You are on page 1of 17

Applied Catalysis A: General 229 (2002) 75–91

Selectivity in hydrocarbon catalytic reforming:


a surface chemistry perspective
Francisco Zaera∗
Department of Chemistry, University of California, Riverside, CA 92521, USA
Received 15 May 2001; accepted 25 July 2001

Abstract
A brief overview of the recent advances in the understanding of the reaction mechanisms of hydrocarbon reforming processes
is provided. Emphasis is placed on the knowledge developed by studies using model single-crystal metals and modern surface
analytical techniques. An argument is presented for the early definition of reaction selectivities in these processes because
of subtle relative changes in dehydrogenation rates. Specifically, the ability of platinum to promote ␥-hydride elimination
steps from chemisorbed alkyl groups is proposed to be the reason for its unique behavior in isomerization and cyclization
reactions. Nickel, in contrast, facilitates dehydrogenation preferentially at the ␣-position, and catalyses hydrogenolysis instead.
Nevertheless, by far the most favorable reaction of alkyl moieties on transition metals is ␤-hydride elimination. It is this
step the one that accounts for the fast equilibrium reached between alkanes and alkenes under most reaction conditions.
Additional mechanistic complications in hydrocarbon reforming under catalytic conditions are introduced by the strongly
bonded carbonaceous deposits that cover the surface of the active catalyst. The role that those deposits play in reforming
is not yet entirely clear, but it is believed to involve passivation of the surface to facilitate the adsorption of ␲-bonded
olefins and other weakly adsorbed intermediates, and also the provision of a reservoir for hydrogen. The working reforming
catalysts is therefore likely to display bifunctional character, with rapid hydrogenation–dehydrogenation steps taking place
on the hydrocarbon-covered surface, and more demanding skeletal rearrangement steps occurring on patches of bare metal.
© 2002 Elsevier Science B.V. All rights reserved.
Keywords: Hydrocarbons; Catalytic reforming; Molecular models; Surface chemistry; Single crystals; Alkyls; Platinum

1. Introduction isomerization, hydrogenation, and polymerization of


organic compounds in 1905. However, it was with the
It could be said that heterogeneous catalysis started discovery of cheap oil in the early 1950s that the de-
with the report by Davy in 1818 on the rapid oxida- velopment of catalytic processes to convert crude oil
tion of hydrogen over a platinum surface [1]. The ad- to fuels and chemicals really took off. Catalysts con-
vance of hydrocarbon processing using heterogeneous taining sulfides of molybdenum and other metals were
catalysis followed soon thereafter. Sabatier proposed a designed to remove sulfur, oxygen, nitrogen, and other
catalytic process for the hydrogenation of ethylene as impurities from oil feedstocks [2]. Zeolites and other
early as 1902, and Ipatieff reported the use of clays as aluminosilicates were introduced for the cracking of
catalysts for several hydrogenation, dehydrogenation, large petroleum molecules into smaller hydrocarbons
[3]. Titanium and zirconium complexes became the
∗ Tel.: +1-909-787-5498; fax: +1-909-787-3962. basis for several polymerization catalytic processes
E-mail address: francisco.zaera@ucr.edu (F. Zaera). [4].

0926-860X/02/$ – see front matter © 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 6 - 8 6 0 X ( 0 2 ) 0 0 0 1 7 - 0
76 F. Zaera / Applied Catalysis A: General 229 (2002) 75–91

Fig. 1. Possible reactions for the conversion of n-hexane during reforming processes [2]. The key to an efficient reforming catalyst is the
promotion of isomerization, cyclization, and aromatization at the expense of undesirable hydrogenolysis (C–C bond scission) steps.

One of the fundamental oil refining processes though many metals promote hydrocarbon hydro-
nowadays is catalytic reforming [5]. In reforming, genation, only platinum-based catalysts lead to the se-
gasoline-range molecules are reconstructed without lective production of reforming products. Nickel, for
changing their carbon content in order to improve instance, favors undesirable cracking instead. More-
the quality of the fuel, as measured by their octane over, unsupported platinum is capable of carrying out
number. Reforming processes involve many reactions, reforming catalysis by itself, without the presence of
including isomerizations, hydrogenations, dehydro- any acidic sites [11,12]. Clearly, the metal itself plays
cyclization, and dehydrogenation. An example of the a central role in all the steps of reforming.
relevant reactions in reforming is provided in Fig. 1 Extensive research has been carried out in the last
for the case of n-hexane. Reforming catalysis was century to develop a better basic understanding of
originally conducted on molybdenum oxide dispersed the unique behavior of platinum for reforming. Re-
on alumina, but a significant advance was made by searchers like Bond, Horiuti, Polanyi, and Wells pro-
the Universal Oil Products Company in 1949 with the vided fundamental ideas early on about the specific
introduction of the use of platinum for this process mechanistic details of the catalytic hydrogenation of
[6]. A number of platinum-based catalysts, mostly al- unsaturated hydrocarbons on transition metals [13,14].
loys with a second metal such as rhenium or iridium, Additional concepts associated with the conversion
have been developed ever since [7]. steps involving hydrocarbons on solid surfaces have
The modern reforming catalysts consist of small come from Anderson, Avery, Boudart, Gault, Kemball,
metal particles finely dispersed on a high surface-area Paál, Rooney, Sachtler, Sinfelt, Tétényi, and many oth-
support such as alumina. These catalysts are bi- ers [2,15–20]. Professor Guczi, to whom this article
functional, possessing two types of sites [8]. On the is dedicated, has been a major player in this endeavor
one hand, the acidic nature of the surface of the as well [21–28]. The collective work of those research
alumina support, a property enhanced by treatment groups have provided a useful, if limited, picture of
with chlorine-containing compounds [9], helps pro- the microscopic details of the chemical surface reac-
mote skeletal rearrangements in carbonium ions [10]. tions involved in catalytic reforming.
Transition metals, on the other hand, are excellent The main difficulty with most of the traditional
hydrogenation–dehydrogenation catalysts. It is be- catalytic investigations cited above is that it has in
lieved that hydrocarbon reforming requires multiple many cases been based on kinetic studies, and, con-
reactions involving both types of sites. However, even sequently, has relied on indirect evidence. A few
F. Zaera / Applied Catalysis A: General 229 (2002) 75–91 77

