthalene in n-heptane. Table 3
shows the corresponding product distribution obtained after 2
and 5 h time-on-stream. The data show that significant changes
in catalyst activity occurred over a period of 5 h. The data also
show that although the oxidative treatment increased the
naphthalene conversion after 5 h time-on-stream, none of the
bulk catalysts had high selectivity to the desired ring-opening
products.
The naphthalene conversion and product distribution also
varied as a function of time-on-stream for the supported
catalysts as shown by the data of Fig. 7 for the series of Mo2C
catalysts supported on the Y2 zeolite. The plots do not include
the selectivity to decalin or cracked (LP) products, since the
Fig. 3. Mo 3d region of XPS spectra for Mo2C catalysts supported on HY
zeolite Y1.

Fig. 4. Mo 3d region of XPS spectra for 5 wt.% Mo2C catalysts supported on


HY zeolites Y1, Y2 and Y3.

Fig. 5. TEM micrograph of Mo2C(5)/Y1 catalyst after passivation.

14

S.J. Ardakani et al. / Applied Catalysis A: General 324 (2007) 919

Fig. 6. Naphthalene conversion over bulk Mo2C catalysts. Reaction conditions:


573 K, 3.0 MPa H2, SV = 7.27 h1, H2/naphthalene = 30 mol/mol.

selectivity to both was less than 5% in all cases. Fig. 7A shows


that the naphthalene conversion increased significantly with
addition of Mo2C to the zeolite support, and the activity loss
with time-on-stream was significantly decreased compared to
Y2 alone. Furthermore, the naphthalene conversion was
dependent upon the Mo2C loading, reaching a maximum with
the Mo2C(7)/Y2. The addition of Mo2C to the zeolite also
shifted the product distribution from heavy products to ringopening products (ROP) and tetralin. Similar trends are shown
by the data of Fig. 8 for the Mo2C supported on zeolite Y1, the
least acidic zeolite. Fig. 9 compares the naphthalene conversion
and product distribution obtained with zeolite Y3, the most
acidic zeolite, and the 5% Mo2C supported on zeolite Y3.
Although similar trends in naphthalene conversion were
observed compared to the data of Figs. 7 and 8, in this case,
addition of Mo2C did not impact the ROP selectivity
significantly.
The ring-opening products from naphthalene are secondary
products, as are some of the heavy products. Consequently,

Table 3
Naphthalene conversion and product selectivities of the bulk catalysts at 573 K and 3 MPa H2 after 2 and 5 h time-on-stream
Catalyst

Bulk Mo2C
Bulk Mo2C-O623
Bulk Mo2C-O648

Conversion (wt.%)

Selectivity

2h

Tetralin (wt.%)

91
87
91

5h

51
97
97

Decalin (wt.%)

Other Products (wt.%)

2h

5h

2h

5h

2h

5h

94
64
71

99
71
81

6
34
22

1
24
12

0
2
7

0
5
7

Fig. 7. Naphthalene conversion (A) and product selectivities (BD) obtained on () zeolite Y2, (&) Mo2C(2)/Y2, (*) Mo2C(5)/Y2, (^) Mo2C(7)/Y2 and (~)
Mo2C(10)/Y2. Reaction conditions: 573 K, 3.0 MPa H2, SV = 7.27 h1, H2/naphthalene = 30 mol/mol.

S.J. Ardakani et al. / Applied Catalysis A: General 324 (2007) 919

15

Fig. 8. Naphthalene conversion (A) and product selectivities (BD) obtained on () zeolite Y1, (&) Mo2C(2)/Y1, (*) Mo2C(5)/Y1 and (^) Mo2C(7)/Y1. Reaction
conditions: 573 K, 3.0 MPa H2, SV = 7.27 h1, H2/naphthalene = 30 mol/mol.

Fig. 9. Naphthalene conversion (A) and product selectivities (BD) obtained on () zeolite Y3 and (*) Mo2C(5)/Y3. Reaction conditions: 573 K, 3.0 MPa H2,
SV = 7.27 h1, H2/naphthalene = 30 mol/mol.

