You are on page 1of 21

Applied Catalysis A: General 294 (2005) 1–21

www.elsevier.com/locate/apcata

Review

The chemistry of selective ring-opening catalysts


Hongbin Du *, Craig Fairbridge, Hong Yang, Zbigniew Ring
National Center for Upgrading Technology, 1 Oil Patch Drive, Devon, Alta., Canada T9G 1A8
Received 6 June 2005; accepted 15 June 2005
Available online 30 August 2005

Abstract

Bitumen-derived crude and heavy oils require severe processing and produce middle distillate product with poor ignition quality. This
becomes a concern to refiners as tighter specifications on transportation fuel are promulgated. One process to address this issue is selective ring
opening of cycloparaffins to reduce the number of ring structures, while retaining the carbon number of a product molecule. The process involves
bifunctional catalysts, both metal and acid sites, working in high-pressure, high-temperature reactor systems in the presence of hydrogen. The
acidic sites catalyze dehydrogenation, cracking, isomerization and dealkylation, while the metal sites promote hydrogenation, hydrogenolysis
and isomerization. Various compounds containing single, double and multiple rings have been used to model the ring-opening reactions and a
number of mechanisms have been proposed. The five-membered ring readily undergoes ring-opening reaction on either acid or metal catalysts
with the selectivity and activity dependent on the nature of the supported metal catalyst. The ring opening of six-membered ring compounds is
secondary, requiring an acidic function to isomerize a six-membered ring cycloparaffin to a five-membered ring. A balanced metallic-acidic
function catalyst is necessary to achieve optimal performance. A system dominated by acid function results in excess cracking, while a catalytic
system with high concentration of metals leads to mainly hydrogenation. Commercial hydrocracking catalysts using transition metal sulfides on
acidic supports usually require severe operating conditions due to their low activities of the metal sulfide compared to metal sites, leading to
extensive cracking of cycloparaffin side chains. Noble metals supported on acidic oxides are the most active catalysts for selective ring opening,
but these catalysts are very sensitive to poisoning by sulfur compounds in petroleum feedstocks. An understanding of the chemistry of selective
ring-opening catalysts, combined with theoretical studies of structure–activity relationships and high throughput experimentation methods,
provides opportunities in searching for new generations of selective ring-opening catalysts with high-performance and sulfur resistance.
Crown Copyright # 2005 Published by Elsevier B.V. All rights reserved.

Keywords: Ring opening; Catalysts; Model compound; Mechanism; Bifunctional catalysts

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Ring-opening catalysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1. Cracking on solid acid catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2. Hydrocracking on bifunctional catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.3. Hydrogen spillover . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3. Selective ring opening of single-ring compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.1. Activity and selectivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.2. Proposed mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.3. Electronic effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.4. Geometric effect. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.5. Support effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

* Corresponding author. Present address: State Key Laboratory of Coordination Chemistry, Nanjing University, Nanjing 210093, China.
Fax: +86 25 8331 4502.
E-mail address: hong_bin_du@hotmail.com (H. Du).

0926-860X/$ – see front matter. Crown Copyright # 2005 Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2005.06.033
2 H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21

4. Selective ring opening of multiple-ring compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10


4.1. Indan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.2. Decalin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.3. Tetralin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
4.4. Naphthalene. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.5. Phenanthrene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.6. Fluorene and fluoranthene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
5. Sulfur tolerance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
6. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

1. Introduction In Europe, the European Commission mandated a


reduction in diesel sulfur levels from 500 to 350 wt ppm in
Health concerns and local environmental awareness have 2000, and a further reduction to 50 wt ppm in 2005. The
prompted worldwide regulatory actions on transportation regulations require higher average cetane number (currently
systems to limit emissions of criteria air contaminants 51 and 53 by 2005) than in the US and an even higher level of
(hydrocarbons, carbon monoxide, nitrogen oxides and 55 is under consideration and will put additional pressure on
particulate matter). In addition, concerns for global climate refiners. For gasoline, the regulations require reduction of
change have highlighted the need to improve transportation sulfur content from 150 to 50 ppm, and aromatics from 45 to
system efficiency to decrease emissions of carbon dioxide. 35 vol% by 2005. Future regulations will likely tighten the
These regulations and concerns call for new technologies to limitation for the aromatic and toxics content and increase
produce high quality liquid fuels and internal combustion fuel efficiency to reduce the resultant vehicle CO2 emissions.
engines with maximum combustion effectiveness and This will probably promote light-duty, diesel-fuelled vehicle
minimum emissions. The regulations primarily involve production and the use of ultra-low sulfur diesel (ULSD).
reductions of sulfur levels, aromatics (particularly benzene), In the last few years, refiners in the US and Canada have
olefins, volatility and the use of oxygenates in gasoline. The invested in major refining expansion and revamping
sulfur compounds in fuels contribute to the emission of construction projects because of the ultra-low sulfur fuel
sulfates with particulate matter, and have a deleterious effect requirements. At the same time, they are also facing
on the ability of future automobile engines to meet the more challenges of meeting other requirements including the
stringent emissions standards. Aromatics in middle-dis- replacement of methyl tertiary butyl ether (MTBE) and
tillate fuels produce particulates in the exhaust gases and, in reduction of aromatics [1–3]. In addition, refiners have to cope
addition, have poor ignition properties, i.e. low cetane with increasingly heavier and poorer quality feedstocks.
number in diesel fuel and high smoke point in jet fuel. According to the US Energy Information Administration, the
Across the world, gasoline and diesel specifications are quality of crudes processed in the US declined between 1997
becoming stricter, especially with regard to sulfur. In the US, and 2001 at the approximate rates of 0.128 API/year and
the Reformulated Gasoline Program Phase II, which took +0.057 wt% sulfur/year. This trend is expected to continue in
effect on 1 January 2002, mandated a reduction in volatile the next 5–10 years, forcing US refineries to import heavier
organic compounds (VOC) by 27%, nitrogen oxides (NOx) crudes due to the worldwide increase in crude demand,
by 7%, and toxic pollutants (aromatics, especially benzene) decreased production of light crudes from mature regions such
by 22%. In 2004, the Environment Protection Agency as the US and North Sea, and geopolitical factors. Increased
started to implement the Tier 2 Gasoline Sulfur Control heavy oil supplies arose from Canada and Venezuela.
Program for cleaner gasoline, which limits the maximum With 174 billion barrels of established bitumen reserves
sulfur level to 300 ppm, and by January 2006 to 80 ppm. The (recoverable in situ and mineable), the Canadian oil sands
current maximum concentration of sulfur in on-road diesel is resource has been increasingly recognized as a strategic
limited to 500 ppm and 2000–5000 ppm for off-road diesel. source of North America energy supply. Alberta bitumen
By June 2006, the new EPA regulations will limit sulfur production reached approximately 1 million b/d (barrels per
content to 15 ppm. Additional specifications for minimum day) in 2003, accounting for 53% of Alberta’s total crude oil
cetane index at 40, and maximum aromatics content at and equivalent production. According to the Alberta Energy
35 vol% will also be imposed. Utilities Board (AEUB), total bitumen production will
Canada has also launched programs for a staged sulfur increase to 3.5 million b/d by 2017, representing 2 million b/
reductions in gasoline that introduced the specification of d of synthetic crude oil (SCO) and 1.5 million b/d of
150 ppm July 1, 2002, then to 30 ppm by January 1, 2005. A unprocessed crude bitumen. Compared to conventional
maximum of 500 ppm sulfur in on-road diesel fuel was crude oil, SCO has the valued quality attributes of low sulfur
mandated on January 1, 1998, with 15 ppm slated for June content and zero residues. However, SCO also has some
2006. major disadvantages, largely related to its high aromaticity.
H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21 3

The heavy gas oil (HGO) bottom cut from SCO (about
38 vol%) contains more than 90% of cyclic hydrocarbons
which causes processing problems, particularly in fluid
catalytic cracking (FCC) units, which constitutes approxi-
mately 80% of the US conversion capacity. This has led to its
limited acceptance in a conventional FCC refinery intake. In
addition to the poor HGO quality, SCO also yields poor
quality middle distillate (potentially jet and diesel fuel
blending components). These factors have resulted in a price
differential between SCO and other light crudes. Similar Fig. 1. Cetane numbers of various compounds [76].
problems exist with other highly aromatic streams, such as
those from cokers and FCC units. These potential motor fuel bond scission in the presence of hydrogen [10,11]. However,
components have low cetane numbers and poor ignition because of its extensive overcracking there is limited net
qualities. This will become a particular difficulty for refiners gain in production of desirable paraffins via this route, and
if higher cetane specifications are introduced worldwide in the improved distillate fuel quality results primarily from a
the near future. In order to accept larger volumes of bitumen- combination of hydrogenation of aromatics and a concentra-
derived crudes, both the producers and refiners will have to tion of paraffins in a reduced volume of distillate product. The
address the quality issues and refining challenges. One way ideal process would be a selective ring-opening catalyst that
to do this is to develop new catalytic processes that will can reduce the number of ring structures while retaining the
improve the quality of these streams. There has been a carbon number of a product molecule.
growing demand for refining catalysts driven primarily by
refiners’ needs to meet the already legislated regulatory
requirements without capital infusion. It will require new 2. Ring-opening catalysis
generation of high-performance hydroprocessing catalysts
to make higher quality fuels of the future. 2.1. Cracking on solid acid catalysts
Hydrogenation and hydrocracking are two commercially
proven refining technologies addressing the high aromaticity Selective ring opening of naphthenes with minimum
issues [4]. The conventional hydrotreating catalysts help cleavage of side chains is very complex and represents a
saturate aromatic rings to naphthenes (cycloparaffins or challenge to catalyst researchers. It is known that ring
cycloalkanes) [5,6]. However, the increase of cetane number opening can be catalyzed by the acid sites, e.g. the Brönsted
via hydrogenation of aromatics is insufficient to significantly sites, via carbenium intermediates [12]. The reaction is
increase the quality of the middle distillates because the cetane initiated by protolytic cracking, accompanied by protolytic
numbers of individual cycloalkanes, such as decalin are rather dehydrogenation, hydride transfer, skeletal isomerization,
low (Fig. 1). Early studies by Wilson et al. demonstrated that b-scission and alkylation, as in the cases of cracking of
the hydrogenation or a combination of hydrogenation and aliphatics over acid catalysts. The latter mechanisms have
aromatics extraction of bitumen-derived middle distillates been extensively studied [13]. The cracking of endocyclic
could only produce a distillate with 40 cetane number [7–9]. C–C bonds in cyclic hydrocarbons, however, is much slower
However, approximately 97% of the aromatic carbon had to be than those of aliphatics, presumably because of a higher
converted by severe hydroprocessing to achieve this. tendency of the alkenyl cation, formed by b-scission of a
Hydrocracking is, on the other hand, a process designed to cycloalkyl cation, to recyclize, or because of a lower b-
produce lighter and higher quality naphtha and middle scission rate in cylcoalkylcarbon ions caused by an
distillate fuel blending stocks by saturating aromatic rings and unfavorable orientation of the p-orbital at the positively
reducing the number of ring structures via carbon–carbon charged carbon atom and the b-bond to be broken (Fig. 2)

Fig. 2. b-Scission of linear and cyclic hydrocarbons [14].