spectroscopic techniques such as infrared and NMR comprised of straight-chain paraffins in the range of
have long been available for the direct characterization five to eight carbons long [2]. Because saturated alka-
of surface intermediates [29,30], but the heteroge- nes are quite stable, the step that limits the activity
neous and complex nature of the surface of supported of reforming catalysts is typically associated with the
catalysts makes the interpretation of the results of scission of one of the C–H bonds in those molecules.
those studies less than straightforward. It was only With this in mind, surface chemists have spent a great
with the advent of new modern surface-sensitive tech- deal of effort on the study of the activation of alkanes
niques in the 1960s that the detailed characterization on single crystals and other model surfaces.
of basic chemical transformations on model systems Both molecular beam and adsorption–desorption
became possible [31–33]. The use of this approach techniques have indicated that the promotion of C–H
for the study of hydrocarbon catalysis was pioneered bond-scission steps in alkanes by metal surfaces may
by Somorjai and co-workers [1,11,34–36], and has occur either directly upon collision of the incoming
been one of the main directions of research in our gas molecule with the solid or, alternatively, via the
group for the last couple of decades. formation of a weakly adsorbed intermediate trapped
The purpose of this report is to summarize some on the substrate [37,38]. It has been shown that the
highlights from the results obtained using such a critical factor for alkane activation is the efficiency
modern surface-science approach to the study of hy- with which energy is transferred from the translational
drocarbon reactions on transition metal surfaces, in and internal degrees of freedom of the gas molecules
particular to reactions related to hydrocarbon reform- to the particular C–H bond to be broken; the energy
ing. Ours does not pretend to be a comprehensive re- stored by the surface seems to be of little importance
view, but rather one that focuses on our contributions for this process [37–39]. In addition, an early study by
to this field, although key advances from other groups Stewart and Ehrlich revealed that vibrational modes
are also included. This article is divided as follows. are the most likely internal source of energy for the
Section 2 provides a discussion on the initial adsorp- dissociation of methane [40]. Later work has indi-
tion and activation of alkanes as the rate-limiting step cated that most often it is the normal component of
in reforming. The chemistry of the resulting surface the translational energy of the alkane that is important
alkyl intermediates and its relevance to reforming se- for the C–H bond-breaking step [41,42].
lectivity are addressed in Section 3. Section 4 focuses Direct alkane activation may occur via a tunneling
on the preferred ␤-hydride elimination step from mechanism [43], at least at the high kinetic ener-
alkyls that leads to the formation of olefins. In Section gies associated with supersonic beam experiments.
5 a look is taken at the alternative dehydrogenation Alkane sticking probabilities often increase mono-
reactions from the ␣- and ␥-position as the ones that tonically with increasing kinetic energy, not suddenly
define selectivity between hydrogenolysis and iso- after a given threshold as expected if the kinetics
merization. Bond forming reactions are addressed in were dominated by an activation barrier. Also, the
Section 6. Section 7 discusses the complications to the observed shifts in the translational energy needed to
understanding of reforming mechanisms introduced activate CH4 versus CD4 (or C2 H6 versus C2 D6 ) are
by the atmospheric pressure environments required usually on the order of 5 kcal/mol, much larger than
for hydrocarbon catalysis. Finally, Section 8 provides the differences in the zero-point energy due to iso-
some concluding remarks on how the knowledge sum- topic substitution [42]. On the other hand, the energy
marized in the previous sections can be put together and angle dependencies of the sticking coefficient of
in a model for the working reforming catalyst. methane observed on some metals may be explained
by a classical model based on the microscopic re-
verse of the reductive elimination of methyl with
2. Alkane activation as the rate limiting step hydrogen [44]. In any case, supersonic beams probe
in reforming only a small fraction of the gas molecules in a typical
catalytic mixture, those at the high-energy tail of the
The feedstock for oil reforming processes is Boltzmann distribution. Consequently, the probability
a low-octane mixture of hydrocarbons mainly of alkane activation under normal catalytic conditions
78 F. Zaera / Applied Catalysis A: General 229 (2002) 75–91

by the same direct collision processes observed in of the catalysts. Alkane activation appears to be
molecular beam experiments is predicted to be quite mainly dependent on entropic factors, hence the low
low. This has been confirmed by regular high-pressure probabilities of reaction at all temperatures. However,
catalytic studies, which have yielded methane sticking alkane activation also displays low activation energies,
probabilities as low as 10−12 [45]. implying that reforming activity should only change
Other molecular beam work has provided evidence moderately with temperature. Other steps must con-
for a precursor-mediated alkane activation mechanism. tribute to the overall temperature dependence of the
The trapping probability of ethane on Ir(1 1 0)-(1 × 2), reaction rates.
for instance, shows only a weak dependence on the
angle of incidence of the beam, suggesting that the
momentum of the incoming molecules is rapidly trans- 3. The chemistry of alkyl surface species and
ferred to the surface [46]. In some cases, the dissoci- its role in reforming selectivity
ation of alkanes is also assisted by a reverse energy
transfer from the surface to the molecule. On both The initial activation of alkanes may be respon-
platinum and rhodium, for example, the sticking prob- sible for determining the overall rate of reforming
ability of methane displays an Arrhenius temperature processes, but the selectivity towards high-octane
dependence in the limit of low kinetic energy, with products is defined by the subsequent surface reac-
an activation energy between 5 and 10 kcal/mol [47]. tions. Specifically, the scission of the first C–H bond
Some heavier chemisorbed alkanes can also be made in the reacting alkanes leads to the formation of sur-
to react directly by heating of the substrate [48–50], face alkyl intermediates. Alkyl surface moieties can
and physisorbed methane can be activated via colli- then follow a number of subsequent reactions, hy-
sions with other impinging molecules as well [51]. dride, alkyl, and reductive eliminations, insertions,
The bonding of the intermediate precursor in and homolytic bond scissions, among others (Fig. 2).
the trapping-mediated dissociative adsorption of However, it is our contention that it is the regiose-
alkanes to the surface most likely involves a lectivity of a second dehydrogenation step on those
three-center two-electron interactions similar to that species (a hydride elimination in the terminology of
in organometallic systems [52,53]. Several pieces of organometallic chemistry) what seals the faith of the
evidence have been put forward to support this hy- final products. In order to better understand this issue
pothesis. For one, the activation barrier associated of selectivity in reforming, we in our laboratory have
with the decomposition of alkanes on metal surfaces embarked into a detailed research of the chemistry
amounts in most cases to less than one-fifth of the of alkyl species on transition metal single-crystal sur-
C–H bond energy in hydrocarbon molecules [54–56], faces [52,61–73]. Some of the main conclusions from
suggesting that those bonds are weakened by an elec- those studies are summarized below.
tronic interaction between the alkane and the surface. As mentioned in the previous section, alkanes are
In fact, infrared C–H stretching soft modes are some- fairly unreactive, and usually display significantly
times observed in adsorbed alkanes [57]. Also, alkane low reactive sticking probabilities when dosed on
activation is facilitated by electron deficiency at the metal surfaces. This makes the production of alkyl
metal center: early transition metals are particularly surface species directly via activation of saturated
efficient in dissociating medium-size alkane chains hydrocarbons under controlled vacuum conditions
[58], and iridium activates both methane and ethane virtually impossible. In addition, because of the se-
more efficiently than platinum [59]. Rough surfaces vere conditions often required for the initial scission
are in general more active than close-packed ones of the C–H bonds in alkanes, those reactants tend
too [59]. Finally, co-adsorbed additives such as alkali to decompose extensively on metal surfaces once
metals or chalcogens can promote alkane decomposi- dissociatively adsorbed. In order to address these
tion on nominally inert substrates [60]. difficulties, a number of alternative procedures have
All these studies support the basic idea that the been developed over the years for the preparation of
overall rate of reforming processes may be controlled hydrocarbon surface intermediates in general, and for
by the initial activation of the reactants on the surface the deposition of alkyl surface moieties in particular,
F. Zaera / Applied Catalysis A: General 229 (2002) 75–91 79

Fig. 2. Potential elementary steps available to alkyl moieties (sec-butyl in this example) when chemisorbed on metal surfaces. All these
reactions, with the exception of ␤-methyl elimination, have been seen in surface science studies [66,70,74,100,183]. However, it is mostly
the regioselectivity of the hydride elimination steps what defines selectivity in reforming.

on clean single-crystal surfaces [66,74]. Fig. 3 depicts the point, alkyl halides have also been shown to un-
schematically some of the methods reported for the dergo oxidative addition reactions in solution to form
production of surface methyl groups. They include organometallic products with alkyl ligands attached
dosing surfaces with gas radicals produced by py- directly to the metal center [80,81]. We have shown
rolysis of appropriate precursors (azomethane in the that the cleavage of the carbon–halogen bonds in alkyl
case of methyl groups) [75], activating of adsorbed halides adsorbed on metal surfaces requires activation
alkanes via hyperthermal atom [51] or electron [76] energies of 5 kcal/mol or less, about the same as in so-
bombardment, soft-landing gas-phase cations [77], lution [82], and that that reaction usually occurs at tem-
and photoactivating of adsorbed precursors [78]. Un- peratures below 200 K [83,84]. The carbon–halogen
fortunately, it has not been possible to extend any bond scission is easier with heavy halogens, hence the
of those procedures to the general study of alkyl preference for using alkyl iodides and alkyl bromides
chemistry. By far, the most popular method for the precursors.
preparation of surface alkyls under vacuum nowadays We have also proven that the initial decomposition
is the thermal activation of adsorbed alkyl halides, an of alkyl halides on metals leads to the formation of
approach that we pioneered in the late 1980s [61,66]. chemisorbed alkyl moieties. Examples of the data sup-
It is well known that alkyl radicals can be pre- porting this conclusion are shown in Fig. 4. The alkyl
pared efficiently in gas or liquid phases via the py- chains initially orient themselves flat on the surface
rolysis of suitable precursors such as alkyl halides, in order to maximize their interaction with the metal,
azo compounds, and peroxides [79]. Perhaps more to but reorient to a perpendicular configuration at higher
80 F. Zaera / Applied Catalysis A: General 229 (2002) 75–91

