Catalyst
Catalyst
73. N.L. Abbott, C.B. Gorman, and G.M. Whitesides. Langmuir 11:16 (1995).
74. S. Abbott et al. Langmuir 15:8923 (1999).
75. B.S. Gallardo et al. Science 283:57 (1999).
76. K. Ichimura, S.K. Oh, and M. Nakagawa. Science 288:1624 (2000).
77. J. Lahann et al. Science 299:371 (2003).
CATALYSIS BY SURFACES
9.1 Introduction
9.1.1 A Brief History of Surface Catalysisc
9.2 Catalytic Action
9.2.1 Kinetic Expressions
9.2.2 Selective Catalysis
9.2.3 Tabulated Kinetic Parameters for Catalytic Reactions on Metal
Surfaces
9.3 Catalyst Preparation, Deactivation, and Regeneration
9.3.1 Catalyst Preparation
9.3.1.1 Industrial Catalysts
9.3.1.2 Model Catalysts for Surface-Science Studies
9.3.2 Catalyst Deactivation
9.3.3 Catalyst Regeneration
9.4 Techniques to Characterize Catalyst Surface and Study the Reactivity of
Catalysts
9.4.1 Turnover Rate Measurements
9.4.2 High-Pressure Reactors
9.4.3 In Situ Characterization Techniques: High-Pressure Scanning
Tunneling Microscope, Sum Frequency Generation Spectroscopy, and
Ambient Pressure X-Ray Photoelectron Spectroscopy
9.5 Metal Catalysis
9.5.1 Trends Across the Periodic Table
9.5.2 Some Frequently Used Concepts of Metal Catalysis
9.5.3 Most Frequently Used Catalyst Materials
9.6 Case Histories of Surface Catalysis
9.6.1 Ethylene Hydrogenation Over Platinum Surfaces
9.6.1.1 Surface Species Involved in Ethylene Hydrogenation
9.6.1.2 The Role of Reaction Intermediates in Ethylene Hydrogenation
9.6.1.3 Structure Insensitivity of Ethylene Hydrogenation
9.6.1.4 Catalysis in the Presence of a Strongly Adsorbed Overlayer
9.6.1.5 Summary
9.6.2 Ammonia Syntheses
9.6.2.1 Thermodynamics and Kinetics
9.6.2.2 Catalyst Preparation
9.6.2.3 Activity for Ammonia Synthesis Using Transition Metals Across the
Periodic Table
9.6.2.4 Surface Science of Ammonia Synthesis
9.6.2.5 Mechanism and Kinetics of Ammonia Synthesis
9.6.2.6 Summary
9.6.3 Oxidation of Carbon Monoxide on Transition Metal Catalysts
9.6.3.1 Carbon Monoxide Oxidation Under UHV Conditions
9.6.3.2 Carbon Monoxide Oxidation Under High-Pressure Conditions
9.6.3.3 Carbon Monoxide Oxidation Under Oxygen-Rich Conditions
9.6.3.4 Carbon Monoxide Oxidation Over Nanoparticles
9.6.3.5 Carbon Monoxide Oxidation at High Temperatures
9.6.3.6 Summary
9.7 Selectivity in Multipath Heterogeneous Catalytic Reactions
9.7.1 Energetic View of a Heterogeneous Catalytic Reaction with Multiple
Products
9.7.2 Surface Structure and Selectivity
9.7.3 Alloy Catalysts and Selectivity
9.7.4 Adsorbate-Induced Surface Restructuring
9.7.5 Strong Metal Support Interaction
9.7.6 Oxidation States of Metal Catalyst and Selectivity
9.7.7 Reaction Intermediates
9.7.8 Reaction Conditions and Selectivity
9.7.9 Other Important Research Directions to Catalytic Reaction Selectivity
9.7.10 Summary
9.8 Summary and Concepts
9.9 Problems
9.10 References
List of Tables
Table 9.1 Ethane Hydrogenolysis
Table 9.2 Propane Hydrogenolysis
Table 9.3 Cyclopropane Ring Opening
Table 9.4 Cyclopropane Hydrogenolysis
Table 9.5 n-Butane Hydrogenolysis
Table 9.6 Isobutane Hydrogenolysis
Table 9.7 Methylcyclopropane Ring Opening
Table 9.8 n-Pentane Hydrogenolysis
Table 9.9 Isopentane Hydrogenolysis
Table 9.10 Neopentane Hydrogenolysis
Table 9.11 Cyclopentane Ring Opening and Hydrogenolysis
Table 9.12 n-Hexane Hydrogenolysis
Table 9.13 2-Methylpentane Hydrogenolysis
Table 9.14 3-Methylpentane Hydrogenolysis
Table 9.15 Cyclohexane Hydrogenolysis
Table 9.16 Methylcyclopentane Ring Opening
Table 9.17 Methylcyclopentane Hydrogenolysis
Table 9.18 Benzene Hydrogenolysis
Table 9.19 n-Heptane Hydrogenolysis
Table 9.20 Toluene Hydrodealkylation and Hydrogenolysis
Table 9.21 Other Hydrogenolysis Reactions
Table 9.22 Cracking Reactions Over Nickel Powder
Table 9.23 Ethylene Hydrogenation
Table 9.24 Hydrogenation Reactions of Terminal Olefins
Table 9.25 Benzene Hydrogenation
Table 9.26 Other Hydrogenation Reactions
Table 9.27 Cyclohexane Dehydrogenation to Benzene
Table 9.28 Other Dehydrogenation (D) Reactions
Table 9.29 n-Butane Isomerization (I)
Table 9.30 Isobutane Isomerization (I)
Table 9.31 n-Pentane Isomerization (I)
Table 9.32 Neopentane Isomerization (I)
Table 9.33 n-Hexane Isomerization (I)
Table 9.34 2-MethyIpentane Isomerization (I)
Table 9.35 3-Methylpentane Isomerization (I)
Table 9.36 n-Heptane Isomerization (I)
Table 9.37 Other Isomerization (I) Reactions
Table 9.38 n-Pentane Dehydrocyclization
Table 9.39 n-Hexane Dehydrocyclization
Table 9.40 Other Dehydrocyclization Reactions