4 H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21

[14]. As a result, the overall reaction is predominated by


isomerization and subsequent hydrocracking (b-scission) of
(2)
side chains of cyclic hydrocarbons, particularly of those
having substituents with more than five carbon atoms,
leading to significant dealkylation of pendant substituents on
the ring. The yields of desired high cetane ring-opened
products are usually low due to a high extent of consecutive
cracking reactions and a fast catalyst deactivation. (3)

2.2. Hydrocracking on bifunctional catalysts

Starting in the 1950s, bifunctional catalysts consisting of 2.3. Hydrogen spillover


highly dispersed metal particles for hydrogenation or
dehydrogenation and an acidic support for cracking or The classical model for bifunctional catalysis proposed
isomerization were introduced in refining processes, such as by Mills et al. [15] and by Weisz and Swegler [16] envisaged
catalytic reforming of naphtha, hydrotreating, hydrocrack- the reaction in three steps (Eqs. (1)–(3)), involving a gas
ing and hydroisomerization [2]. These processes improved phase diffusion of olefinic intermediates between the two
fuel quality and at the same time avoided catalyst decay and catalytic sites. The model successfully explained a number
the formation of coke deposits. Early studies by Mills et al. of experimental observations, but failed to account for the
[15] and by Weisz and Swegler [16] described the reactions role of hydrogen and the synergy between the two catalyst
over the bifunctional catalysts as proceeding over two components in controlling the selectivity and activity [17].
distinct catalytic sites: the reactants are first converted into The extension of the classical model by incorporating the
olefins on the metal site via hydrogenation/dehydrogenation hydrogen spillover concept allows a better interpretation of
reactions (Eq. (1)); then the formed olefins are protonated at experimental results, including those that did not fit into the
the acid sites, leading to the formation of carbenium ions, old theorem. The new model involves the formation of
which subsequently undergo skeletal isomerization, crack- mobile hydrogen surface species, i.e. spilt-over hydrogen
ing or alkylation (Eq. (2)). The products are finally desorbed that allows the hydroconversion of the hydrocarbon at a
from the acid sites as olefins, which are hydrogenated on the certain distance from the metal so that all reaction steps can
metal sites in the presence of hydrogen (Eq. (3)). It was later occur on one reaction site.
recognized that the reactions could also occur on one Spillover is now a well-known phenomenon in hetero-
reaction site, owing to activated hydrogen species, i.e. geneous catalysis, involving the transport of active species
spillover hydrogen [17]. The metal co-catalyst provides sorbed or formed on one surface onto another that does not
spillover hydrogen, which migrates to the acid sites and sorb or form the active species under the same conditions.
saturates the carbenium intermediates. In other words, the Several small molecules have been known to exhibit
acidic support cannot only initiate the formation of spillover effects upon interactions with noble metals,
carbenium ions pyrolytically or by the addition of protons including hydrogen, oxygen, nitrogen, carbon monoxide
to olefins, but also hydrogenate carbenium ions with hydride and organic species [18]. Hydrogen spillover plays an
ions and promote their desorption as the saturated products. important role in petroleum processes, e.g. in hydrotreating,
In addition, the ring opening of naphthenes can also proceed hydrocracking, hydrogenation and hydroisomerization.
on certain metal sites via direct hydrogenolysis of an A number of excellent reviews on this subject have been
endocyclic C–C bond [12], i.e. cleavage of a C–C bond with published [17–21].
the addition of hydrogen. Overall, the activity and selectivity The nature of the spilt-over hydrogen species has been
of ring opening on bifunctional catalysts are strongly discussed in the literature. Depending on the system studied,
dependent on the properties of the metal and support as well different species have been claimed, including H atoms, H
as on reaction conditions. These parameters include the type ions, ion pairs, H3 species and bound species. Roland et al.
of metal, metal particle size, acidity of the support, pore size [18] proposed a model that describes the spilt-over species
of the support, the interface length and strength balance as electron donors adsorbed on the surface, which
between the metal and acid sites, working conditions, such corresponds to H atoms (occupied, weakly chemisorbed)
as temperature, hydrogen pressure, etc. This has provided or H+ ions (empty, strongly chemisorbed). Their ratio is
opportunities to devise catalysts for highly selective ring determined by the chemical properties (e.g. the presence of
opening of naphthenic compounds in fuel without significant Lewis and Brönsted acid sites) and the electronic properties
dealkylation of any pendant substituents on the ring. (e.g. electron density) of the solid. In bifunctional catalysts,
spilt-over hydrogen can donate an electron to the support to
form a proton on a Brönsted site, and a hydride ion may form
(1) on a Lewis site in the process of charge balancing [13]. An
olefin formed on the metal by dehydrogenation can react
H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21 5

with the proton on a Brönsted site to form a carbenium ion, supplies spilt-over hydrogen to acidic sites, where it forms
which may then combine with the migrated hydride ion from acidic hydroxyl groups and promotes acid-catalyzed
a Lewis site to produce a saturated compound. Alternatively, reactions [21].
the carbenium ion formed by addition of a proton to an olefin
on a Brönsted site can migrate to a Lewis site where it reacts
with a hydride ion. 3. Selective ring opening of single-ring compounds
The creation of catalytic active sites by spilt-over
hydrogen has been recognized and used in supported metal 3.1. Activity and selectivity
catalysts for hydrogenation (or dehydrogenation), hydro-
cracking and hydroisomerization. For instance, spillover of When adsorbed on metal surface at certain conditions,
hydrogen from a metal under certain conditions can make hydrocarbon molecules undergo dehydrogenation/hydroge-
the inert silica, or alumina, or carbon support active for nation, skeletal isomerization and hydrogenolysis, including
hydrogenation of olefin or benzene [22–27]. A number of ring opening/cyclization, and ring contraction/enlargement.
studies showed that mechanical mixtures of supported metal Certain noble metals, such as Pt, Pd, Ir, Ru and Rh have been
catalyst and a support exhibited higher hydrogenation found to be selectively active for the ring opening for cyclic
activities than the supported metal catalyst alone. The hydrocarbons to the corresponding paraffins with the same
increased activities have been attributed to hydrogen carbon number. The activity and selectivity depend mainly
spillover. Without activation by hydrogen spillover, neither on the metal catalysts, such as the type of metal, particle
silica nor alumina will adsorb hydrogen and both are inert size, crystal morphology, etc. The ring opening of cyclic
for hydrogenation catalysis. hydrocarbons on noble metals is highly sensitive to the
Several studies have implied that hydrogen spillover can catalyst structure, particularly alkylcyclopentanes and
give rise to Brönsted acid sites on oxide supports, including alkylcyclobutanes, which have been used as molecular
zeolites. Fujimoto [26] found that mechanically mixed probes to characterize various types of catalysts. Several
catalyst Pt/SiO2 + H-ZSM-5 showed both high selectivity excellent reviews have been published on this subject
and high conversion for benzene hydrogenation, equal to [14,28–33].
those obtained over a Pt/ZSM-5 catalyst, while Pt/SiO2 or The ring opening of methylcyclopentane (MCP) on
H-ZSM-5 alone were not effective catalysts. He attributed supported metal catalysts has been extensively studied
the activity of the mechanical mixture to the generation of [30,34–37]. The reaction produces n-hexane (nHx), 2-
Brönsted acid sites by hydrogen spillover. Ohgoshi et al. methylpentane (2MP) and 3-methylpentane (3MP). The
[27] demonstrated that the hybrid catalysts Pt/KA + NaY or product distribution is dependent on the properties of the
H-ZSM-5 showed high activity for isobutene hydrogena- metal. Over supported Pt catalysts with small particle sizes,
tion, while the Pt/KA catalyst and zeolite Y or H-ZSM-5 e.g. low loading and highly dispersed Pt, the ring opening of
alone failed to hydrogenate isobutene because isobutene is MCP is non-selective (the rupture of endocyclic C–C bonds
too big to fit into the pores of KA zeolite (3 Å), and H2 is statistical, producing 2MP, 3MP and nHx in a ratio of
cannot be activated on NaY or H-ZSM-5 zeolite. They 2MP:3MP:nHx = 40:20:40 related to the number of bond
concluded that H2 was activated on Pt/KA. The resulting H types in the molecule) (Fig. 3). On the other hand, the ring
atoms spilt-over onto the NaYor H-ZSM-5 zeolite, and then opening of MCP on large metal particles and metal surfaces
hydrogenated isobutene on the zeolite. Further study using a with low Miller index produces selectively 2MP and 3MP,
H-ZSM-5 zeolite that contained almost no SiOH groups but no nHx. In some cases, the ring opening does not follow
mixed with Pt/KA yielded zero conversion of isobutene, the ‘selective’ or ‘non-selective’ mechanism, instead
indicating the OH groups play an important role for producing unusually high amount of nHx or 3MP relative
migration of spilt-over hydrogen. In the bifunctional CoMo/ to its statistical ratio. These ‘partially selective’ mechanisms
zeolite hydrocracking catalysts, Co-Mo adsorbs H2 and compete with the non-selective and selective mechanism,

Fig. 3. Non-selective and selective hydrogenolysis of methylcyclopentane.