Fig. 3. Schematic representation of some of the methods reported in the literature for the preparation of methyl surface groups on
solid surfaces [52,66,74]. From left to right: deposition of gas methyl radicals prepared by azomethane pyrolysis [75], activation of
adsorbed methane by hyperthermal atoms [51], soft-landing of gas-phase methyl cations [77], electron activation of adsorbed methane
[76], photoactivation of adsorbed methyl bromide [78], and thermal activation of methyl iodide [63,65]. From these, the latter is by far
the most versatile and widely used procedure for the isolation of hydrocarbon intermediates on metals.

coverages, presumably because of the prevalence of contain low-energy C–N bonds and are used exten-
van der Waals interactions among the hydrocarbon sively as precursors in gas and liquid phase studies,
chains in that regime [67,72,85]. Unfortunately, there but display fairly complex surface chemistry which
is no direct information on either the bonding geom- do not lead to the clean production of adsorbed alkyl
etry or the coordination of alkyl groups to the surface groups [107,108].
available to date. Both the low metal–carbon stretch-
ing frequency observed in vibrational studies [38,86]
and some theoretical calculations [87,88] suggest that 4. ␤-hydride elimination: olefin formation
alkyls may adsorb on hollow sites, but other computer
estimates explain the experimental results by an on-top The thermal chemistry of alkyl groups on transition
configuration instead [89]. metal surfaces is dominated by the elimination of a
The activation of chemisorbed halohydrocarbon hydrogen atom from the ␤-position, that is, the carbon
precursors has been successfully extended to the pro- adjacent to that bonded to the surface [52,69,74,109].
duction of surface carbenes [90–93], vinyls [94,95], One of the earliest indications of the occurrence of
allyls [96–99], metallacycles [100–104], and oxamet- this step came from temperature-programmed desorp-
allacycles [105,106]. It should be pointed out, how- tion studies on the decomposition of partially labeled
ever, that ease of activation of potential precursors in ethyl iodides on Pt(1 1 1) [110]. That work showed that
the gas or liquid phases does not guarantee success the dehydrogenation of CH3 CD2 I only removes nor-
on surfaces. For instance, azo and diazo compounds mal hydrogen atoms, while the thermal activation of
F. Zaera / Applied Catalysis A: General 229 (2002) 75–91 81

Fig. 4. Evidence for the production of alkyl groups on metal surfaces by thermal activation of chemisorbed alkyl iodides. Left:
reflection-absorption infrared spectra (RAIRS) from methyl, ethyl, 1-propyl, and 2-propyl moieties on Cu(1 1 0), prepared by annealing of
the corresponding adsorbed alkyl iodides at 180 K [69]. Right: static secondary ion mass spectra (SSIMS) from perdeutero methyl and
ethyl fragments chemisorbed on Ni(1 0 0), again prepared by thermal activation of the appropriate alkyl iodide precursors [65,68,135].

CD3 CH2 I leads to the exclusive release of a deuterium was also proven by the results from experiments
atom instead. More direct evidence is now available with 2-iodohexane on Cu(1 0 0), where the selective
for this reaction in the form of the detection of the production of 2-hexene over 1-hexene was observed
resulting olefin, both adsorbed on the surface (Fig. 5) [73,112]. The activation barrier for ␤-elimination
and in the gas phase after desorption. from chemisorbed alkyl groups was determined to be
In terms of the details of the transition state during less than 10 kcal/mol on most transition-metal sur-
the ␤-hydride elimination step, an elegant series of ex- faces [70,113], and to display a large normal kinetic
periments using fluorine-substituted propyl groups on isotope effect, on the order of a factor of 10 when sub-
Cu(1 1 1) was used to demonstrate the anionic nature stituting normal hydrogen by deuterium atoms [114].
of the leaving hydrogen [111]. The fact that the posi- This implies that either tunneling or secondary effects
tive charge that develops on the ␤-carbon in the tran- must play a role in the dehydrogenation step. Some
sition state is stabilized by electron donating groups other significant trends have been observed for the
82 F. Zaera / Applied Catalysis A: General 229 (2002) 75–91

␤-H elimination in organometallic complexes. (4) ␤-H


elimination is easier on earlier and lighter transition
metals: it occurs below 200 K on nickel, palladium,
and platinum surfaces, at around 230–250 K on cop-
per, and at about 260–310 K on gold. (5) Most metals
are efficient at promoting ␤-hydride elimination reac-
tions, but a few late transition metals tend to favor the
reductive elimination route instead [115,116]. On sil-
ver in particular, alkyl groups couple with 100% effi-
ciency [117], perhaps because the low affinity of silver
for hydrogen inhibits any hydride elimination steps.
␤-hydride elimination explicitly explains olefin
production during reforming. This is a facile reaction,
commonly leading to a rapid alkane–alkene equilib-
rium [13,118]. Together with the reverse C=C insertion
into metal–hydrogen bonds (the half hydrogenation
of adsorbed olefins to alkyls) and the reductive elim-
ination between alkyl and hydrogen adsorbates (the
second hydrogenation of alkyls to alkanes), it also
accounts for other catalytic processes such as H–D
exchange and double bond migration [71,90,96,99,
100,118–130]. H–D isotope exchange in particu-
lar may not be particularly important in practical
catalysis, but plays a central role in mechanis-
tic studies of hydrocarbon conversion processes
Fig. 5. RAIRS evidence for the formation and subsequent ther- [12,13,21,131–133], and has provided a conve-
mal chemistry of propene from ␤-hydride elimination in 1-propyl
groups adsorbed on Pt(1 1 1) surfaces [71]. The bottom three traces
nient way to estimate metal–carbon bond energies
correspond to the progression of the intermediates that form dur- in surface-science studies [52,66,74,113,123,126,
ing the thermal activation of propyl surface species, propylene first 134,135]. It has also allowed for the calculation of
and propylidyne above room temperature. The spectra for low and relative rates between hydrogenation and dehydro-
high propylene coverages are also provided to illustrate the possi- genation steps [100].
bility of formation of two types, di-␴- and ␲-bonded of adsorbed
olefins [177].
It is worth emphasizing here that selectivity in hy-
drocarbon reforming is frequently controlled by the
relative rates among a number of alternative path-
␤-hydride elimination reaction, as illustrated by the ways. In particular, the nature of the final products is
data in Fig. 6 [52,70,74,112]: (1) The desorption tem- typically determined by competition among different
perature of the alkenes produced by alkyl dehydro- hydrogenation and dehydrogenation steps. It is here
genation shifts to higher temperature with increasing where seemingly simple reactions such as isotope ex-
chain length, perhaps because of the increase in ad- change can help. It has consistently been reported that
sorption energy of the products. (2) The alkene yield H–D scrambling in alkanes and alkenes, which takes
is higher (relative to alkane production by the com- place via the interconversion of alkyls and olefins,
peting reductive elimination reaction) for branched is much more extensive on platinum than on nickel
than for linear alkyls. Also, the products desorb at [12,13,132]. Given that the rate of ␤-hydride elimi-
lower temperatures in the former cases. These effects nation, although somewhat dependent on the nature
can be explained at least in part by steric and entropic of the catalyst, is quite facile on most late transition
arguments. (3) Open surfaces are normally more ac- metals [52,61,67,68,71,109,113,118,129,130,136], it
tive than close-packed ones, a difference that may be must be concluded that the final exchange product
correlated to the requirement of low coordination for distribution is determined by the relative rates of
F. Zaera / Applied Catalysis A: General 229 (2002) 75–91 83

Fig. 6. Dependence of surface alkyl dehydrogenation rates on the nature of the surface (left) and of the reactants (right) [52,66,74]. Left:
temperature maxima for methane (clear bars) and ethylene (filled bars) desorption as a function of metal and crystallographic orientation.
Those data provide an indication of the relative reactivity for ␣- and ␤-hydride eliminations, respectively, across the series of surfaces
reported. Right: TPD temperature maxima for the production of olefins from surface alkyls as a function of hydrocarbon chain length and
degree of branching. Data are shown for reactions on Ni(1 0 0) (filled bars) [70] and Cu(1 0 0) (clear bars) [112]. In general, increases in
reactivity are seen with early transition metals and with decreasing hydrocarbon chain lengths.