16

S.J. Ardakani et al. / Applied Catalysis A: General 324 (2007) 919

Fig. 10. ROP yields at 2 and 5 h time-on-stream for Mo2C supported on (*) Y1, (&) Y2 and (~) Y3. Reaction conditions: 573 K, 3.0 MPa H2, SV = 7.27 h1, H2/
naphthalene = 30 mol/mol.

their selectivity will depend on the level of conversion, which in


the present work was dependent upon the catalyst, the process
conditions and the time-on-stream because of the significant
deactivation observed for some of the catalysts. Hence, the
effect of Mo2C loading among the different catalysts, shown in
Fig. 10, is based on ROP yield, defined as the product of the
naphthalene conversion and ROP selectivity. Fig. 10 compares
the ROP yields among the different catalysts after 2 h time-onstream and 5 h time-on-stream. Overall, the optimum Mo2C
loading was dependent upon the support, with the optimum
loading generally increasing as the zeolite acidity increased,
assuming the trend shown by the limited data for Y3 can be
extrapolated to higher loadings. For the catalysts examined in
the present study, the highest ROP yield occurred for the
Mo2C(5)/Y1 and the Mo2C(7)/Y2 catalysts.
Fig. 11 shows the effect of reaction temperature and
residence time (1/SV in Fig. 11) on the naphthalene conversion
and product distribution over the Mo2C(5)/Y1 catalyst, after 5 h
time-on-stream. The changes in space velocity (SV) were a
result of changes in both liquid and gas (H2) flow such that the
H2:reactant ratio remained fixed. The data of Fig. 11 show that
at low temperature (553 K), the conversion increased with
increased residence time, whereas at 573 and 613 K, the
conversion reached a plateau at higher residence times. The
conversion data are clearly influenced by the equilibrium
conversion established between naphthalene and tetralin at
higher temperature [6]. The decrease in tetralin selectivity with
increased temperature and contact time, shown in Fig. 11,
confirmed this assertion. The ROP selectivity increased with
temperature and the effect of residence time was most
significant at low temperature. The increased ROP selectivity
with increased residence time observed at 553 K is in
agreement with ROP being generated by secondary reactions.
At higher temperature (>553 K), further reaction of ROP
occurred, decreasing selectivity to the desired product at higher
residence time. Tetralin selectivity showed a decreasing trend
with temperature and as the contact time increased, the initial
hydrogenation of naphthalene to tetralin also decreased. The
selectivity to heavy products also increased with residence time

for all temperatures, indicative of a terminal product from a


series of reactions. The above observations are consistent with a
reaction scheme, proposed in the literature [1], for naphthalene.
Accordingly, naphthalene undergoes hydrogenation as the first
step toward ring opening. A subsequent set of hydrogenation,
isomerization, hydrogenolysis and cracking reactions yield
ring-opening products, heavy products, cracked products and
coke. Note that although the rate of catalyst deactivation was
dependent upon the operating conditions (SV and temperature),
similar trends to those shown in Fig. 11 were observed after the
first 2 h time-on-stream. Consequently, the trends shown in
Fig. 11 do allow one to deduce the sequence of reactions that
leads to the various products, even though the conversion data
are marginally influenced by deactivation.
Finally, the thio-resistence of Mo2C(5)/Y2 catalyst was
investigated by adding 1000 ppmw dibenzothiophene (DBT) to
the naphthalene feed and the results are shown in Fig. 12. The
data show that the naphthalene conversion decreased with
addition of the DBT, whereas the ROP selectivity and tetralin
selectivity was not much affected after 5 h time-on-stream.
4. Discussion
Although Mo2C catalysts have been shown to have good
hydrogenation and hydrogenolysis activities [12], the present
study shows that their ability as bulk catalysts to hydrogenate
naphthalene is limited. Partial oxidation increased the
naphthalene hydrogenation rate (after 5 h time-on-stream),
but no ring-opening reactions occurred. The catalyst characterization data confirmed the presence of Mo2C and oxygen
treatment at elevated temperature all but removed the acidity.
The acidity of the bulk catalysts was an order of magnitude
lower than that of the HY zeolites and this apparently limits
acid catalysed ring opening via C6 ! C5 isomerization and
hence no ROP were observed with the bulk catalysts.
Mo2C supported on HY zeolites significantly increased both
the hydrogenation and ring opening of naphthalene compared
to the bulk materials. The presence of Mo2C in the supported
catalysts was confirmed by XPS, and the Mo2C particle size (as

S.J. Ardakani et al. / Applied Catalysis A: General 324 (2007) 919

17

Fig. 11. Effect of residence time (1/SV) on naphthalene conversion and product selectivities over the Mo2C(5)/Y1 catalyst at 5 h time-on-stream for (*) 553 K, (&)
573 K and (~) 613 K and 3.0 MPa H2, H2/naphthalene = 30 mol/mol.