6 H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21

occurring on high loaded 10% Pt/Al2O3 at high-temperature adsorbed intermediates at different hydrogen pressures. A
[30], or Pt/zeolite [38–40]. dissociative adsorption of hydrocarbon via the rupture of the
The ring opening of MCP on Ir [12] and Rh [41], on the C–H bond(s) was assumed at low hydrogen pressure. An
other hand, is less sensitive to the particle size of the metals increase in hydrogen pressure inhibits this C–H dissociation,
in comparison with Pt catalysts. However, these metals slowing down the reaction rate. In the case of Rh catalyst,
exhibit higher activity and selectivity to open the C5 ring in new adsorbed species were presumably formed via scission
bisecondary positions [36,42]. Other metal catalysts, such as of the ring C–C bond at high hydrogen pressure. These
Co, Ni, Ru and Os, show extensive hydrogenolysis, yielding species were desorbed to produce ring-opening compounds.
a significant amount of fragments [41,42]. Compared to the five-membered ring cyclopentanes, the
Generally, the ring-opening activities of alkylcyclopen- ring opening of the four-, and six-membered ring
tanes on metal catalysts decrease with increased number of compounds are less extensively studied. Cyclobutanes are
ring substitutents [34]. This is especially the case for Ir, and more reactive towards ring opening than cyclopentanes due
to lesser extent Ru, Rh and Ni [36], all of which show a to higher ring strain. The reactions can take place at low
preference for cleaving unsubstituted ring C–C bonds. A temperature (<100 8C), exhibiting relatively low selectivity,
significant decrease in activity was observed for cleaving i.e. statistical product distribution [44,46]. On the other
substituted ring C–C bonds over these metals. In compar- hand, the ring opening of cyclohexanes over metal catalysts
ison, Pt is better able to break substituted ring C–C bonds [36,48,49] is much slower than five-membered rings, even
than Ir, Ru, Rh and Ni. However, the ring-opening rate over over the most active metal Ir [36]. Instead, cracking and
Pt catalysts is sensitive to the cis-/trans-isomer ratio, aromatization dominate the total reaction. For a significant
decreasing with increasing trans-isomer concentration. In conversion of six-membered rings into alkanes without loss
other words, Pt requires a cis-substitution of isomers when of molecular weight, additional acid function is required to
breaking a tertiary–tertiary C–C ring bond. convert the six-membered rings to the five-membered rings
The reaction conditions have also been found to affect the via skeletal isomerization. The latter then readily undergo
product distribution. Changes in selectivity and activity of selective ring opening on acid or metal sites.
MCP ring opening on supported Pt metal catalysts with
hydrogen pressure have been reported [34,43]. Higher 3.2. Proposed mechanism
hydrogen pressures favor ring opening further from the
substituents. The increase in hydrogen partial pressure leads There have been several mechanisms proposed to account
to increased ring-opening activity on supported Pt catalysts at for ring opening on metal surfaces, two of which are well
relatively low temperatures (<300 8C). This reaches a recognized: the multiplet mechanism and the dicarbene
maximum and thereafter shows a steady decline, or a very mechanism [30,33]. The two mechanisms differ mainly in
broad plateau with further increase in hydrogen partial how the reaction intermediate forms on the metal surface.
pressure. The hydrogen pressure at which the maximum In the multiplet mechanism, cyclic hydrocarbons
activity is obtained is a function of the support. However, physically adsorb either edgewise on two metal atoms, or
these volcano-type plots of ring-opening activity versus flat-lying on the metal surface. The former is called the
hydrogen pressure are independent of the support and the ‘doublet’ mechanism, and the latter called the ‘sextet-
reduction temperature. The latter determines the particle size, doublet’ mechanism. Both mechanisms compete with each
morphology and structure of the metal particle on the support, other in the reaction. The doublet mechanism is thought to
and influences the activity. These results were attributed to the occur on small metal particles, and is used to explain the
competitive adsorption of hydrogen and hydrocarbon on the selective hydrogenolysis of bisecondary C–C bonds.
metal surface. Increase in hydrogen pressure diminishes According to this mechanism, cyclic hydrocarbon is
dehydrogenation activity, leading to an increase in the overall adsorbed perpendicular to the metal surface via a
selectivity towards the ring-opening reaction. bisecondary C–C bond, which then reacts with chemisorbed
Hydrogen pressure dependence on activity and selectivity hydrogen in a push–pull manner to produce ring-opening
of ring opening has also been observed over other transition products. Due to steric hindrance, the edge-wise adsorption
metal catalysts [41,42,44–47]. The relationships between of the tertiary–secondary or tertiary–tertiary C–C bonds is
the product formation rate and hydrogen pressure varies with limited. In the ‘sextet-doublet’ mechanism, on the other
catalytic systems, reflecting differences in reaction mechan- hand, the cyclic molecule is physically adsorbed flat-lying
ism. Over Rh catalysts [45,46], the ring opening of 1,2- on the metal surface, with the carbon atoms of the ring
dimethylcyclopropane goes through a maximum, and then a located over the interstices of the metal plane, e.g. Pt(1 1 1).
minimum and then a continuous increase along with the For the five-membered ring cyclopentanes, one C–C bond
increase in hydrogen pressure. On Pd, Ni and Cu catalysts has to be stretched (Fig. 4), and this bond is readily attacked
[45], the formation rates of one product increases and levels by neighboring, adsorbed hydrogen, leading to hydrogeno-
off with increasing hydrogen pressure, while for the other lysis of the ring. The tertiary–secondary C–C bonds in
product the curve passes through a maximum. These alkylcyclopentane could be ruptured via the five-atom
observations were attributed to the changes of formation of adsorbed mode. For cyclobutanes, all four C–C bonds are
H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21 7

Fig. 4. Hydrogenolysis of methylcyclopentane via the multiplet mechanism [30].

stretched, resulting in higher reactivity but lower selectivity. 3.3. Electronic effect
In contrast, there is no stretching when cyclohexanes and
paraffin adsorb on the metal surface, as all the carbon atoms As various mechanisms were proposed to account for
can fit the interstices of the Pt(1 1 1) plane. As a result, experimental facts such as selectivity, structural effect and
cyclohexanes and paraffin are usually not hydrogenolysized, kinetic data, the nature of the catalyst–reactant interactions
but preferably dehydrogenated or isomerized. remains unknown. In ring opening (cyclization), and
Alternatively, the dicarbene mechanism involves the isomerization catalysis, metallocyclobutanes, metallocar-
chemisorption of the cyclic hydrocarbon molecules on the benes and metallocarbynes were postulated as the possible
metal surface after the rupture of several C–H bonds, forming reaction intermediates, depending on the nature of the sites.
carbon–metal bonds or p-adsorbed olefins (Fig. 5). The non- To generalize and simplify the reaction mechanisms, Garin
selective ring opening of MCP on metal catalysts can be and Maire [29] proposed an agostic precursor as the first
accounted for by the p-adsorbed olefin mechanism. Similar to species adsorbed on the metal surface, which consists of an
that in the sextet-doublet mechanism, MCP is adsorbed M H–C with a hydrogen atom bonded simultaneously to
parallel to the metal surface, but in a quasi-planar manner both a carbon atom of a reactant and a transition metal atom
involving only one metal atom. The selective ring opening of (Fig. 6). This species initiates the formation of a s-alkyl or
MCP, on the other hand, involves 1,2-dicarbene complexes carbene species that further reacts to give rise to products of
that bond to several metal atoms and stand perpendicular to ring opening (cyclization), hydrogenation (dehydrogena-
the metal surface. Owing to steric hindrance, the hydro- tion) and/or isomerization. It was proposed that an
genolysis of the tertiary–secondary C–C bonds is retarded. On electronic (or ligand) effect governs the relative contribu-
some occasions, however, an exocyclic alkyl substituent tion of s-alkyl or carbene precursor, while a geometric (or
participates in the formation of a metallocyclobutane ensemble size) effect determines the pathway to the bond
intermediate, resulting in the selective breaking of tertiary– shift (isomerization), the cyclic, and the hydrogenolysis
tertiary and tertiary–secondary C–C bonds (Fig. 5b). reactions.

Fig. 5. Hydrogenolysis of methylcyclopentane via the dicarbene mechanism [30].


8 H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21

Based on the product distribution, hydrogenolysis


reaction on metal catalysts can be classified into two groups
[29]: the C2-unit mode, involving rupture of C–C bonds
between primary and secondary carbon atoms; and the iso-
unit mode, involving rupture of a C–C bond with a tertiary
carbon atom. In the C2-unit mode, metallocarbene species
may be formed upon removing at least two hydrogen atoms,
which lead to hydrogenolysis over isomerization. On the
other hand, in the iso-unit mode where only one H atom is
available on the tertiary carbon, s-alkyl species are formed,
leading to higher isomerization than hydrogenolysis. The
relative contribution of the two modes is determined by the
electronic effect, i.e. the nature of the metal catalyst.
The C2-unit mode is favored on monometallic Ru, Ir and
bimetallic Pt-Ru, Pt-Co, Pt-Ir and Pt-Mo catalysts. The C2-
unit mode-dominated reactions, e.g. catalyzed by Ir catalysts,
were thought to proceed via a dicarbene (or doublet)
mechanism. The reactant bonds perpendicularly to the metal
surface. The reactions are usually not sensitive to the particle
size of metals but to the number of substituents because of
Fig. 6. Proposed reaction mechanism of n-C5H12 on a metal catalyst, steric hindrance. On the other hand, the iso-unit mode is
involving an agostic precursor species [29]. dominant on Ni, Pt, Pd, Pt-Ni, Pt-Wand Pt-Pd catalysts. In this
case, the reactant is adsorbed parallel to the metal surface, and
the reaction proceeds via a multiplet mechanism. The particle
In hydrocracking/isomerization of 2-methyl- and 3- size of metal influences the selectivity of the reactions.
methylpentane on several group VIII metal catalysts [29], It is now widely accepted that the reactions of
studies have shown that Co and Ni tend to cleave multiple hydrocarbon on metal are initiated by dissociative adsorp-
bonds, leading to extensive cracking, while Pd, Pt and Ir tion of the alkanes on the metal surface and followed by
selectively rupture primary–secondary/tertiary, secondary– subsequent hydride elimination from adsorbed alkyl
secondary/tertiary (demethylation) and secondary–second- intermediates. The former is the rate-determining step in
ary C–C bonds, respectively. The differences in the the reaction, while the latter is thought to determine the
selectivity and activity of metals in hydrogenolysis and selectivity. Zaera [53–55] proposed that b-hydride elimina-
isomerization may be related to their electronic structure tion from the surface intermediate accounts for the
(density of states), as proposed by Saillard and Hoffmann production of olefins, g-hydride elimination is responsible
[50]. When a H2 or a hydrocarbon molecule (e.g. CH4) for isomerization and cyclization, and a-hydride elimination
approaches a metal surface, the electron transfer from d leads to hydrogenolysis products (Fig. 7). The relative rates
orbitals of the metal M to a C–H s* antibond (M ! s*) for a-, b- and g-dehydrogenations determine selectivity of
dominates the early stages of the reaction, which weakens metal catalysts in hydrocarbon reforming. In general, the
the C–H s bond and forms the M–H bond. This is different reactivity for a-, b- and g-dehydrogenations increases with
from transition metal complexes in which s ! M electron early transition metals and with decreasing hydrocarbon
transfer leads the reaction because of the higher energy of
the occupied metal orbitals on the surface. For transition
metals at the left side of the periodic table, the surface is
positive relative to the bulk, while for metals at the right side
of the periodic table, the metal surface is negative. This is in
accordance with the fact that the heats of chemisorption of
hydrogen on metal surfaces decrease from left to right across
a periodic row [50,51]. These metals are more likely to form
metallocarbynes or carbenes, leading to multiple C–C
rupture. Within the same series of transition metals, e.g.
group VIII metal, the charge transfer from the bulk to the
surface decreases as d level occupancy for a metal surface
increases in traversing a periodic row [50,52]. Consistently,
the heats of chemisorption for hydrogen atoms on the metal
surface decrease in going from 3d to 4d to 5d (i.e. decrease Fig. 7. Simplified reaction mechanism for the catalytic reforming of
with increased atomic weight). hydrocarbons on transition metal surfaces [54].
H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21 9