Table 9.41 Hydro- and Dehydroisomerization (DI) Reactions
Table 9.42 n-Hexane Conversion
Table 9.43 Cyclopropane Hydrogenation
Table 9.44 Propene Hydrogenation
Table 9.45 Neohexane Hydrogenolysis
Table 9.46 Neopentane Conversion
Table 9.47 Toolbox for Studying 2D or 3D Nanoparticles
Table 9.48 Several Structure-Sensitive and Structure-Insensitive Catalytic
Reactions
Table 9.49 Chemical Processes that are the Largest Users of Heterogeneous
Catalysts at Present and the Catalysts that are Utilized Most Frequently
Table 9.50 Chemical Processes that are the Largest Users of Homogeneous
Catalysts at Present and the Catalysts that are Utilized Most Frequently
Table 9.51 Kinetic Data for C2H4 Hydrogenation Over Various Pt Surfaces
Table 9.52 Bandgaps of Several Oxides
9.1 INTRODUCTION
In a surface catalytic process, the catalytic reaction occurs repeatedly by a
sequence of elementary steps, which include adsorption, surface diffusion,
chemical rearrangements (bond breaking, bond-forming, molecular
rearrangement) of the adsorbed reaction intermediates, and finally desorption of
the products.
Catalytic reactions play all important roles in our life. Most biological
reactions that build the human body, as well as the reactions that control the
functioning of the brain and other vital organs, are catalytic. Photosynthesis and
the majority of chemical processes that are utilized in chemical technology are
also catalytic reactions. These range from oil refining and the production of
chemicals by hydrogenation, dehydrogenation, partial oxidation, and organic
molecular rearrangements (isomerization, cyclization), to ammonia (NH3)
synthesis and fermentation.
deactivation.
The energy crisis in the early 1970s renewed interest in chemicals and fuels,
producing technologies using feedstocks other than crude oil. Intensive research
was carried out utilizing coal, shale, and natural gas to develop new technologies
and to improve on the activity and selectivity of older catalyst-based processes.
Increasing concern about environmental quality led to the development of the
catalytic converter for automobiles and to other, NO-reducing catalysts.
Modern surface science developed during the same period and has been
applied intensively to explore the working of catalysts on the molecular level, to
characterize the active surface, and to aid the development of new catalysts for
new chemical reactions. Indeed, surface science provides the means to explore
the molecular structure and mechanisms of elementary reaction steps and to
provide for rational design for modification of catalyst activity and selectivity.
This is carried out usually by altering the structure of the surface and by using
coadsorbed additives as bonding modifiers for reaction intermediates on the
surface.
This chapter describes the important macroscopic and molecular concepts of
surface catalysis that emerged from studies of recent decades. We review what is
known about a few important catalytic reactions that provide case histories of the
state of modern surface science of catalysis and of catalytic science.
During the later part of the 20th century, surface-science research of metal
catalysis were mainly focused on increasing catalyst activity for one-product
reactions (e.g., NH3 synthesis, CO oxidation, and C2H4 hydrogenation). With
increasing concern for the need of clean air and water, minimizing chemical
waste by reaction selectivity becomes the frontier of surface-science research in
the 21st century. To form only one desired product out of several
thermodynamically possible waste byproducts is the motivation of “green
chemistry”, which attempts to achieve 100% molecular selectivity as fuels and
chemicals are produced from methane and carbon solids. The challenge of
catalysis science is how to build the catalyst architecture to steer the reaction
intermediates along the desired reaction path while inhibiting the formation of
other waste products.
The need to produce fuels and chemicals from renewable energy sources (e.g.,
biomass) and to harness sunlight to dissociate H2O and CO2 provides an
unprecedented challenge and opportunity for catalysis science. Exploring how
fuels and chemicals are produced in this way promises to lead to new science
and technologies through the intellectual ferment that always accompanies the
use of new feedstocks, such as when coal was introduced in the 19th century and
oil in the 20th century.