determined by TEM) indicated that the crystallites were


supported on the HY zeolite rather than anchored within the
micro-pores of the zeolite. A similar conclusion has been
reported for Mo2C supported on ZSM-5 zeolite [20]. By using
HY as the support, the acidity of the catalysts was increased.
The HY zeolites alone converted naphthalene to heavy products
and tetralin, but the conversion decreased very rapidly, likely
due to coke formation that is a consequence of acid catalysed
polymerization reactions [1]. For all three HY zeolites,
naphthalene conversion, ring-opening selectivity and tetralin
selectivity increased significantly with the addition of the
Mo2C. Furthermore, there was a corresponding decrease in
selectivity to heavy products, consistent with the decrease in
acidity observed upon addition of Mo2C to the HY zeolites. The
comparison of data among the supports and the Mo2C catalysts,
however, clearly shows that both the Mo2C and the HY zeolite
are required to obtain significant catalyst activity and stability,
and that there is an optimum Mo loading that is dependent on
the HY zeolite. A moderate acidity provides the appropriate
compromise between too much cracking or polymerization but
sufficient ring opening to obtain a catalyst with reasonable ringopening selectivity. The initial yield of ROP obtained on
Mo2C(5)/Y1 and Mo2C(7)/Y2 is comparable to that reported by

others on noble metal catalysts supported on various acidic


supports [410].
The mechanism for naphthalene hydrogenation and ring
opening on bifunctional catalysts has been discussed in the
literature [1]. Hydrogenation can occur via conventional metal
catalysed hydrogenation or by acid site induced hydrogenation
involving migration of spillover hydrogen from metal sites. The
latter mechanism has been invoked to explain increased
hydrogenation activity of metal catalysts on acidic supports.
Subsequent reactions of tetralin include acid catalysed
isomerization and ring opening or metal catalysed hydrogenation or hydrogenolysis. In the present study, all of the HY
zeolites showed significant conversion of naphthalene, albeit
with rapid catalyst deactivation and loss in conversion. The
conversion was primarily to heavy products and tetralin, the
latter most likely a consequence of hydride transfer reactions
linked to the acid catalysed polymerization reactions that lead
to heavy products and coke deposits. In the presence of a
monofunctional acid catalyst and at typical cracking conditions
the direct attack of a Brnsted acid site is assumed to be the
initial activation step in the conversion of naphthalene [2123].
For the Mo2C(5)/Y1 catalyst, the change in tetralin
selectivity with residence time shown in the data of Fig. 11

18

S.J. Ardakani et al. / Applied Catalysis A: General 324 (2007) 919

Fig. 12. Naphthalene conversion and product selectivities over the Mo2C(5)/Y1 catalyst in the presence of DBT: (&) 0 ppm and (&) 1000 ppm DBT. Reaction
conditions: 573 K, 3.0 MPa H2, SV = 7.27 h1, H2/naphthalene = 30 mol/mol.

confirm that tetralin is a primary product and that ring-opening


products and heavy products are formed by secondary reaction.
At 573 K, the ROP selectivity decreased as residence time
increased, a consequence of conversion of the ROP to HP. The
product distribution obtained was consistent with a bifunctional
catalyst reaction mechanism proposed by others using metal
catalysts supported on acidic zeolites [1].
An interesting feature of the bifunctional catalysts is also
revealed by the deactivation profiles shown in Figs. 7 and 8. In
the case of HY zeolites, as the activity declined, the selectivity
to heavy products increased, whereas selectivity to tetralin and
ROP decreased with time-on-stream. In the case of the
Mo2C(i)/Zj catalysts, however, the rate of deactivation was far
lower than in the case of the corresponding HY zeolite, the
selectivity to tetralin increased and the heavy product
selectivity remained unchanged with time-on-stream. The data
suggest that in the presence of Mo2C, the hydrogenation of
naphthalene to tetralin is readily achieved (as reported for metal
catalysts) and this reaction is not much affected by coke
formation on acid sites. The tetralin can undergo acid catalysed
isomerization and ring opening or hydrogenation and hydro-

genolysis on the Mo2C. If deactivation is primarily associated


with the catalyst acid sites because of acid catalysed
polymerization reactions, then ROP selectivity would decrease
as the acid sites are deactivated and the selectivity to tetralin
would increase as less teralin were converted to ROP.
5. Conclusion
Naphthalene conversion over Mo2C supported on HY
zeolite catalysts is shown to have ROP yields comparable to
noble metal catalysts dispersed on acidic supports. The
catalysts are bifunctional with hydrogenation/hydrogenolysis
activity provided by the Mo2C and isomerization, ring opening,
cracking and polymerization occurring on acid sites associated
with the HY zeolite. The acid strength and Mo loading are both
contributing factors to the control of product selectivity, with
moderate acidity required to limit heavy product and coke
formation. Preliminary data also showed that the thio-tolerance
of these catalysts was significant, with the Mo2C(5)/Z2 catalyst
showing almost no effect in the presence of 1000 ppmw of DBT
after 5 h time-on-stream.