chain lengths. The b-hydride elimination is the most icosahedron (20 faces) might become more stable [31]. As
favorable. However, the relative rates for a-, b- and g- the particle size decreases, the number of atoms of low
dehydrogenations within the metal also change across the coordination at the edges or corners of the crystallites
periodic table, which results in the different catalytic increases, while the fraction of face atoms diminishes.
performance of different metals in hydrocarbon reforming. Therefore, small particles would favor catalysis by atoms of
For instance, Pt(1 1 1) displays comparable rates for a- and low coordination, e.g. the non-selective, statistical hydro-
g-dehydrogenations, while Ni(1 0 0) only shows a-dehy- genolysis, giving higher rates. Large particles would favor
drogenation. These results explain the unique ability of Pt catalysis by face atoms of high coordination numbers, e.g.
for catalyzing hydrocarbon reforming as opposed to Ni, selective hydrogenolysis. The effect of surface atom
which facilitates extensive cracking. coordination on the catalytic properties of the metal can
also account for the changes in the selectivity of the ring
3.4. Geometric effect opening of MCP over different metal surfaces. In controlled
tests using metal films with oriented faces [58,59] or
The geometric effect on the catalytic properties is supported metal catalysts where the specific facets of the
exemplified by the particle size effects, which have been metal particles were selectively prepared [41,60], studies
extensively studied [28,30,31]. In hydroconversion of showed that the hydrogenolysis activity and product
hydrocarbons, large metal particle size promotes selective selectivity varied from facet to facet of metal crystal. High
hydrogenolysis, and small particle size favors non-selective activities and non-selective hydrogenolysis were observed
isomerization and hydrogenolysis. For instance, ring open- on small polycrystalline particles with disordered surfaces,
ing on supported Pt metal catalysts exhibit pronounced or on flat, stepped, and kinked noble-metal single-crystal
particle size effects on the selectivity. As illustrated by the surfaces of high Miller index containing high concentrations
hydrogenolysis of MCP [56,57], catalysts with large Pt of low-coordination sites. This particle size phenomenon
particle size lead to selective ring opening, favoring 2MP, observed in heterogeneous catalysis is the basis of the
and 3MP with the formation of nHx suppressed. Small Pt current interest in nanotechnology related to chemistry and
particle size produces non-selectively nHx, 2MP and 3MP in materials.
a statistical ratio. Other metals, e.g. Rh and Ir, favor the
selective hydrogenolysis with little effect of particle size. 3.5. Support effect
The particle size has also affected the turnover rate
(activity) in the hydrogenolysis of cyclic hydrocarbons Besides the above-mentioned electronic and geometric
[30,31]. On Pt/Al2O3, the turnover rate of the hydrogeno- effects, catalyst supports also play an important role in
lysis of cyclopentane was shown to decrease with decreased heterogeneous catalysis. As demonstrated in many catalytic
particle size. Similar results were observed on Rh/Al2O3. On processes, the supports are not inert in selective ring-
Ir/Al2O3, however, the hydrogenolysis of cyclopentane is opening catalysis as previously thought, but alter the
insensitive to the particle size. Pd/Al2O3, on the other hand, catalytic properties of the metal particles through electronic
presented a slight increase in the hydrogenolysis of interactions, or by space confinements. The support effect
cyclopentane with decreasing particle size. has attracted much attention and stimulated extensive
The origin of particle size effects is not clear. Some research worldwide. Excellent reviews have been published
possible explanations include electronic effects, morphol- on this subject [32,33,61,62].
ogy, and support effects [28,30,31]. As the particle size is A ‘partly’ selective mechanism for ring opening of MCP
reduced, the electronic bands of the metal particle become was observed on noble metals supported on acidic oxides,
distinct, and the electron energies increase. For most metal which produced a higher proportion of nHx than the
particles, the electron energies begin to increase as the theoretical ratio, according to a non-selective mechanism.
particle size is reduced below 5 nm. For some metals, the The formation of nHx appeared to increase with increasing
energies level off below a certain size, e.g. Pd at about 1 nm, support acidity [33,63,64]. Kramer and Zuegg [57] used the
while for others, e.g. Pt, the increase in energies continue adlineation model, coined by Schwab and Pietsch [65], to
with the decrease in particle size. As a result, the electronic account for the partial selectivity for nHx and 2MP. The
configurations of the surface atoms are different from those model assumed that ring opening occurs at the phase
of large particles, leading to changes in the catalytic boundary of metal and support. The support, i.e. an acidic
properties. On the other hand, the decrease in particle size site, interacts with the tertiary C–H of MCP because this H
likely changes the morphology of the metal particle, which atom is most susceptible to attack by an acidic or cationic
could influence the catalytic properties. For instance, a large site, while the neighboring carbon is attached to the metal
fcc metal particle crystallizes in an octahedron geometry (8 site. Ring opening of MCP adjacent to the methyl group
faces), while most small fcc metal particles (d = 10 nm) occurs, leading to the formation of additional nHx (Fig. 8).
adopt cubooctahedral geometry (14 faces) as demonstrated This partly selective mechanism competes with a non-
by quantum-mechanical calculations and experimental selective mechanism occurring exclusively on the metal
studies. For even smaller particles a non-fcc arranged surface. Alternatively, the partly selective ring opening of
10 H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21

metal catalysts, ring opening of fused C6 naphthenic rings


requires the addition of an acidic function to promote the
isomerization of the C6 rings into the more reactive C5 rings
for achieving reasonable ring-opening activity. The ring-
opening reactions are complex, usually accompanied by
hydrogenation, isomerization and cracking.

4.1. Indan

The indan molecule consists of a C5 ring fused to a


benzene ring, and represents a probable reaction intermediate
of reactions of heavy oil fractions on ring-opening catalysts.
The ring opening of indan was recently evaluated using noble
Fig. 8. Partly selective ring opening of methylcyclopentane.
metal catalysts supported on boehmite by Nylen et al. [69].
They found that the hydrogen pressure plays a key role in the
reaction. At atmospheric pressure the ring opening of the C5
MCP may be explained by electronically asymmetric pairs ring was dominant, leading to 2-ethyltoluene (60–70 mol%)
of metal atoms resulting from strong metal–support and minor amounts of n-propylbenzene (7–10%) (Eq. (4)).
interactions. The cationic sites of the support in bifunctional The Ir and Pt-Ir catalysts showed superior catalytic activities
catalysts exert an electric field on adjacent metal sites, to the Pt catalyst. Pt was shown to selectively open the C5
leading to the formation of partially charged metal atoms naphthenic ring into 2-ethyltoluene with a selectivity
that have been confirmed by various methods [33,66]. The approximately seven times higher than for n-propylbenzene,
resulting Md–Md+ exhibit functionalities similar to those of while Ir displays relatively high cracking activity. At high-
the acid site–metal pair to selectively produce nHx and 2MP pressure (40 bar) and low-to-moderate temperature, hydro-
but not 3MP (Fig. 8) [63]. genation of indan into hexahydroindan was favored.
Another ‘partly’ selective mechanism for ring opening of Increasing temperature resulted in increasing both the ring
MCP has been observed over zeolite-supported metal opening and subsequent undesired cracking products.
catalysts, for example, on SAPO-11 [67], LTL [38,39,43]
and FAU [43,68]. Over these catalysts, significantly higher
selectivities to 3MP were observed compared to other (4)
supported metal catalysts of similar particle size. These
enhanced selectivities were attributed to the constraint of the
one-dimension pores of zeolites that result in a preferred
orientation of the incoming MCP molecule with its long axis 4.2. Decalin
parallel to the direction of the pores. When the MCP
molecule reaches a metal particle inside a zeolite pore, ring The six-member hydrocarbon rings are more stable than
opening preferentially takes place through one of its ends, five-member rings, with the ring strain energies of about
resulting in a 3MP/2MP product ratio higher than the 1 kcal/mol and 6–7 kcal/mol, respectively. McVicker et al.
statistical value of 0.5 (Fig. 8). Similarly, a higher 3MP/nHx [36] showed that the ease of ring opening of two-ring
ratio than the statistical value of 0.5 can be explained by the naphthenes is directly related to the relative number of
steric hindrance of the methyl group of MCP in zeolite pores saturated five-member rings in the molecule. Therefore, the
that restrict the rotation of the molecule when it approaches ring opening of decalin (4.4% conversion on 0.9% Ir/Al2O3)
the metal particle with its methyl end. with two saturated six-member rings is much slower than
perhydroindan (68%) with one six- and one five-member
ring, and bicyclo[3,3,0]octane of two five-member rings. An
4. Selective ring opening of multiple-ring acidic function in the catalyst has to be used to achieve a
compounds significant amount of ring-opening products by promoting
the isomerization of six-member ring structures into the five-
Compared to the number of studies using mono C4, C5 member ones.
and C6 naphthenic rings, there are fewer studies on ring On acidic catalysts, the direct ring opening of decalin was
opening of more complex molecules containing multiple thought to proceed via cleavage of the corresponding
rings. Recently, model compounds containing two fused carbonium ion, which is initiated by the attack of a Brönsted
rings, such as indan, decalin, tetralin and naphthlene have acid site on a C–C bond of decalin [70]. The alkylnaphthene
been used to model the upgrading of heavy oil fractions. carbenium ion can further undergo b-scission and/or
Unlike ring opening of alkylcyclopentanes, which can isomerization to form lighter naphthenes or gaseous
readily occur at low temperature by hydrogenolysis on noble molecules. Meanwhile, bimolecular reactions may occur,
H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21 11

Fig. 9. Proposed mechanism for direct ring opening over acid catalyst [70]. PC, pyrolytic cracking; ISO, isomerization; TA, transalkylation; b, b-scission; HT,
hydride transfer.

such as hydrogen transfer, transalkylation or disproportio- reactive than trans-decalin, and converts much more
nation to produce various naphthenic species, as shown in selectively to ring-opening products than trans-decalin.
Fig. 9. Alternatively, ring opening of decalin can take place The latter mainly converts to cracking products. The skeletal
as a result of isomerization (ring contraction), which isomerization was initiated by ring contraction of decalin
produces five-member ring hexahydroindan. In other words, carbocations or carbenium ions to methylbicyclo[4,3,0]no-
the ring-opening products of decalin originated from the nanes, followed by their isomerization or further ring
isomerization products. Indeed, stereo- and skeletal iso- contraction (Fig. 11). The isomers can undergo further
merizations of decalin were found to dominate the reaction isomerization to yield a complicated mixture of isomers.
of decalin on proton-form zeolites including medium-pore Among those isomers, methylbicyclo[4,3,0]nonanes,
ZSM-5, MCM-22, large-pore HY, H-Beta, H-Mordenite, dimethylbicyclo[3,2,1]octanes, dimethylbicyclo[3,3,0]oc-
UTD-1 and extra-large-pore H-MCM-41 [71,72]. The tanes and methylbicyclo[3,3,1]nonanes are the most
decalin isomers were then consumed by consecutive abundant. The pore size of zeolites influences the isomer
reactions of ring opening and cracking, yielding more than product distribution, while the acidity and temperature do
200 products. Stereoisomerization of decalin was thought to not affect the selectivity of isomerization. Decalin hardly
occur on a Brönsted site via a pentacoordinated carbocation penetrates the channels of the medium-pore zeolites (ZSM-
(Fig. 10b), or carbenium ion formed by protonic dehy- 5, MCM-22); these zeolites are much less active than large-
drogenation, or proton exchange, or on Lewis acid sites by pore zeolites as the reaction takes place only on external
hydride abstraction (Fig. 10c). The cis-decalin is much more surfaces [70]. In addition, the pore topology of zeolite has a
strong influence on diffusion and, consequently, on activity
and selectivity. As a result, dimethylbicyclooctanes were

Fig. 10. Isomerization of cis-decalin to trans-decalin [71]. Fig. 11. Proposed isomerization reaction network for decalin [71].
12 H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21

Fig. 12. Proposed mechanism for isomerization and ring opening of decalin over acidic catalyst [71]. PC, pyrolytic cracking; ISO, isomerization; DS,
desorption; b, b-scission; HT, hydride transfer.