olefin re-hydrogenation versus further decomposition. dehydrogenation steps are required to explain the more
Therefore, it is the particular ease with which nickel demanding reforming processes. A number of studies
promotes further dehydrogenation steps what limits have been carried out by us and others to tackle this
H–D exchange on that metal. Conversion on platinum, issue. The results from that work lead us to believe
on the other hand, yields extensive H–D scrambling, that the selectivity in reforming processes is defined
as manifested not only by the high degree of deu- by the regiospecificity of the removal of the next hy-
terium substitution in the final products, but also by drogen atom from alkyl surface intermediates. While
the low temperatures at which the reaction proceeds dehydrogenation at the ␣-position is likely to lead to
[113,130]. Fig. 7 displays laser-induced desorption eventual C–C bond scissions [15], reactivity at the
data for the case of ethylene exchange with adsorbed ␥-carbon may be responsible for isomerization and cy-
deuterium on a Pt(1 1 1) surface [129]. It is shown clization steps [20,137,138].
there that extensive exchange can be induced at tem- Hydride elimination from the ␣- and ␥-position are
peratures as low as 215 K. Similar arguments also generally much less favorable than from the ␤-carbon,
apply to exchange at the ␣- and ␥-position (see later). but can still be probed by using adsorbates with no
␤-hydrogens such as methyl, neopentyl, and benzyl
moieties. The simplest case of ␣-H elimination is the
5. ␣- and ␥-hydride eliminations: prelude to conversion of methyl to methylene groups, a reaction
hydrogenolysis and isomerization first established on Pt(1 1 1) by reflection-absorption
infrared spectroscopy [63]. The multiple H–D ex-
␤-hydride elimination may be the dominant step change seen for methyl groups on Pt(1 1 1) suggests
in the conversion of alkyl surface species, but other an activation barrier of less than 10 kcal/mol for the
84 F. Zaera / Applied Catalysis A: General 229 (2002) 75–91

but of only about a factor of five around 600 K. Again,


this works out to a barrier difference of approximately
2 ± 1 kcal/mol.
Hydrogenolysis after ␣-hydride elimination has
been corroborated in the case of neopentyl iodide con-
version on Ni(1 0 0) [140]. In that system, the initial
elimination of one of the a hydrogen atoms from the
neopentyl moiety occurs below 200 K. The resulting
neopentylidene intermediate is quite stable, react-
ing further only above 350 K to ultimately produce
isobutene. The mechanism by which neopentylidene
decomposes to isobutene appears to involve the con-
certed scission of the C␣ –C␤ bond in conjunction with
the elimination of hydrogen atoms from the ␣- and
␥-positions. The interesting observation here in terms
of selectivity in reforming is that the hydrogenolysis
of alkyl intermediates may be rate limited by the C–C
bond breaking step, but is decided by a much earlier
and facile dehydrogenation reaction.
As stated above, elimination at positions fur-
Fig. 7. Typical mass spectra from laser-induced thermal desorp-
tion/Fourier transform mass spectrometry (LITD/FTMS) experi- ther along the carbon chain, whenever possible, are
ments designed to characterize the kinetics of H–D exchange re- believed to yield the cyclic products proposed in iso-
actions between C2 D4 and H atoms on Pt(1 1 1) [129]. In this merization reactions (Fig. 8). The selectivity between
case, the adsorption of 0.13 ML of C2 D4 was followed by a dos- ␥- and ␣-hydride elimination depends on the nature
ing of 20 L of H2 , and the reaction was carried out at 225 K. The
of the metal surface: both steps are possible on plat-
exchange is manifested by both a decrease in signal intensity for
32 amu (C2 D4 ) and a simultaneous growth of the 31 (C2 D3 H) inum [141], but only the latter is observed on nickel
and 29 and 27 (multiply-exchanged C2 X4 ) amu peaks during the [140]. This trend explains the unique ability of plat-
course of the reaction. inum for catalyzing reforming processes, as opposed
to nickel, which leads to exclusive cracking instead
[100]. It should be emphasized again that it is the
initial a dehydrogenation from those moieties [122]. relative rates among the different available paths on
In the case of the thermal activation of methyl iodide a given surface what matters in terms of defining se-
on either Ni(1 0 0) or Ni(1 1 1) surfaces, experiments lectivity. All dehydrogenation steps follow the same
both with labeled methyl iodide and with methylene qualitative trends, that is, they become faster on early
iodides co-adsorbed with hydrogen provided a lower transition metals or with heavier hydrocarbon chains.
limit for the temperature of the ␣-hydride elimination Some examples of this are provided for ␣- and
reaction in the conversion of methyl to methylidyne at ␤-hydride elimination reactions in Fig. 6. However,
above 200 K [90], over 40 K higher than that needed the relative rates for ␣-, ␤-, and ␥-dehydrogenation
for ␤-hydride elimination from ethyl groups. This within the same metal also change across the periodic
represents a difference of about 2 kcal/mol in activa- table, because the magnitude of the variations in rate
tion energy, or about a 25% increase in barrier height for each of those steps can be quite different on differ-
[65,124,135]. Additional comparisons between ␣- ent metals. Notice, for instance, the large difference
and ␤-hydride elimination have been achieved indi- in maximum temperature in temperature programmed
rectly by isotope scrambling during high-temperature desorption (TPD) experiments for ␣ (>450 K) versus
ethylidyne formation [139] and by catalytic studies ␤ (∼250 K) eliminations on copper reported in Fig. 6.
with ethane on platinum [127]. Those experiments A significantly smaller difference between the two
indicated a difference in rate between the two steps (<100 K) is seen on platinum, palladium, or nickel
of over four orders of magnitude at low temperatures, crystals. It is also for this reason that while only
F. Zaera / Applied Catalysis A: General 229 (2002) 75–91 85

Fig. 8. Schematic representation of the elementary surface reaction steps proposed for the hydrogenolysis and isomerization of hydrocarbons
during catalytic reforming [15,20]. The example provided here corresponds to neopentyl intermediates. The arguments is made that while
␣-hydride elimination eventually leads to C–C bond scission and yields hydrogenolysis products, hydride elimination at the ␥-position
results in isomerization and cyclization instead. The steps identified in studies with Pt and Ni single crystal surfaces are indicated in the
figure as well [100,140,141].

␣-hydride elimination can be detected from neopentyl section, we summarize what we have learned about
groups on Ni(1 0 0), both ␣- and ␥-hydride elimina- reductive eliminations and coupling steps in the last
tions display comparable rates on Pt(1 1 1). The same decade.
argument holds true when contrasting hydrogenation Hydrogenation reactions, the reverse of C–H bond
versus dehydrogenation steps, as mentioned in the scissions, are possible under vacuum as long as the
previous discussion on isotope exchange. surfaces are predosed with large coverages of hy-
drogen [118]. Indeed, most olefins are capable of
hydrogenating to their corresponding alkanes dur-
6. Bond forming steps ing temperature-programmed desorption experiments
[118,120,142]. Extensive information has been ob-
So far in this review we have addressed dehy- tained on the mechanism of that reaction by a com-
drogenation reactions, that is, C–H bond-breaking bination of experiments with deuterium isotopic
steps. Reforming catalysis also involves bond-forming labeling and with alkyl and other potential interme-
processes. In general, though, those are much more diates [113,126,143]. It has become quite clear that
difficult to emulate under vacuum conditions. In this alkenes hydrogenate in a stepwise manner, via the
86 F. Zaera / Applied Catalysis A: General 229 (2002) 75–91