S.J. Ardakani et al. / Applied Catalysis A: General 324 (2007) 919

Acknowledgements
Funding for the present study from Natural Resources
Canada through the Climate Change Technology and Innovation Initiative is gratefully acknowledged. The authors also
acknowledge the support of Dr. Zbigniew Ring, CANMET
Energy Technology Center, Devon, Alberta.
References
[1] H. Du, C. Fairbridge, H. Yang, Z. Ring, Appl. Catal. A: Gen. 294 (2005) 1.
[2] J.P. van der Berg, J.P. Lucien, G. Germaine, G.L.B. Thielemans, Fuel
Process. Technol. 35 (1993) 119.
[3] A. Stanislaus, B.H. Cooper, Catal. Rev.-Sci. Eng. 36 (1994) 75.
[4] G.B. McVicker, M. Daage, M.S. Touvelle, C.W. Hudson, D.P. Klein, W.C.
Baird, B.R. Cook, J.G. Chen, S. Hantzer, D.E.W. Vaughn, E.S. Ellis, O.C.
Feeley, J. Catal. 210 (2002) 137.
[5] F. Locatelli, J.-P. Candy, B. Didillon, G.P. Niccolai, D. Unzo, J.-M. Basset,
JACS 123 (2001) 1658.
[6] M. Jacquin, D.J. Jones, J. Roziere, S. Albertazzi, A. Vaccari, M. Lenarda,
L. Storaro, R. Ganzerla, Appl. Catal. 251 (2003) 131.
[7] E. Rodriguez-Catellon, J. Merida-Robles, L. Diaz, P. Maireles-Torres, D.J.
Jones, J. Roziere, A. Jimenez-Lopez, Appl. Catal. 260 (2004) 9.

19

[8] U. Nylen, J.F. Delgado, S. Jaras, M. Boutonnet, Appl. Catal. 262 (2004)
189.
[9] M.A. Arribas, P. Concepcion, A. Martinez, Appl. Catal. 267 (2004) 111.
[10] S. Albertazzi, G. Busca, E. Finocchio, R. Glocker, A. Vaccari, J. Catal. 223
(2004) 372.
[11] M.K. Neylon, S. Choi, H. Kwon, K.E. Curry, L.T. Thompson, Appl. Catal.
A 183 (1999) 253.
[12] S.T. Oyama, The Chemistry of Transition Metal Carbides and Nitrides:
Preparation, Properties and Reactivity, Chapman & Hall, 1996, pp. 127
(Chapter 1).
[13] M.A. Arribas, A. Martinez, Appl. Catal. A 230 (2002) 203.
[14] C. Pham-Huu, M.J. Ledoux, J. Guille, J. Catal. 143 (1993) 249.
[15] F.G. Sheriff, US Patent 5,451,389 (1995).
[16] E.A. Blekkan, C. Pham-Huu, M.J. Ledoux, J. Guille, Ind. Eng. Chem. Res.
33 (1994) 1657.
[17] L. Leclercq, M. Provost, H. Pastor, J. Grimblot, A.M. Hardy, L. Gengembre, J. Catal. 117 (1989) 371.
[18] T.I. Koranyi, I. Manninger, Z. Paal, O. Marks, J.R. Gunter, J. Catal. 116
(1989) 422.
[19] S.F. Ho, S. Contarini, J.W. Rabalais, J. Phys. Chem. 91 (1987) 4779.
[20] A. Szechenyi, F. Solymosi, Appl. Catal. A: Gen. 306 (2006) 149.
[21] A. Corma, V. Gonzalez-Alfaro, A.V. Orchilles, J. Catal. 200 (2001) 34.
[22] A. Corma, F. Mocholi, V. Orchilles, G. Koermer, R.J. Madon, Appl. Catal.
67 (1991) 307.
[23] H.B. Mostad, T.U. Riis, O.H. Ellestad, Appl. Catal. 63 (1990) 345.

You might also like