found to be more abundant over H-Y than over H-Beta A mechanism for decalin isomerization and ring opening has
zeolites, while more methylbicyclononanes are formed over been proposed (Fig. 12), involving pyrolytic dehydrogena-
Beta zeolites [71]. The isomers are then converted into ring- tion, pyrolytic cracking, hydride transfer, b-scission and
opening and cracking products. The ring strain and the olefin desorption and adsorption.
number of alkyl substituents in the isomers affect the ring Introduction of noble metal onto acidic zeolite catalysts
opening. Strain in the hydrocarbon ring results in ring reduces the strength of Brönsted acid sites, and significantly
opening following the order C3 > C4  C5  C7  C6. enhances isomerization and ring opening of decalin [72,73],
The isomers with more alky substitutes are more prone to while the cracking is reduced. Stereosiomerization of cis-
ring opening by acid because these compounds consist of decalin to trans-decalin proceeds on a metal site via the
more tertiary carbon atoms that more easily form carbenium dehydrogenation–hydrogenation mechanism (Fig. 10a), in
ions than the secondary carbon. As a result, the main ring- addition to the stereoisomerization facilitated solely by acid
opening products on large-pore zeolites HY, H-Beta, and H- sites (Fig. 10b and c). The skeletal isomerization occurs on
Mordenite [71] are propyl-ring-opened products (ROP) Brönsted acid sites as shown in Fig. 13, which is an essential
(from methylbicyclo[4,3,0]nonanes), followed by ethyl- step for the ring opening of decalin. In the absence of Brönsted
ROP (from dimethyl bicycle[3,2,1]octane with 1 or 2 methyl acid sites, neither isomerization nor ring-opening reactions
groups on the five member ring), methyl-ROP (from occur. The main isomerization products are methylbicy-
dimethyl bicycle[3,2,1]octane with 2 methyl groups on clo[4,3,0]nonanes and dimethylbicyclo[3,2,1]octanes, simi-
the six member ring), ethylpropyl-ROP and butyl-ROP. The lar to those on proton-form zeolites [70,71]. However, the
different isomer product distributions on different zeolites distributions of ring-opening products over Pt-modified
are reflected in the ring-opening product (ROP) distribu- zeolites are different from those obtained on H-zeolites.
tions, i.e. relatively more ethyl-ROPs and less propyl-ROPs The dominant ring-opening products are methyl-cyclohex-
are observed over H-Y zeolites in comparison with H-Beta. anes, followed by ethyl-cyclohexanes. Ring opening yields of

Fig. 13. Proposed mechanism for isomerization and ring opening of decalin over bifunctional catalyst [73]. PC, pyrolytic cracking; ISO, isomerization; DS,
desorption; b, b-scission; HT, hydride transfer; HYG, de-/hydrogenation; RO, ring opening.
H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21 13

as high as 30 mol% were obtained; in comparison just 8 mol% methylindanes and methylindenes. Subsequently, dealkyla-
of ring-opening products were achieved on H-zeolites. The tion occurred to produce cracking products, while dis-
differences were attributed to the addition of Pt. In addition to proportionation between carbenium and olefins led to heavy
ring opening initiated by Brönsted acid sites as observed on H- products and coke. Fast hydride transfer between tetralin and
zeolites, Pt particles participated in the ring-opening reaction: the adsorbed carbenium ions was attributed to the formation
(1) to form olefin by dehydrogenation of decalin isomers or of naphthalene, which was found to be one of the primary
ring-opening products, which were then protonated over products.
Brönsted acid sites and underwent isomerization and ring On bifunctional catalysts, ring opening of tetralin was
opening; (2) to directly open rings of isomers and (3) to believed to occur exclusively on Brönsted acid sites, while
provide spillover hydrogen to suppress the secondary metal sites hydrogenated tetralin to decalin and promoted its
reactions (e.g. cracking) and to prevent the catalyst isomerization (ring contraction) which is a crucial step in the
deactivation. ring opening of decalin as described in the previous section.
Hydrogenation of tetralin to decalin was found to dominate
4.3. Tetralin on supported noble metal catalysts, especially at low
temperature [72,74–79]. The hydrogenation of tetralin was
Ring opening of tetralin via cracking on acidic supports reported to take place not only on metal centers but also on
produced mainly C1 to C4 fractions, benzene, naphthalene acid sites by hydrogen spillover from the metal surface. The
and C10 olefin/naphthene aromatics (mainly methylindanes formed decalin underwent skeletal isomerization, and then
and naphthalene) [70,72]. Their relative selectivity varied subsequent ring-opening and cracking reactions. The rate of
from one zeolite to another depending on pore topology and isomerization, and hence the ring-opening rate increased
size. Extensive cracking dominated on zeolites with medium with the decrease in the distance between the metal and acid
pore sizes (ZSM-5, MCM-22, ITQ-2) to form C3 and C4 sites and with the increase in metal loading. The ring-
fractions. Large pore zeolites (Beta, Y) were able to open the opening product yields and product distribution were related
naphthenic ring and minimize the dealkylation of the formed to the topology of zeolite structure, i.e. diffusional
alkylaromatics. Extra-large pore zeolites (UTD-1) and limitations. The acidity, on the other hand, shifted the
mesoporous MCM-41 favored ring opening. A mechanism optimal (maximum yield) reaction temperature for ring
was proposed to account for the reaction, involving the opening. Mild acidity favored ring opening.
attack of a Brönsted acid site either on the aromatic ring or Supported noble metal catalysts are very sensitive to
on the naphthenic ring (Fig. 14). The resulting carbenium sulfur compounds. Efforts have been made to improve the
and carbonium ions underwent isomerization (ring contrac- sulfur resistance of noble metal-based catalysts for
tion), b-scission and protolytic cracking to form ring- hydrogenation. For instance, bimetallic catalysts, e.g. Pd-
opening products and isomerization products such as Pt, have demonstrated moderate sulfur tolerance in

Fig. 14. Proposed reaction networks for catalytic cracking of tetralin [70]. Iso, isomerization; PC, pyrolytic cracking; b, b scission.
14 H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21

hydrogenation [74,77,79]. However, the influence of sulfur compounds, consisting of products of cracking (alkane
on the activity for ring opening of naphthenes is still unclear. gases, benzene), hydrogenation (decalin), isomerization
Rodriguez-Castellon et al. [74] found that the ring opening (decalin isomers), ring opening (alkylbenzene, monocyclo-
and cracking of tetralin on Pt-Pd catalysts supported on paraffins), dehydrogenation (naphthalene) and alkylation
zirconium doped mesoporous silica were suppressed in the (alkyltetralin, tricyclic compounds). NiW was found to
presence of sulfur compounds, but not as much as hydrogenate aromatic or olefinic compounds that were
hydrogenation. produced in the reactions, and to supply spillover hydrogen
Ring opening of tetralin on transition metal catalysts, e.g. to the acid sites. Due to an insufficient supply of hydrogen
NiMo and NiW supported on acidic materials, such as with NiW sulfide, hydrocracking of tetralin proceeded
silica–alumina or zeolites has been investigated [80–86]. mainly via a bimolecular pathway (Fig. 15), in addition to a
These catalysts are able to catalyze the ring-opening reaction unimolecular mechanism, at the initial reaction stage,
of naphthalenes in the presence of sulfur compounds. Such leading to the formation of heavy compounds. In the long
catalysts are usually used at severe operating conditions, run, the unimolecular process was dominant, and the heavy
such as high hydrogen pressure and high-temperature, due to compounds were gradually converted to light products. A
their relatively low activities. Unlike noble metal catalysts, careful balance between the metal active sites of hydro-
the loading of NiW onto zeolites increases the amounts of genation and acid sites of cracking was necessary for
Brönsted acid sites [80,85]. However, the additional acid optimal performance for selective ring opening. Under
sites themselves do not catalyze hydrocracking of tetralin. conditions with high hydrogen/tetralin ratios, on the other
The increased conversion of tetralin over pure zeolites can hand, hydrocracking of tetralin proceeds via a monomole-
be attributed to the bifunctionality. Sato et al. [80–82] cular mechanism over transition metal or metal sulfide
reported hydrocracking of tetralin over bifunctional NiW catalysts [83–85], producing cracking compounds (volatiles,
sulfide catalysts supported on zeolites USY, HY and benzene, toluene, xylene), ring opening (n-propylbenzene,
mordenite (MOR) under typical hydrocracking conditions n-butylbenzene), isomerization (indan), hydrogenation
at moderate temperature (623 K) and high hydrogen (decalin) and dehydrogenation (naphthalene) compounds.
pressure (6.1 MPa) with a low H2/feed ratio. They found No products heavier than decalins were observed [83–85].
that the ring opening of tetralin required relatively strong
acid sites, and was related to the hydrogen transferability of 4.4. Naphthalene
supports: NiW/USY > NiW/HY > NiW/MOR > NiW/
Al2O3. The NiW/Al2O3 catalyst functions mainly as a Hydrogenation of naphthalene to tetralin is dominant in
hydrogenation catalyst, yielding decalin with small propor- the hydroconversion of naphthalene over bifunctional
tions of heavy compounds, while zeolites and NiW/zeolite catalysts at low temperatures [40,87–95]. The hydrogena-
catalysts produced a complex mixture of more than 300 tion rate of naphthalene to tetralin is an order of magnitude

Fig. 15. Proposed bimolecular mechanism for catalytic cracking of tetralin [81]. A, acid, DS, desorption.
H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21 15

operating temperature exists, e.g. at ca. 350 8C for ring


opening of 1-methylnaphthalene over Pt/USY [40].
The ring-opening product yields are also dependent on
the Brönsted acidity. Arribas and Martinez [40] studied the
influence of zeolite acidity for the coupled hydrogenation
and ring opening of 1-methylnaphthalene on Pt/USY
catalysts with varied aluminum content (i.e. acidity). They
found that selective ring opening was favored over zeolites
with a low density of Brönsted acid sites. Strong and high
density of Brönsted acid sites lead to extensive cracking and
Fig. 16. Hydrogenation networks for naphthalene over bifunctional cata-
lysts [88]. M, metal sites; A, acid sites; Hsp, spilt-over hydrogen.
dealkylation. They also found that the selective ring opening
of 1-methylnaphthalene increased with decreasing space
velocity and increasing total pressure with both cases
higher than that of tetralin to decalin, due to the strong exhibiting a maximum for ring-opening conversion. The
adsorption of naphthalene on the metal surface [87]. unimolecular mechanism was proposed for the hydrocrack-
Because of a thermodynamic limitation, the hydrogenation ing of 1-methylnaphthalene (Fig. 17).
conversions decreased with the increase in reaction
temperature, while the isomerization and ring-opening 4.5. Phenanthrene
reaction increased [91]. In a bifunctional catalyst system,
the hydrogenation takes place via conventional metal- Few studies on selective ring opening of three-ring cyclic
catalyzed hydrogenation, or acid site induced hydrogena- compounds have been reported. Several early studies were
tion, involving migration of spillover hydrogen from metal carried out to elucidate the reaction networks for hydro-
sites (Fig. 16). The latter was thought to account for the conversion of three- or four-ring compounds on bifunctional
increased hydrogenation activity of metal catalysts on acidic catalysts under hydrocracking conditions. Nonetheless,
supports [88,94]. these studies shed insights into understanding of the ring
Ring opening of naphthalene is a secondary reaction, opening of these model compounds. The studies are
converting saturated and/or isomerized products on metal or summarized as follows.
Brönsted sites. Albertazzi et al. [90] showed that Pd(4)-Pt(1) Phenanthrene was considered a suitable model compound
supported on Mg/Al mixed oxide, which is a basic support for hydrogenation and hydrocracking of liquefied coal over
with no Brönsted acid sites, produced an appreciable amount bifunctional catalysts. Lemberton and Guisnet [96] carried
of ring opening (18 mol% yield), mainly alkycyclohexanes, out a model test of hydroconversion of phenanthrene over
at moderate conditions around 300 8C. In the presence of NiMo sulfide supported on alumina at 430 8C, 10 MPa of H2
acid sites, the yield of ring opening increased up to 30 mol% in a stainless steel autoclave. They found that phenanthrene
[91]. hydroconversion proceeded through multi-step hydrogena-
The ring opening increased with the increase in reaction tion, isomerization and cracking reactions. Phenanthrene
temperature. However, consequent cracking reactions of initially was hydrogenated into dihydrophenanthrene, which
ring-opening products are favored at high temperatures in turn was hydrogenated into tetrahydro- and octahydro-
(>375 8C), leading to lighter products. The optimal phenanthrenes. But further hydrogenation into perhydro-