slow formation of an alkyl intermediate [119,139]. In methyl moieties yields ethylidyne moieties on Pt(1 1 1)
analogy with organometallic chemistry, the first step, [153] and Ru(0 0 0 1) [154] surfaces only at high cover-
the hydrogenation of the olefin to an alkyl moiety, ages. It is also conceivable that, at least in some cases,
can be described as the insertion of an olefin into the actual C–C surface recombination step is fast, and
a metal–hydrogen bond, while the second, the con- that the rate for alkyl dimerization is controlled by the
version of the alkyl to an alkane, corresponds to the diffusion of the reactants on the surface.
reductive elimination of alkyl groups with hydrogen Thermal activation of surface metallacycles can also
atoms. In the case of platinum, the reversibility of the lead to a C–C coupling, in that case to eliminate cyclic
first step (the ␤-hydride elimination from alkyls dis- hydrocarbons. Such steps have been reported on sil-
cussed before [61,66]) leads to the rapid interchange ver [117] and nickel [102,103], but were not seen on
of hydrogen atoms within the adsorbed hydrocarbons. aluminum [155], palladium [156], or platinum [104].
Therefore, when deuterium is used instead of nor- The particular fact that cyclopropane is produced from
mal hydrogen in TPD experiments, extensive H–D thermal activation of 1,3-diiodopropane on Ni(1 0 0)
exchange is observed [119,120]. [102] but not on Pt(1 1 1) [102] is somewhat odd,
Similar conclusions can also be reached in refer- since platinum is a better cyclization catalyst than
ence to alkanes incapable of direct dehydrogenation nickel. However, benzene formation from nickellacy-
to olefins such as methane or neopentane. In those cloheptane intermediates appears to require an exten-
cases, the relative rates of hydrogenation versus de- sive prior dehydrogenation of those surface moieties
hydrogenation and the selectivity in dehydrogenation [103]; again, nickel is a particularly able surface when
from the different positions still control both the type it comes to dehydrogenation steps, ␣-H elimination in
of intermediates that forms on the surface [100] and particular. Perhaps the most interesting C–C coupling
the extent of the H–D exchange [90,121,122]. For in- reaction on single-crystal surfaces reported so far is the
stance, neopentyl groups dehydrogenate exclusively trimerization of acetylene on Pd(1 1 1) [157], a reac-
to neopentylidene on Ni(1 0 0) [140,144] but are able tion takes place via an initial oxidative coupling of two
to also loose hydrogen atoms from the ␥ position acetylene molecules to form a Pd–C4 H4 metallacycle
(and presumably form a metallacyclic surface moi- [158]. Acetylene trimerization has also been observed
ety) on Pt(1 1 1) [141,145], and the re-hydrogenation on nickel [159], copper [160], and tin-modified plat-
of those intermediates is easier in relative terms on inum [161] surfaces, but no mechanistic details were
platinum than on nickel. Consequently, more exten- provided in the reports of those cases.
sive exchange is seen for neopentane on the former In terms of relative rates for C–H versus C–C
metal [146–148]. bond-forming reactions, it has been reported that
There have been only a handful of clean examples co-adsorption of hydrogen with phenyl groups on
for C–C bond forming steps under vacuum. Methylene Cu(1 1 1) completely inhibits biphenyl formation at
insertion and alkyl coupling steps have been identified the expense of benzene production [162]; C–H reduc-
on late transition metals, most noticeably on coinage tive elimination steps are clearly favored in that case.
metals [52,74,115,149]. Alkyl coupling takes place We believe that this conclusion may be quite general,
with 100% selectivity on silver [117], but competes applying to most metals relevant to reforming. In ad-
with hydride elimination steps, especially at low alkyl dition, phenyl groups act as efficient traps for alkyl
coverages, on palladium [150], copper [73,74,115], radical intermediates, an indication that alkyl–aryl
and gold [116,151]. It appears that the selectivity to- coupling is favored over alkyl–alkyl formation [163].
wards reductive elimination reactions may at least in
part be determined by the ability of the metal to acti-
vate dehydrogenation reactions: only when dehydro- 7. Added complications from the use of
genation is slow it is possible for the metal to promote atmospheric pressures in catalysis
such bond forming steps [152]. Also, coupling reac-
tions are bimolecular, and therefore display kinetics An additional level of complication is added to
strongly dependent on the concentrations of the sur- the study of hydrocarbon conversion when atmo-
face reactants. For instance, the thermal activation of spheric pressures are considered. This is so because
F. Zaera / Applied Catalysis A: General 229 (2002) 75–91 87

hydrocarbon catalysis does not occur on clean metal hydrogenation and dehydrogenation processes. TPD
surfaces, but rather on metals covered with strongly and 14 C isotope labeling experiments have indicated
bonded carbonaceous residues [11,164]. In fact, the that their residence time and composition depend
initial dehydrogenation steps discussed above take on temperature, becoming more hydrogen deficient
place at quite low temperatures, at least on clean metal and strongly held with increasing temperature, but
surfaces, and are followed by additional decomposi- that they always turn over at rates orders of mag-
tion reactions. In particular, the olefins produced via nitude slower than those of the catalytic conversion
␤-hydride elimination react around room temperature [118,174,175]. Indirect evidence suggests that similar
to yield alkylidyne species (Fig. 5). The mechanism species may also be present on the surface during
for this reaction has been highly contested and is more demanding hydrogenolysis and reforming pro-
not yet fully understood [165–168], but appears to cesses, but that they may be more directly involved
involve an alkylidene intermediate [139,169,170]. in the mechanism of those reactions [127,176].
Regardless, the same alkylidyne surface moieties ap- The net result from all of this is that clean metal
pear to form during the catalytic hydrogenation of surfaces become covered with carbonaceous species
olefins [12,118,171–173]. right at the beginning of most reforming reactions,
In retrospect, it is easy to understand why such a changing the chemisorption characteristics of the
hydrocarbon layer is needed to passivate the surface; reactants. In order to understand the hydrogena-
otherwise, dehydrogenation steps would dominate the tion of unsaturated hydrocarbons in particular, olefin
conversion of hydrocarbons even on platinum sur- ␲-bonding needs to be considered in the mechanistic
faces. It has also become clear that strongly bonded picture. Such species have indeed been identified by
carbonaceous species only play an indirect role in vibrational spectra both in vacuum [177] and in situ

Fig. 9. Left: typical isothermal kinetic raw data obtained for the conversion of ethylene on Pt(1 1 1) at 270 K by using a molecular beam
technique [128]. The rates of ethylene consumption, hydrogenation to ethane and ethylidyne formation were measured as a function of time
by following the mass spectroscopy signals for 25, 30, and 2 amu, respectively. The initial uptake of ethylene is followed by an additional
small reversible adsorption of weakly-bonded olefins, as manifested by the spikes observed right after blocking (flag up) and unblocking
(flag down) of the beam. Right: dependence of the initial rate of ethane formation on the coverage of weakly-adsorbed ␲-ethylene. The
solid line through the data points, a polynomial fit corresponding to a kinetic order of 1.2 ± 0.3, indicated that it is the weakly adsorbed
ethylene the one involved in the hydrogenation reaction.
88 F. Zaera / Applied Catalysis A: General 229 (2002) 75–91