Fig. 17. Hydroconversion networks for 1-methylnaphthalene over bifunctional catalysts [40]. ISO, isomerization; DS, desorption; b, b-scission; HT, hydride
transfer; HYG, de-/hydrogenation.
16 H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21

Fig. 18. Proposed mechanism for phenanthrene hydroconversion [96]. ISO, isomerization; DS, desorption; b, b-scission; HT, hydride transfer.

phenanthrene was not observed. On NiMo/Al2O3 with only those of the central ring, and isomerization of sym-
weak acidity, ring opening took place through the opening of octahydrophenanthrene was favored over those of asym-
the saturated central ring of phenanthrene, yielding mainly octahydrophenanthrene. Isomerizations and ring opening
2-ethyl biphenyl, or the opening of a saturated terminal ring, preferably occurred at position a to an aromatic ring or a
yielding mainly butyl naphthalene, methylpropyl naphtha- tertiary carbon because of the ability of the parent structure
lene and probably diethyl naphthalene. The former was with tertiary carbons or aromatic rings to stabilize a
proposed to occur on an acid site, initiated by an acid attack carbocation intermediate. Therefore, isomerizations proceed
to a benzene ring (Fig. 18), via cleavage of aromatic C– in the following orders: tetrahydrophenanthrene > sym-
secondary C bond rather than cleavage of the 9–10 C–C octahydrophenanthrene > asym-octahydrophenanthrene >
bond because of the formation of stable conjugated tetralin. Similar trends have been observed for ring opening
carbocation (2-ethyl biphenyl). The ring opening of the and dealkylation.
saturated terminal ring, on the other hand, was thought to
proceed via a bifunctional mechanism, involving isomer- 4.6. Fluorene and fluoranthene
ization of the C6 ring into the C5 ring, followed by ring
opening of the isomers on acid sites. Cleavage of aromatic The hydrogenation of fluorene initially occurred over
C–secondary C bond was highly disfavored because of the bifunctional catalysts [98]. The primary product hexahydro-
involvement of an unstable primary carbocation. fluorene was then converted by isomerization, ring opening
Similar results were obtained by Korre et al. [97] on a and dealkylation. However, isomerization of hexahydrofluor-
presulfided NiW/USY catalyst at 350 8C, 68.1 atm of H2 in a ene took place in parallel with ring opening, and the ring
batch autoclave. They presented detailed kinetic studies on opening of hexahydrofluorene and its isomers occurred
the hydrocracking of phenanthrene. A reaction network through the central five-member ring (Fig. 20). In the presence
consisting of hydrogenation, isomerization, ring opening of a strong acid (e.g. on NiMo/zeolite Y), ring opening of
and dealkylation was established (Fig. 19), which was then hexahydrofluorenes occurred preferably in parallel with
used to obtain various reaction rates, equilibrium and isomerization and dealkylation, yielding mainly cyclohexyl-
adsorption parameters. Di- and tetrahydrophenanthrenes methylenebenzene and cyclohexyltoluene.
were found as the primary products. The latter was The hydroconversion of fluoranthene follows a similar
hydrogenated further to octahydrophenanthrene, but with mechanism. Initial hydrogenation of fluoranthene on NiMo/
no perhydrophenanthrene observed. Terminal ring saturated zeolite Y yields tetrahydrofluoranthene as the major primary
tetra- and octahydrophenanthrenes then underwent isomer- product, which is then converted by cracking and
ization and subsequent ring opening, while the ring-opening isomerization to produce mainly 1-phenyltetralin, 2-phe-
products (dimethyl- or ethylbiphenyls) of central ring nyltetralin, 2-phenlymethylindan, tetralin and benzene. Ring
saturated dihydrophenanthrene were not observed. The opening of the five-member ring of tetrahydrofluoranthene
ring-opening conversion of the latter was thought to proceed immediately followed the hydrogenation of fluoranthene
via reversible dehydrogenation/hydrogention to the terminal due to its high ring strain. The formed 1-phenyltetralin can
ring saturated compounds. Kinetic studies showed that isomerize to the thermodynamically more stable isomer 2-
hydrogenations of terminal aromatic rings were favored over phenyltetralin. 1-Phenyltetralin can recombine with a proton
H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21 17

are the most active catalysts for selective ring opening of


naphthenes. However, these catalysts are very sensitive to
sulfur compounds in real hydrocarbon feedstocks, which
normally need a severe hydrotreating pretreatment to reduce
the sulfur concentration to below a few ppm. The application
of these catalysts would thus be limited by such severe
pretreatment conditions, unless the sulfur tolerance pro-
blems could be solved. It is believed that the sulfur tolerance
of a noble metal catalyst is related to the electron density of
metal clusters. The sulfur poisoning is mainly due to the fact
that adsorption of H2S decreases the metal–support
interaction, which promotes platinum migration and leads
to a growth of platinum particle size [100]. Therefore, the
sulfur tolerance can be increased by decreasing Pt electron
density and/or enhancing metal–support interactions. By
choosing tetralin hydrogenation as a model reaction for
aromatics reduction, Chiou et al. [101] investigated the
effect of sulfur on the deactivation of g-Al2O3 supported Pt
catalysts. They found that the reversible reaction scheme is
unable to explain the decreased reactivation rate with the
severity of sulfur poisoning. TPR, fast FT-IR and electron
probe microanalysis (EPMA) suggested that the interactions
between CO and Pt, and H2 and Pt were weakened due to the
increase in sulfur poisoning. A sintering of platinum
particles occurred, promoted by the H2S adsorption at the
Pt–alumina interface, which may have decreased the Pt–
alumina interaction.
It is possible to improve sulfur resistance through the
addition of a second metal component to Pt or Pd catalyst,
i.e. via bimetallic interaction. There are well-documented
examples that alloying of metals can increase the activity,
selectivity and stability in catalytic reactions [102]. The
formation of the bimetallic phase could inhibit the Pt or Pd
migration. In addition, electron transfer from Pt or Pd to the
second metal would increase the sulfur tolerance of the
catalyst. Fujikawa et al. [103] performed screening tests of
Fig. 19. Phenanthrene hydroconversion networks over a presulfided NiW/
various SiO2-Al2O3 noble metal catalysts (Pt, Pd, Re, Sn, Ir,
USY catalyst [97]. Relative pseudo-first order reaction constants are shown.
Ni, Mo and Ge) with hydrotreated light cycle oil (LCO)/
straight-run light gas oil (SRLGO) feedstocks containing
to form a carbenium ion, which rearranges to a more stable 32–34 vol% aromatics and 172–474 ppm sulfur at a
benzylic carbocation by 1,5- and/or 1,3-hydride transfer temperature of 573 K, H2 pressure of 4.9 MPa, and LHSV
from the saturated ring. The latter is subject to a of 1.5 h1. They found that the Pt-Pd/SiO2-Al2O3 catalyst
hydrogenolytic cleavage of the C–C bond between two was the most highly active catalyst for aromatic hydro-
aromatic rings to give tetralin and benzene. 2-Phenyltetralin, genation under the applied conditions. The co-existence of
on the other hand, does not form benzylic carbocation Pt with Pd on SiO2-Al2O3 remarkably enhanced the catalytic
because of unfavorable hydride transfer in this conforma- activity, which depended on the Pd/(Pt + Pd) weight ratio
tion. As a result, isomerization of 2-phenyltetralin took and reached a maximum at about 0.7 Pd/(Pt + Pd) weight
place, yielding 2-phenylindan. The reaction networks for ratio. TEM studies indicated that all active metals form Pt-
hydrogenation and ring opening of fluoranthene are Pd bimetallic agglomerates on the catalyst. Long-term
summarized in Fig. 21. stability tests demonstrated the excellent stability of the Pd-
Pt/SiO2-Al2O3 catalyst. Similar results have been reported
by Lin et al. [100]. The supported Pt catalysts on g-Al2O3
5. Sulfur tolerance were modified by adding a second metal, such as Co, Mo, Ni,
Re, Ag and Pd. The results showed that Pd-Pt catalyst had
The discussion thus far has considered ring-opening the highest sulfur resistance. Long-term stability tests over
reactions in model compounds where noble metal catalysts Pt and Pd-Pt catalysts indicated the catalyst exhibited much
18 H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21

Fig. 20. Postulated mechanism for fluorene hydroconversion [98]. ISO, isomerization; DS, desorption; b, b-scission.

Fig. 21. Postulated mechanism for fluoranthene hydroconversion [99].

better catalytic performance for hydrogenation of diesel between sulfur and metal and increases the sulfur tolerance
feedstocks containing 28.4% aromatics, 369 ppm sulfur, of the catalyst. Song and Schmitz [89] studied the
44.8 ppm nitrogen under the following commercial diesel hydrogenation of naphthalene in n-tridecane at 200 8C in
hydrotreating conditions: WHSV 3.0 h1, H2 to oil mole the absence and presence of benzothiophene over mordenite
ratio 2.5, temperature 340 8C, pressure 580 psig. Pd-Pt and zeolite Y-supported Pt and Pd, compared with Al2O3-
bimetallic interaction played a crucial role in improving and TiO2-supported Pt and Pd catalysts. Both zeolite-
sulfur resistance. Fast FT-IR studies of CO adsorption supported catalysts are substantially more active than the
suggested that electrons are transferred from Pt to Pd by Al2O3- and TiO2-supported catalysts. The presence of
bimetallic interaction, inhibiting the adsorption of H2S. In benzothiophene decreases the activity of all the catalysts.
addition, it was inferred that the Pd inhibited agglomeration However, there are significant improvements in sulfur
of Pt particles during hydrogen regeneration. tolerance of the noble metals when supported on zeolites.
Another approach to enhance the sulfur resistance of Mordenite-supported Pd catalyst has shown the best activity
noble metal catalysts is to fine tune the metal–support and the highest sulfur resistance [89].
interaction. It has been reported that the sulfur resistance of
Pt catalyst can be increased by using acidic support, such as
SiO2-Al2O3 or zeolites. The interaction between the strong 6. Concluding remarks
acid site and the small cluster of noble metal results in the
electrons being withdrawn from the noble metal, thus Currently, there is increasingly interest in selective ring-
creating an electron-deficient metal particle [89,104–107]. opening catalysis that can produce middle distillate with
This decreases the strength of the bonding interaction high cetane number with bitumen-derived crude and heavy
H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21 19

oils. Ring opening of cycloparaffins can be catalyzed via catalysts using transition metal sulfides on acidic supports
carbenium intermediates by protolytic cracking on the usually require severe operating conditions due to their low
Brönsted sites. The cracking of endocyclic C–C bonds in activities of the metal sulfide compared to the metal sites,
cyclic hydrocarbons is much slower than those of aliphatics. leading to extensive cracking of cycloparaffin side chains.
As a result, the overall reaction is predominated by Noble metals supported on acidic oxides are highly active
isomerization and subsequent hydrocracking (b-scission) catalysts for selective ring opening, but these catalysts are
of side chains of cyclic hydrocarbons, particularly of those very sensitive to poisoning by sulfur compounds in
having substituents with more than five carbon atoms, petroleum feedstocks. There is a need to develop new
leading to significant dealkylation of pendant substituents on generations of selective ring-opening catalysts with high-
the ring. The yields of desired high cetane ring-opened performance and sulfur resistance that could meet the future
products are usually low due to a high extent of consecutive increasing cetane specification.
cracking reactions and a fast catalyst deactivation. The search for such catalysts through the conventional
On the other hand, the ring opening of cycloparaffins can trial-and-error method would be tedious. During the last
proceed on certain metal catalysts via direct hydrogenolysis decade, rapid development of modern computing techniques
of an endocyclic C–C bond, i.e. cleavage of a C–C bond with and computational chemistry have enabled one to run
the addition of hydrogen. The reaction is accompanied by theoretical calculations at a relatively low cost, and at the
dehydrogenation/hydrogenation, skeletal and stereo isomer- same time allowed for the accurate descriptions of the
ization. The activity and selectivity depend mainly on the molecular structures, spectroscopy and predictions of
type of metal, particle size, crystal morphology, etc. Certain chemical reactivity. The application of computational
noble metals, such as Pt, Pd, Ir, Ru and Rh have been found chemistry in heterogeneous catalysis has grown rapidly in
to be selectively active for the ring opening for cyclic recent years, which has helped to better understand the
hydrocarbons to the corresponding paraffins with the same reaction mechanism, to explain the experimental data, and to
carbon number. However, these metal catalysts lack activity devise new catalysts. For instance, quantum theoretical
for ring opening of the six-member hydrocarbon rings. calculations of electron densities of alloy catalysts, aided by
The presence of the acid function is necessary for modern characterization techniques, such as X-ray photo-
selective ring opening of six-membered ring naphthenes in spectroscopy, scanning tunnel microscopy, XANES, TPD,
feedstocks. The acid function promotes the formation of etc., have provided insights into understanding the origin of
intermediate carbenium ions, which isomerize into five- alloying effects of the metal catalysts. These studies have
membered rings that readily undergo subsequent ring yielded qualitative relationships of the chemisorption bond
opening on acid or metal sites. The metal function supplies strength of the reactants on the metal catalysts and catalytic
spilt-over hydrogen to the acid sites to saturate the activity [28,53,102,108]. Recently, Jacobsen et al. [109]
intermediate carbenium ions and prevent the formation of found that the catalytic activity for ammonia synthesis is a
coke. The acid number, strength and distance between the function of nitrogen adsorption energy on the surfaces of the
acid site and the metal site strongly influence the catalytic various metals, which exhibits a typical volcano curve. The
performance. Insufficient supply of spilt-over hydrogen ideal catalyst should possess an optimal binding energy; the
leads to mainly acidic cracking as in the monofunctional metals that bind nitrogen either too weakly or too strongly
catalyst, and favors coke formation, whereas in a catalytic are poor catalysts. Guided by the theoretical studies, the
system with spilt-over hydrogen in a sufficiently high groups found theoretically, and confirmed experimentally,
concentration, hydrogenation dominates, and admixing of a that the alloys Co–Mo, Fe–Ru, Fe–Co and Ni–Mo were
metal-free component can further increase the conversion. more active than pure Co, Mo, Ru or Fe. Similar strategies
The optimal catalytic system requires a well-balanced have been applied to search for hydrodesulfurization
metallic-acidic function. In selective ring-opening catalysis, catalysts from sulfur binding energy–activity correlations
the preferred acid function isomerizes six-membered ring [110–117], and hydrogenation catalysts from hydrogen
naphthenes to five-membered ring naphthenes with the dissociation energy–activity correlations [118]. It is
minimum number of ring substituents [36]. Such non- expected that such strategies will provide opportunities in
branching ring contraction allows maximal ring-opening searching for new metal catalysts with high-performance
rates and product selectivities and minimal undesired and sulfur resistance for selective ring-opening catalysis.
hydrocracking and secondary acyclic paraffin isomerization In comparison, hydrogenolysis of C–C bonds in a
reactions. hydrocarbon is much more complex, and is usually
The commercial process for selective ring opening accompanied by dehydrogenation, isomerization, cycliza-
involves bifunctional catalysts, both metal and acid sites, tion, etc. Nonetheless, Zaera [53–55] discovered a relation-
working together in high-pressure, high-temperature reactor ship between the selectivity of metal catalysts in
systems in the presence of hydrogen. The acidic sites hydrocarbon reforming and the ability of metal to catalyze
catalyze dehydrogenation, cracking, isomerization and a-, b- and g-dehydrogenations from chemisorbed surface
dealkylation, while the metal sites promote hydrogenation, alkyl intermediates. The latter varies across the periodic
hydrogenolysis and isomerization. The widely available table, and can be correlated to temperature-programmed
20 H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21

desorption of alkyl halides on the metal surface. The [10] G. Valavarasu, M. Bhaskar, K.S. Balaraman, Pet. Sci. Technol. 21
temperature-programmed desorption data, which corre- (2003) 1185.
[11] J.W. Ward, Fuel Process. Technol. 35 (1993) 55.
spond to the temperature maxima for the desorption of [12] L.B. Galperin, J.C. Bricker, J.R. Holmgren, Appl. Catal. A 239
methane, ethylene or iso-butane via a-, b- or g-hydride (2003) 297.
eliminations, respectively, indicate the degree of difficulty of [13] K.A. Cumming, B.W. Wojciechowski, Catal. Rev. Sci. Eng. 38
the dehydrogenation reactions. Metals with the ability to (1996) 101.
[14] J. Weitkamp, S. Ernst, H.G. Karge, Erdol und Kohle Erdgas 37 (1984)
promote a-hydride elimination facilitate hydrogenolysis in
457.
hydrocarbon reforming. Such data, though scarce and [15] G.A. Mills, H. Heinemann, T.H. Milliken, A.G. Oblad, Ind. Eng.
scattered in the literature, have yielded some useful insights Chem. 45 (1953) 134.
into the hydrocarbon reforming catalysis. Research into this [16] P.B. Weisz, E.W. Swegler, Science 126 (1957) 31.
direction and establishment of quantitative structure– [17] F. Roessner, U. Roland, J. Mol. Catal. A 112 (1996) 401.
activity relationships could provide opportunities to discover [18] U. Roland, T. Braunschweig, F. Roessner, J. Mol. Catal. A 127 (1997)
61.
new ring-opening catalysts that could meet future fuel [19] W.C. Conner Jr., J.L. Falcone, Chem. Rev. 95 (1995) 759.
specifications. [20] B. Delmon, G.F. Froment, Catal. Rev. Sci. Eng. 38 (1996) 69.
High throughput experimentation methods, invented in [21] B. Delmon, Solid State Ionics 101–103 (1997) 655.
the pharmaceutical industry for fast discovery of new drugs [22] K.M. Sancier, J. Catal. 20 (1971) 106.
has attracted a lot of attention in the catalysis research field [23] S.T. Srinivas, P. Kanta Rao, J. Catal. 148 (1994) 470.
[24] S. Ceckiewicz, B. Delmon, J. Catal. 108 (1987) 294.
[119]. Many research institutes and companies have adopted [25] P. Antonucci, N.V. Truong, N. Giordano, R. Maggiore, J. Catal. 75
this technology when searching for new catalysts or for (1982) 140.
improving existing catalyst performance. This method [26] K. Fujimoto, in: T. Inui, K. Fujimoto, T. Uchijima, M. Massi (Eds.),
allows the synthesis and catalyst testing of tens and even Stud. Surf. Sci. Catal., vol. 77, Elsevier, Kyoto, 1993, p. 9.
[27] S. Ohgoshi, I. Nakamura, Y. Wakushima, in: T. Inui, K. Fujimoto, T.
hundreds of materials at one time, greatly speeding up the
Uchijima, M. Masai (Eds.), Stud. Surf. Sci. Catal., vol. 77, Elsevier,
catalyst development process. Such techniques will provide Kyoto, Japan, 1993, p. 289.
new opportunities in searching for bi- or multi-metallic [28] B. Coq, F. Figueras, Coord. Chem. Rev. 178–180 (1998) 1753.
metal catalyst with high-performance and sulfur resistance [29] F. Garin, G. Maire, Acc. Chem. Res. 22 (1989) 100.
for selective ring-opening catalysis, and/or improving [30] F.G. Gault, Adv. Catal. 30 (1981) 1.
hydrocracking catalysts with well-balanced acid and metal [31] M. Che, C.O. Bennett, Adv. Catal. 36 (1989) 55.
[32] G.L. Haller, D.E. Resasco, Adv. Catal. 36 (1989) 173.
functions. With the guidance of today’s advanced computa- [33] K. Hayek, R. Kramer, Z. Paal, Appl. Catal. A 162 (1997) 1.
tional methods, the search will be even more markedly [34] H. Zimmer, Z. Paal, J. Mol. Catal. 51 (1989) 261.
shortened by zeroing in a limited number of promising [35] Y. Zhuang, A. Frennet, Appl. Catal. A 134 (1996) 37.
candidates. [36] G.B. McVicker, M. Daage, M.S. Touvelle, C.W. Hudson, D.P. Klein,
W.C. Baird Jr., B.R. Cook, J.G. Chen, S. Hantzer, D.E.W. Vaughan,
E.S. Ellis, O.C. Feeley, J. Catal. 210 (2002) 137.
[37] M. Chow, S.H. Park, W.M.H. Sachtler, Appl. Catal. 19 (1985) 349.
Acknowledgements [38] G. Jacobs, F. Ghadiali, A. Pisanu, A. Borgna, W.E. Alvarez, D.E.
Resasco, Appl. Catal. A 188 (1999) 79.
Partial funding for the National Centre for Upgrading [39] W.E. Alvarez, D.E. Resasco, J. Catal. 164 (1996) 467.
[40] M.A. Arribas, A. Martinez, Appl. Catal. A 230 (2002) 203.
Technology (NCUT) has been provided by the Canadian
[41] D. Teschner, L. Pirault-Roy, D. Naud, M. Guerin, Z. Paal, Appl.
Program for Energy Research and Development (PERD), Catal. A 252 (2003) 421.
the Alberta Research Council (ARC) and the Alberta Energy [42] D. Teschner, K. Matusek, Z. Paal, J. Catal. 192 (2000) 335.
Research Institute (AERI). The authors thank Mr. Norman [43] M. Vaarkamp, P. Dijkstra, J. van Grondelle, J.T. Miller, F.S. Modica,
Sacuta for proofreading the manuscript. D.C. Koningsberger, R.A. van Santen, J. Catal. 151 (1995) 330.
[44] B. Torok, M. Bartok, J. Catal. 151 (1995) 315.
[45] I. Palinko, J. Catal. 168 (1997) 543.
[46] B. Torok, I. Palinko, A. Molnar, M. Bartok, J. Catal. 159 (1996) 500.
References [47] D. Teschner, Z. Paal, D. Duprez, Catal. Today 65 (2001) 185.
[48] F. Figueras, B. Coq, C. Walter, J.-Y. Carriat, J. Catal. 169 (1997) 103.
[1] T.G. Kaufmann, A. Kaldor, G.F. Stuntz, M.C. Kerby, L.L. Ansell, [49] G. Onyestyak, G. Pal-Borbely, H.K. Beyer, Appl. Catal. A 229 (2002)
Catal. Today 62 (2000) 77. 65.
[2] C. Marcilly, J. Catal. 217 (2003) 47. [50] J.-Y. Saillard, R. Hoffmann, J. Am. Chem. Soc. 106 (1984) 2006.
[3] S. Rossini, Catal. Today 77 (2003) 467. [51] E. Schustorovich, R.C. Baetzold, E.L. Muetterties, J. Phys. Chem. 87
[4] A. Nishijima, T. Kameoka, T. Sato, N. Matsubayashi, Y. Nishimura, (1983) 1100.
Catal. Today 45 (1998) 261. [52] E. Schustorovich, R.C. Baetzold, J. Am. Chem. Soc. 102 (1980)
[5] J. Barbier, E. Lamy-Pitara, P. Marecot, J.P. Boitiaux, J. Cosyns, F. 5989.
Verna, Adv. Catal. 37 (1990) 279. [53] F. Zaera, Catal. Lett. 91 (2003) 1.
[6] A. Stanislaus, B.H. Cooper, Catal. Rev. Sci. Eng. 36 (1994) 75. [54] F. Zaera, J. Phys. Chem. B 106 (2002) 4043.
[7] M.F. Wilson, J.F. Kriz, Fuel 63 (1984) 190. [55] F. Zaera, Appl. Catal. A 229 (2002) 75.
[8] M.F. Wilson, I.P. Fisher, J.F. Kriz, Ind. Eng. Chem. Prod. Res. Dev. 25 [56] R. Kramer, H. Zuegg, J. Catal. 85 (1984) 530.
(1986) 505. [57] R. Kramer, H. Zuegg, J. Catal. 80 (1983) 446.
[9] M.F. Wilson, I.P. Fisher, J.F. Kriz, Energy Fuels 1 (1987) 540. [58] G. Rupprechter, K. Hayek, H. Hofmeister, J. Catal. 173 (1998) 409.
H. Du et al. / Applied Catalysis A: General 294 (2005) 1–21 21