under reaction conditions [178–181]. Infrared spectra discussed above where the surface is mostly covered
for both (strongly bonded) di-␴- and (weakly bonded) with three-dimensional carbonaceous islands but still
␲-propylene on Pt(1 1 1) are shown in Fig. 5. Addi- retains some patches of uncovered bare metal [164].
tional molecular beam kinetic work on the ethylene The initial and final hydrocarbon hydrogenation and
hydrogenation/Pt(1 1 1) system has revealed that: (1) dehydrogenation steps are likely to involve weakly
weakly-bonded ␲-olefins are indeed the ones that adsorbed species, ␲-bonded in the case of olefins, and
intervene directly in the hydrogenation process; (2) to be aided by the strongly-bonded deposits. Those
the dissociative adsorption of hydrogen on the sur- moieties not only moderate the dehydrogenating high
face is rate-limiting for the overall reaction; and (3) activity of the clean metal surface, but also act as a
strongly-adsorbed alkenes and other carbonaceous reservoir of hydrogen atoms [11]. The more demand-
deposits inhibit the latter step [128]. Key kinetic data ing intermediate processes, on the other hand, are
for this ethylene/Pt(1 1 1) system are shown in Fig. 9. likely to occur directly on the exposed metal sites.
Hydrogen also plays a unique role in hydrocar- There, the selectivity between the C–C scission steps
bon conversion processes. It has been long recognized that lead to undesirable hydrogenolysis and the skele-
that one of the key functions of metals as catalysts tal rearrangements involved in reforming is defined
in hydrogenation–dehydrogenation and skeletal rear- by the relative rates of the initial dehydrogenation
rangement reactions is to activate molecular hydrogen of alkyl intermediates between the ␣-position versus
to produce chemisorbed hydrogen atoms [13]. How- ␥-position.
ever, the exact manner in which the resulting atomic
hydrogen participates in the hydrocarbon conversion
steps remains unsettled. In fact, it is quite possible that Acknowledgements
hydrogen transfer from the metal to the reactants may
be mediated by the surface carbonaceous deposits. Financial support for this research was provided by
What is clear is that hydrogen pressures affect the se- a grant from the National Science Foundation, Chem-
lectivity of many reactions, as well as hydrogenation istry Division (CHE-9819652).
versus dehydrogenation and more complex reforming
networks, in a significant way [12]. For instance, it has
been shown that the selectivity during the conversion References
of cyclohexene on Pt(1 1 1) changes dramatically with
pressure, with the benzene (dehydrogenation) to cyclo- [1] G.A. Somorjai, Science 227 (1985) 902.
hexane (hydrogenation) production ratio going from [2] B.C. Gates, J.R. Katzer, G.C.A. Schuit, Chemistry of
about 100:1 below 10−7 Torr total pressure to 1:100 at Catalytic Processes, McGraw-Hill, New York, 1979.
10 Torr [182]. In that case, it was argued that this selec- [3] H.H. Voge, in: P.H. Emmett (Ed.), Catalysis, Vol. VI,
Reinhold, New York, 1958, p. 407.
tivity change is a reflection of an approximately con- [4] P.J. Flory, Principles of Polymer Chemistry, Cornell
stant reaction probability with changing pressure for University Press, Ithaca, 1953.
hydrogenation, in contrast to the strong dependence [5] F.G. Ciapetta, R.M. Dobres, R.W. Baker, in: P.H. Emmett
of dehydrogenation rates on pressure due to the fact (Ed.), Catalysis, Vol. VI, Reinhold, New York, 1958, p. 495.
that those reactions require large surface ensembles [6] V. Haensel, in: B.T. Brooks, C.E. Boord, S.S. Kurtz,
L. Schmerlin (Eds.), The Chemistry of Petroleum
of empty sites which become blocked by other adsor-
Hydrocarbons, Vol. II, Reinhold, New York, 1955, p. 189.
bates at high pressures. Hydrogen adsorption may also [7] J.H. Sinfelt, Bimetallic Catalysts: Discoveries, Concepts and
be inhibited by competition with gas-phase hydrocar- Applications, Wiley, New York, 1983.
bons, hence the often negative dependence of reaction [8] G.A. Mills, H. Heinemann, T.H. Milliken, A.G. Oblad, Ind.
rates on hydrocarbon partial pressures [12,118]. Eng. Chem. 45 (1953) 134.
[9] J.H. Sinfelt, in: J.R. Anderson, M. Boudart (Eds.),
Catalysis—Science and Technology, Vol. 1, Springer, Berlin,
8. Conclusions 1981, p. 257.
[10] H. Pines, The Chemistry of Catalytic Hydrocarbon
A molecular model of the working reforming cat- Conversions, Academic Press, New York, 1981.
alyst has been advanced based on the information [11] G.A. Somorjai, F. Zaera, J. Phys. Chem. 86 (1982) 3070.
F. Zaera / Applied Catalysis A: General 229 (2002) 75–91 89

[12] F. Zaera, G.A. Somorjai, in: Z. Paál, P.G. Menon (Eds.), [43] H. Yang, J.L. Whitten, J. Chem. Phys. 96 (1992) 5529.
Hydrogen Effects in Catalysis: Fundamentals and Practical [44] V.A. Ukraintsev, I. Harrison, J. Chem. Phys. 101 (1994)
Applications, Marcel Dekker, New York, 1988, p. 425. 1564.
[13] G.C. Bond, Catalysis by Metals, Academic Press, London, [45] L. Hanley, Z. Xu, J.T. Yates Jr., Surf. Sci. 248 (1991) L265.
1962. [46] C.B. Mullins, W.H. Weinberg, J. Chem. Phys. 92 (1990)
[14] J. Horiuti, K. Miyahara, Hydrogenation of Ethylene on 3986.
Metallic Catalysts, National Bureau of Standards, Report [47] A.C. Luntz, D.S. Bethune, J. Chem. Phys. 90 (1989) 1274.
NSRDS-NBC no. 13, 1968. [48] P.D. Szuromi, J.R. Engstrom, W.H. Weinberg, J. Phys. Chem.
[15] C. Kemball, Catal. Rev. 5 (1971) 33. 89 (1985) 2497.
[16] J.R. Anderson, Adv. Catal. 23 (1973) 1. [49] D.P. Land, C.L. Pettiette-Hall, R.T. McIver, J.C. Hemminger
[17] J.K.A. Clarke, J.J. Rooney, Adv. Catal. 25 (1976) 125. Jr., J. Am. Chem. Soc. 111 (1989) 5970.
[18] Z. Paál, Adv. Catal. 29 (1980) 273. [50] J.F. Weaver, M.A. Krzyzowski, R.J. Madix, J. Chem. Phys.
[19] J.H. Sinfelt, Adv. Catal. 23 (1980) 91.
112 (2000) 396.
[20] F.G. Gault, Adv. Catal. 30 (1981) 1.
[51] J.D. Beckerle, A.D. Johnson, Q.Y. Yang, S.T. Ceyer, J.
[21] L. Guczi, P. Tétényi, Acta Chim. Acad. Sci. Hung. 71 (1972)
Chem. Phys. 91 (1989) 5756.
341.
[52] F. Zaera, Chem. Rev. 95 (1995) 2651.
[22] L. Guczi, P. Tetenyi, Ann. New York Acad. Sci. 213 (1973)
[53] H. Burghgraef, A.P.J. Jansen, R.A. van Santen, J. Chem.
173.
[23] Z. Karpinski, L. Guczi, J. Chem. Soc., Chem. Commun. Phys. 101 (1994) 11012.
(1977) 563. [54] F.C. Schouten, O.L.J. Gijzeman, G.A. Bootsma, Surf. Sci.
[24] P. Tétényi, L. Guczi, A. Sárkány, Acta Chim. Acad. Sci. 87 (1979) 1.
Hung. 97 (1978) 221. [55] S.G. Brass, G. Ehrlich, J. Chem. Phys. 87 (1987) 4285.
[25] L. Guczi, A. Frennet, V. Ponec, Acta Chim. Hung. 112 [56] T.P. Beebe Jr., D.W. Goodman, B.D. Kay, J.T. Yates Jr.,
(1983) 127. J. Chem. Phys. 87 (1987) 2305.
[26] L. Guczi, J. Mol. Catal. 25 (1984) 13. [57] J.E. Demuth, H. Ibach, S. Lehwald, Phys. Rev. Lett. 40
[27] G. Diaz, P. Esteban, L. Guczi, F. Garin, P. Bernhardt, (1978) 1044.
J.L. Schmitt, G. Maire, in: C. Morterra, A. Zecchina, G. [58] A.C. Liu, C.M. Friend, J. Chem. Phys. 87 (1987) 4975.
Costa (Eds.), Proceedings of a European Conference on [59] W.H. Weinberg, Langmuir 9 (1993) 655.
Studies in Surface Science and Catalysis Series, Structure [60] J.G. Chen, M.D. Weisel, J.H. Hardenbergh, F.M. Hoffmann,
and Reactivity of Surfaces, Vol. 48, Trieste, Italy, 13–16 C.A. Mims, R.B. Hall, J. Vac. Sci. Technol. A 9 (1991)
September 1988, Elsevier, Amsterdam, 1989, p. 363. 1684.
[28] L. Guczi, T. Hoffer, Z. Zsoldos, S. Zyade, G. Maire, F. [61] F. Zaera, Surf. Sci. 219 (1989) 453.
Garin, J. Phys. Chem. 95 (1991) 802. [62] F. Zaera, H. Hoffmann, P.R. Griffiths, Vacuum 41 (1990)
[29] R.B. Anderson, P.T. Dawson (Eds.), Experimental Methods 735.
in Catalytic Research, Vol. III, Characterization of Surface [63] F. Zaera, H. Hoffmann, J. Phys. Chem. 95 (1991) 6297.
and Adsorbed Species, Academic Press, New York, 1976. [64] H. Hoffmann, P.R. Griffiths, F. Zaera, Surf. Sci. 262 (1992)
[30] J.W. Niemantsverdriet, Spectroscopy in Catalysis: An 141.
Introduction, VCH, Weinheim, 1993. [65] S. Tjandra, F. Zaera, Langmuir 8 (1992) 2090.
[31] G. Ertl, J. Küppers, Low Energy Electrons and Surface [66] F. Zaera, Acc. Chem. Res. 25 (1992) 260.
Chemistry, VCH, Weinheim, 1985. [67] C.J. Jenks, B.E. Bent, N. Bernstein, F. Zaera, J. Am. Chem.
[32] D.P. Woodruff, T.A. Delchar, Modern Techniques of Soc. 115 (1993) 308.
Surface Science, 2nd Edition, Cambridge University Press, [68] S. Tjandra, F. Zaera, Surf. Sci. 289 (1993) 255.
Cambridge, 1994.
[69] F. Zaera, J. Mol. Catal. 86 (1994) 221.
[33] C.B. Duke (Ed.), Surface Science: The First Thirty Years,
[70] S. Tjandra, F. Zaera, J. Am. Chem. Soc. 117 (1995) 9749.
North-Holland, Amsterdam, 1994.
[71] D. Chrysostomou, C. French, F. Zaera, Catal. Lett. 69 (2000)
[34] G.A. Somorjai, Acc. Chem. Res. 9 (1976) 248.
117.
[35] G.A. Somorjai, Science 201 (1978) 489.
[36] F. Zaera, A.J. Gellman, G.A. Somorjai, Acc. Chem. Res. 19 [72] C.J. Jenks, B.E. Bent, N. Bernstein, F. Zaera, J. Phys. Chem.
(1986) 24. B 104 (2000) 3008.
[37] C.T. Rettner, H.E. Pfnür, D.J. Auerbach, Phys. Rev. Lett. 54 [73] C.J. Jenks, B.E. Bent, F. Zaera, J. Phys. Chem. B 104 (2000)
(1985) 2716. 3017.
[38] S.T. Ceyer, J.D. Beckerle, M.B. Lee, S.L. Tang, Q.Y. Yang, [74] B.E. Bent, Chem. Rev. 96 (1996) 1361.
M.A. Hines, J. Vac. Sci. Technol. A 5 (1987) 501. [75] X.D. Peng, R. Viswanathan, G.H. Smudde, P.C. Stair Jr.,
[39] A.V. Hamza, R.J. Madix, Surf. Sci. 179 (1987) 25. Rev. Sci. Instrum. 63 (1992) 3930.
[40] C.N. Stewart, G. Ehrlich, J. Chem. Phys. 62 (1975) 4672. [76] J.M. White, Langmuir 10 (1994) 3946.
[41] M.C. McMaster, R.J. Madix, Surf. Sci. 275 (1992) 265. [77] D.R. Strongin, J.K. Mowlem, K.G. Lynn, Y. Kong, Rev. Sci.
[42] R.W. Verhoef, D. Kelly, C.B. Mullins, W.H. Weinberg, Surf. Instrum. 63 (1992) 175.
Sci. 311 (1994) 196. [78] J.M. White, J. Mol. Catal. A 131 (1998) 71.
90 F. Zaera / Applied Catalysis A: General 229 (2002) 75–91