[59] D. Kalakkad, S.L. Anderson, A.D. Logan, J. Pena, E.J. Braunsch- [89] C. Song, A.D. Schmitz, Energy Fuels 11 (1997) 656.
weig, C.H.F. Peden, A.K. Datye, J. Phys. Chem. 97 (1993) 1437. [90] S. Albertazzi, G. Busca, E. Finocchio, R. Glockler, A. Vaccari, J.
[60] L. Pirault-Roy, D. Teschner, Z. Paal, M. Guerin, Appl. Catal. A 245 Catal. 223 (2004) 372.
(2003) 15. [91] M. Jacquin, D.J. Jones, J. Roziere, S. Albertazzi, A. Vaccari, M.
[61] J.T. Miller, B.L. Mojet, D.E. Ramaker, D.C. Koningsberger, Catal. Lenarda, L. Storaro, R. Ganzerla, Appl. Catal. A 251 (2003) 131.
Today 62 (2002) 101. [92] C. Petitto, G. Giordano, F. Fajula, C. Moreau, Catal. Commun. 3
[62] M. Breysse, P. Afanasiev, C. Geantet, M. Vrinat, Catal. Today 86 (2002) 15.
(2003) 5. [93] S. Albertazzi, R. Ganzerla, C. Gobbi, M. Lenarda, M. Mandreoli, E.
[63] J.B.F. Anderson, R. Burch, J.A. Cairns, J. Catal. 107 (1987) 351. Salatelli, P. Savini, L. Storaro, A. Vaccari, J. Mol. Catal. A 200 (2003)
[64] T.J. McCarthy, G.-D. Lei, W.M.H. Sachtler, J. Catal. 159 (1998) 90. 261.
[65] G.M. Schwab, E. Pietsch, Z. Phys. Chem. B 1 (1928) 385. [94] A. Corma, A. Martinez, V. Martinez-Soria, J. Catal. 169 (1997) 480.
[66] R. Kramer, M. Fischbacher, J. Mol. Catal. 51 (1989) 247. [95] A.D. Schmitz, G. Bowers, C. Song, Catal. Today 31 (1996) 45.
[67] M. Hoffmeister, J.B. Butt, Appl. Catal. 82 (1992) 169. [96] J.-L. Lemberton, M. Guisnet, Appl. Catal. 13 (1984) 181.
[68] F. Garin, R. Girard, G. Maire, G. Lu, L. Guczi, Appl. Catal. A 152 [97] S.C. Korre, M.T. Klein, R.J. Quann, Ind. Eng. Chem. Res. 36 (1997)
(1997) 237. 2041.
[69] U. Nylen, J.F. Delgado, S. Jaras, M. Boutonnet, Appl. Catal. A 262 [98] A.T. Lapinas, M.T. Klein, B.C. Gates, A. Macris, J.E. Lyons, Ind.
(2004) 189. Eng. Chem. Res. 30 (1991) 42.
[70] A. Corma, V. Gonzalez-Alfaro, A.V. Orchillesy, J. Catal. 200 (2001) 34. [99] A.T. Lapinas, M.T. Klein, B.C. Gates, A. Macris, J.E. Lyons, Ind.
[71] D. Kubicka, N. Kumar, P. Maki-Arvela, M. Tiitta, V. Niemi, T. Salmi, Eng. Chem. Res. 26 (1987) 1026.
D.Y. Murzin, J. Catal. 222 (2004) 65. [100] T.B. Lin, C.A. Jan, J.R. Chang, Ind. Eng. Chem. Res. 34 (1995) 4284.
[72] M. Santikunaporn, J.E. Herrera, S. Jongpatiwut, D.E. Resasco, W.E. [101] J.F. Chiou, Y.L. Huang, T.B. Lin, J.R. Chang, Ind. Eng. Chem. Res.
Alvarez, E.L. Sughrue, J. Catal. 228 (2004) 100. 34 (1995) 4277.
[73] D. Kubicka, N. Kumar, P. Maki-Arvela, M. Tiitta, V. Niemi, H. [102] V. Ponec, Appl. Catal. A 222 (2001) 31.
Karhu, T. Salmi, D.Y. Murzin, J. Catal. 227 (2004) 313. [103] T. Fujikawa, K. Idei, T. Ebihara, H. Mizuguchi, K. Usui, Appl. Catal.
[74] E. Rodriguez-Castellon, J. Merida-Robles, L. Diaz, P. Maireles- A 192 (2000) 253.
Torres, D.J. Jones, J. Roziere, A. Jimenez-Lopez, Appl. Catal. A [104] J. Wang, L. Huang, Q. Li, Appl. Catal. A 175 (1998) 191.
260 (2004) 9. [105] L.J. Simon, J.G. van Ommen, A. Jentys, J.A. Lercher, J. Catal. 201
[75] M.A. Arribas, P. Concepcion, A. Martinez, Appl. Catal. A 267 (2004) (2001) 60.
111. [106] J. Zheng, M.J. Sprague, C. Song, Pet. Chem. Div. Prepr. 47 (2002)
[76] M.A. Arribas, A. Corma, M.J. Diaz-Cabanas, A. Martinez, Appl. 100.
Catal. A 273 (2004) 277. [107] J.T. Miller, D.C. Koningsberger, J. Catal. 162 (1996) 209.
[77] H. Yasuda, Y. Yoshimura, Catal. Lett. 46 (1997) 43. [108] C&EN, November 29, 2004, p. 25.
[78] J.-R. Chang, S.-L. Chang, J. Catal. 176 (1998) 42. [109] C.J.H. Jacobsen, S. Dahl, B.S. Clausen, S. Bahn, A. Logadottir, J.K.
[79] H. Yasuda, T. Sato, Y. Yoshimura, Catal. Today 50 (1999) 63. Norskov, J. Am. Chem. Soc. 123 (2001) 8404.
[80] K. Sato, Y. Iwata, T. Yoneda, A. Nishijima, Y. Miki, H. Shimada, [110] S. Harris, R.R. Chianelli, J. Catal. 86 (1984) 400.
Catal. Today 45 (1998) 367. [111] P. Raybaud, J. Hafner, G. Kresse, S. Kasztelan, H. Toulhoat, J. Catal.
[81] K. Sato, Y. Iwata, Y. Miki, H. Shimada, J. Catal. 186 (1999) 45. 190 (2000) 128.
[82] T. Sato, Y. Nishimura, K. Honna, N. Matsubayashi, H. Shimada, J. [112] P. Raybaud, J. Hafner, G. Kresse, S. Kasztelan, H. Toulhoat, J. Catal.
Catal. 200 (2001) 288. 189 (2000) 129.
[83] R. Hernandez-Huesca, J. Merida-Robles, P. Maireles-Torres, E. [113] X. Ma, H.H. Schobert, J. Mol. Catal. A 160 (2000) 409.
Rodriguez-Castellon, A. Jimenez-Lopez, J. Catal. 203 (2001) 122. [114] L.S. Byskov, J.K. Norskov, B.S. Clausen, H. Topsoe, J. Catal. 187
[84] D. Eliche-Quesada, J. Merida-Robles, P. Maireles-Torres, E. Rodri- (1999) 109.
guez-Castellon, A. Jimenez-Lopez, Appl. Catal. A 262 (2004) 111. [115] I.I. Zakharov, A.N. Startsev, J. Phys. Chem. B 104 (2000) 9025.
[85] D. Li, A. Nishijima, D.E. Morrisz, J. Catal. 182 (1999) 339. [116] I.I. Zakharov, A.N. Startsev, G.M. Zhidomirov, V.N. Parmon, J. Mol.
[86] E. Rodriguez-Castellon, L. Diaz, P. Braos-Garcia, J. Merida-Robles, Catal. A 137 (1999) 101.
P. Maireles-Torres, A. Jimenez-Lopez, A. Vaccari, Appl. Catal. A [117] I.I. Zakharov, A.N. Startsev, G.M. Zhidomirov, J. Mol. Catal. A 119
240 (2003) 83. (1997) 437.
[87] K. Ito, Y. Kogasaka, H. Kurokawa, M. Ohshima, K. Sugiyama, H. [118] J. Greeley, M. Mavrikakis, Nat. Mater. 3 (2004) 810.
Miura, Fuel Process. Technol. 79 (2002) 77. [119] R.J. Hendershot, C.M. Snively, J. Lauterbach, Chem. Eur. J. 11
[88] K.C. Park, D.J. Yim, S.K. Ihm, Catal. Today 74 (2002) 281. (2005) 806.

You might also like