[79] M. Lazár, J. Rychly, V. Klimo, P. Pelikán, L. Valko, Free [113] F. Zaera, J. Phys. Chem. 94 (1990) 8350.
Radicals in Chemistry and Biology, CRC Press, Boca Raton, [114] C.J. Jenks, M. Xi, M.X. Yang, B.E. Bent, J. Phys. Chem.
1989. 98 (1994) 2152.
[80] M.P. Brown, A. Yavari, R.H. Hill, R.J. Puddephatt, J. Chem. [115] C.-M. Chiang, T.H. Wentzlaff, B.E. Bent, J. Phys. Chem.
Soc. Dalton Trans. (1985) 2421. 96 (1992) 1836.
[81] J.P. Collman, L.S. Hegedus, J.R. Norton, R.G. Finke, [116] A. Paul, M.X. Yang, B.E. Bent, Surf. Sci. 297 (1993) 327.
Principles and Applications of Organotransition Metal [117] X.-L. Zhou, J.M. White, J. Phys. Chem. 95 (1991) 5575.
Chemistry, University Science Books, Mill Valley, [118] F. Zaera, Langmuir 12 (1996) 88.
California, 1987. [119] F. Zaera, J. Phys. Chem. 94 (1990) 5090.
[82] H.R. Rogers, C.L. Hill, Y. Fujiwara, R.J. Rogers, H.L. [120] F. Zaera, J. Catal. 121 (1990) 318.
Mitchell, G.M. Whitesides, J. Am. Chem. Soc. 102 (1980) [121] F. Zaera, Langmuir 7 (1991) 1998.
217. [122] F. Zaera, Catal. Lett. 11 (1991) 95.
[83] S. Tjandra, F. Zaera, J. Vac. Sci. Technol. A10 (1992) 404. [123] F. Zaera, Surf. Sci. 262 (1992) 335.
[84] J.-L. Lin, B.E. Bent, J. Phys. Chem. 96 (1992) 8529. [124] S. Tjandra, F. Zaera, J. Catal. 147 (1994) 598.
[85] F. Zaera, H. Hoffmann, P.R. Griffiths, J. Electron Spectrosc. [125] T.V.W. Janssens, F. Zaera, Surf. Sci. 344 (1995) 77.
Relat. Phenom. 54-55 (1990) 705. [126] S. Tjandra, F. Zaera, Surf. Sci. 322 (1995) 140.
[86] J.-L. Lin, B.E. Bent, Chem. Phys. Lett. 194 (1992) 208. [127] A. Loaiza, M. Xu, F. Zaera, J. Catal. 159 (1996) 127.
[87] P.E.M. Siegbahn, I. Panas, Surf. Sci. 240 (1990) 37. [128] H. Öfner, F. Zaera, J. Phys. Chem. 101 (1997) 396.
[88] H. Yang, J.L. Whitten, Surf. Sci. 255 (1991) 193. [129] T.V.W. Janssens, D. Stone, J.C. Hemminger, F. Zaera, J.
[89] H. Burghgraef, A.P.J. Jansen, R.A. van Santen, Surf. Sci. Catal. 177 (1998) 284.
324 (1995) 345. [130] F. Zaera, D. Chrysostomou, Surf. Sci. 457 (2000) 89.
[90] S. Tjandra, F. Zaera, J. Catal. 144 (1993) 361. [131] T.I. Taylor, in: P.H. Emmett (Ed.), Catalysis, Vol. 5,
[91] F. Solymosi, Catal. Today 28 (1996) 193. Reinhold, New York, 1957, p. 257.
[92] T.V.W. Janssens, F. Zaera, J. Phys. Chem. 100 (1996) 14118. [132] C. Kemball, Adv. Catal. Relat. Subjects 11 (1959) 223.
[93] G. Wu, M. Kaltchev, W.T. Tysoe, Surf. Rev. Lett. 6 (1999) [133] V. Ponec, in: D.A. King, D.P. Woodfuff (Eds.), The Chemical
13. Physics of Solid Surfaces and Heterogeneous Catalysis:
[94] Z.-M. Liu, X.-L. Zhou, D.A. Buchanan, J. Kiss, J.M. White, Fundamental Studies of Heterogeneous Catalysis, Vol. 4,
J. Am. Chem. Soc. 114 (1992) 2031. Elsevier, Amsterdam, 1982, p. 365.
[95] F. Zaera, N. Bernstein, J. Am. Chem. Soc. 116 (1994) 4881. [134] M.L. Burke, R.J. Madix, J. Am. Chem. Soc. 113 (1991)
[96] S. Tjandra, F. Zaera, J. Catal. 164 (1996) 82. 3675.
[97] H. Ihm, J.M. White, Langmuir 14 (1998) 1398. [135] S. Tjandra, F. Zaera, Langmuir 9 (1993) 880.
[98] H. Celio, K.C. Smith, J.M. White, J. Am. Chem. Soc. 121 [136] F. Zaera, Isr. J. Chem. 38 (1998) 293.
(1999) 10422. [137] J.R. Anderson, N.R. Avery, J. Catal. 5 (1966) 446.
[99] D. Chrysostomou, F. Zaera, J. Phys. Chem. B 105 (2001) [138] F. Garin, F.G. Gault, J. Am. Chem. Soc. 97 (1975) 4466.
1003. [139] F. Zaera, T.V.W. Janssens, H. Öfner, Surf. Sci. 368 (1996)
[100] F. Zaera, S. Tjandra, T.V.W. Janssens, Langmuir 14 (1998) 371.
1320. [140] F. Zaera, S. Tjandra, J. Am. Chem. Soc. 50 (1996) 12738.
[101] T.B. Scoggins, J.M. White, J. Phys. Chem. B 103 (1999) [141] T.V.W. Janssens, G. Jin, F. Zaera, J. Am. Chem. Soc. 119
9663. (1997) 1169.
[102] S. Tjandra, F. Zaera, J. Phys. Chem. B 101 (1997) 1006. [142] D. Godbey, F. Zaera, R. Yates, G.A. Somorjai, Surf. Sci.
[103] S. Tjandra, F. Zaera, J. Phys. Chem. A 103 (1999) 2312. 167 (1986) 150.
[104] D. Chrysostomou, A. Chou, F. Zaera, J. Phys. Chem. B 105 [143] M. Salmerón, G.A. Somorjai, J. Phys. Chem. 86 (1982) 341.
(2001) 5968. [144] F. Zaera, S. Tjandra, J. Am. Chem. Soc. 115 (1993) 5851.
[105] G.S. Jones, M. Mavrikakis, M.A. Barteau, J.M. Vohs, J. [145] T.V.W. Janssens, F. Zaera, Surf. Sci., in press.
Am. Chem. Soc. 120 (1998) 3196. [146] C. Kemball, J.C. Kempling, Proc. R. Soc. London A 329
[106] G. Wu, D. Stacchiola, M. Kaltchev, W.T. Tysoe, Surf. Sci. (1972) 391.
463 (2000) 81. [147] S.M. Davis, W.D. Gillespie, G.A. Somorjai, J. Catal. 83
[107] L. Hanley, X. Guo, J.T. Yates Jr., J. Phys. Chem. 93 (1989) (1983) 131.
6754. [148] R. Brown, C. Kemball, J.A. Oliver, I.H. Sadler, J. Chem.
[108] N.R. Gleason, C.J. Jenks, C.R. French, B.E. Bent, F. Zaera, Res. (M) (1985) 3201.
Surf. Sci. 405 (1998) 238. [149] X.-L. Zhou, F. Solymosi, P.M. Blass, K.C. Cannon, J.M.
[109] S. Tjandra, F. Zaera, Langmuir 10 (1994) 2640. White, Surf. Sci. 219 (1989) 294.
[110] F. Zaera, J. Am. Chem. Soc. 111 (1989) 8744. [150] A. Berko, F. Solymosi, J. Phys. Chem. 93 (1989) 12.
[111] J.G. Forbes, A.J. Gellman, J. Am. Chem. Soc. 115 (1993) [151] A.M. Paul, B.E. Bent, J. Catal. 147 (1994) 264.
6277. [152] J.-L. Lin, B.E. Bent, J. Am. Chem. Soc. 115 (1993) 6943.
[112] J.-L. Lin, A.V. Teplyakov, B.E. Bent, J. Phys. Chem. 100 [153] D.H. Fairbrother, X.D. Peng, R. Viswanathan, P.C. Stair, M.
(1996) 10721. Trenary, J. Fan, Surf. Sci. 285 (1993) L455.
F. Zaera / Applied Catalysis A: General 229 (2002) 75–91 91

[154] Y. Zhou, M.A. Henderson, W.M. Feng, J.M. White, Surf. [169] P. Cremer, C. Stanners, J.W. Niemantsverdriet, Y.R. Shen,
Sci. 224 (1989) 386. G. Somorjai, Surf. Sci. 328 (1995) 111.
[155] B.E. Bent, R.G. Nuzzo, B.R. Zegarski, L.H. Dubois, J. Am. [170] F. Zaera, C.R. French, J. Am. Chem. Soc. 121 (1999) 2236.
Chem. Soc. 113 (1991) 1143. [171] F. Zaera, G.A. Somorjai, J. Am. Chem. Soc. 106 (1984)
[156] C.H. Patterson, J.M. Mundenar, P.Y. Timbrell, A.J. Gellman, 2288.
R.M. Lambert, Surf. Sci. 208 (1989) 93. [172] T.P. Beebe Jr., J.T. Yates Jr., J. Am. Chem. Soc. 108 (1986)
[157] W.T. Tysoe, Isr. J. Chem. 38 (1998) 313. 663.
[158] R.M. Ormerod, R.M. Lambert, H. Hoffmann, F. Zaera, J.M. [173] P.S. Cremer, X. Su, Y.R. Shen, G.A. Somorjai, J. Phys.
Yao, D.K. Saldin, L.P. Wang, D.W. Bennett, W.T. Tysoe, Chem. 100 (1996) 16302.
Surf. Sci. 295 (1993) 277. [174] A. Wieckowski, S.D. Rosasco, G.N. Salaita, A. Hubbard,
[159] J.C. Bertolini, J. Massardier, G. Dalmai-Imelik, J. Chem. B.E. Bent, F. Zaera, D. Godbey, G.A. Somorjai, J. Am.
Soc. Faraday I 74 (1978) 1720. Chem. Soc. 107 (1985) 5910.
[160] N.R. Avery, J. Am. Chem. Soc. 107 (1985) 6711. [175] S.M. Davis, F. Zaera, B. Gordon, G.A. Somorjai, J. Catal.
[161] C. Xu, J.W. Peck, B.E. Koel, J. Am. Chem. Soc. 115 (1993) 92 (1985) 240.
751. [176] F. Zaera, G.A. Somorjai, J. Phys. Chem. 89 (1985) 3211.
[162] M. Xi, B.E. Bent, J. Am. Chem. Soc. 115 (1993) 7426. [177] F. Zaera, D. Chrysostomou, Surf. Sci. 457 (2000) 71.
[163] M. Xi, B.E. Bent, Langmuir 10 (1994) 505. [178] J. Kubota, S. Ichihara, J.N. Kondo, K. Domen, C. Hirose,
[164] S.M. Davis, F. Zaera, G.A. Somorjai, J. Catal. 77 (1982) Langmuir 12 (1996) 1926.
439. [179] T. Ohtani, J. Kubota, J.N. Kondo, C. Hirose, K. Domen, J.
[165] J.E. Demuth, Surf. Sci. 93 (1980) L82. Phys. Chem. B 103 (1999) 4562.
[166] L.L. Kesmodel, L.H. Dubois, G.A. Somorjai, Chem. Phys. [180] P.S. Cremer, X. Su, Y.R. Shen, G.A. Somorjai, J. Am. Chem.
Lett. 56 (1978) 267. Soc. 118 (1996) 2942.
[167] A.M. Baró, H. Ibach, J. Chem. Phys. 74 (1981) 4194. [181] P.S. Cremer, X. Su, G.A. Somorjai, Y.R. Shen, J. Mol. Catal.
[168] P. Skinner, M.W. Howard, I.A. Oxton, S.F.A. Kettle, D.B. A 131 (1998) 225.
Powell, N. Sheppard, J. Chem. Soc. Faraday Trans. 277 [182] S.M. Davis, G.A. Somorjai, J. Catal. 65 (1980) 78.
(1981) 1203. [183] F. Zaera, S. Tjandra, J. Phys. Chem. 98 (1994) 3044.

You might also like