9.2 CATALYTIC ACTION
One of the major functions of a catalyst is to aid in rapidly achieving chemical
equilibrium for certain chemical reactions. Two of the simpler, although
important, reactions that demonstrate this type of catalytic action are the
formation of water from oxygen and hydrogen ( ) and the
formation of ammonia from hydrogen and nitrogen (N2 + 3H2 → 2NH3). Water
has a standard free energy of formation ΔG0 = –58kcal mol−1 Yet O2 and H2 gas
mixtures may be stored indefinitely in a glass bulb without showing signs of any
chemical reaction. Just by dropping a high-surface-area Pt gauze into the
mixture, the reaction occurs instantaneously and explosively—as demonstrated
to the delight of freshman chemistry students in introductory chemistry courses.
The reason for this striking effect can be explained as follows: To form H2O,
one of the diatomic molecules, H2 or O2 must be dissociated first. In the gas
phase, the dissociation energies are 103 kcal mol−1 for H2 and 117 kcal mol−1 for
O2 [3], and are much larger than the thermal energy, RT ( ~0.6 kcal mol−1 at 300
K). Even after the dissociation of one of the diatomic molecules, the subsequent
atom–molecule reactions (H + O2 or H2 + O) still require an activation energy of
~ 10 kcal mol−1 [4]. Thus the gas-phase reaction is very improbable under any
circumstances. On a properly structured Pt surface, however, both molecules
dissociate to atoms with near-zero activation energies (H2 + 2Pt → 2H—Pt and
O2 + 2Pt → 2O—Pt), as shown by low-pressure surface studies [5, 6]. In
addition, the atom-atom or atom-molecule reactions that subsequently take place
on the surface have very low or no activation energies in contrast to the gas
phase [5]. Thus the surface catalytic action involves the ability to atomize the
diatomic molecules with large bonding energies by forming chemisorbed atomic
intermediates on the surface and to lower the activation energy for the
subsequent reactions on the surface.
Similarly, the synthesis of ammonia from dinitrogen and hydrogen (N2 + 3H2
→ 2NH3) requires the “activation” of the N≡N triple bond to dissociate the
molecule. The N atoms then must react with H atoms or molecules to produce
NH3. The very large dissociation energy of N2 (280 kcal mol−1) makes it
virtually impossible for this reaction to occur in the gas phase. On an Fe surface,
however, N2 dissociates on a properly structured surface [e.g., the (111) crystal
face] with a small activation energy (3 kcalmol−1), which is the key initiation
step for the catalytic reaction. Iron also readily atomizes the H molecules. The
chemisorbed N atoms then react with H atoms on the surface to produce NH,
NH2, and finally NH3 molecules that desorb into the gas phase.
(9.1)
Reaction probability can be readily obtained by dividing by the rate of molecular
incidence F, which is obtained from the kinetic theory expression, F =
P/(2πMRT)1/2 with P is the gas pressure of the reactant and M is the mass of the
reactant molecule.
The specific catalytic reaction rate R can often be expressed as the product of
the rate constant k and a reactant pressure (or concentration)-dependent term
(9.2)
where Pi is the partial pressure of the reactants. The rate constant for the overall
catalytic reaction may be a function of the rate constants of many of the
elementary reaction steps that precede the rate-determining step. Because the
slowest rate-reaction step may change as the reaction conditions vary
(temperature, pressure, relative surface concentrations of reactants, catalyst
structure), k also may change to reflect the changing reaction mechanism.
Nevertheless, an effective reaction rate k can be described by the Arrhenius
expression
(9.3)
where A is the temperature-independent pre-exponential factor and ΔE* is the
apparent activation energy measured under catalytic reaction conditions.
Ranges of turnover rates for hydrocarbon reactions are shown in Figure 9.1.
Turnover rates between 10−4 and 100 are used in the various technologies. Thus
the temperature employed is adjusted to obtain the desired rates. The more
complex isomerization, cyclization, dehydrocyclization, and hydrogenolysis
reactions have activation energies ΔE* in the range of 35–45 kcal mol−1; and
thus according to the Arrhenius expression, Eq. 9.3, high temperatures are
required to carry them out at the desired rates. Hydrogenation reactions have
activation energies of 6–12 kcal mol−1 and therefore may be performed at high
rates of 300 K or below. Thus, there are at least two classes of reactions
distinguishable by their very different activation energies that may be carried out
at high and at low temperature, respectively.
Figure 9.1. Block diagram of hydrocarbon conversion over a Pt catalysts
showing the approximate range of reaction rates and temperature ranges that are
most commonly studied.
Combustion reactions are highly exothermic, and can be carried out at a higher
turnover rate (102–103) so that the mass transport of the reactants becomes the
rate-limiting step. Basically, the diffusion of the reactants in the gas phase to the
surface is much slower than the surface reaction process. Under these high-
temperature reaction conditions, the reactions between the gas-phase free
radicals may compete with the catalytic combustion on the surface. In order to
study the intrinsic properties of surface reactions, high turnover rates should be
avoided by controlling the catalytic reaction conditions, so that the effects of
mass transport in the gas phase can be neglected.
Figure 9.1 shows that the reaction probabilities can be very low for many
hydrocarbon conversion processes. The main reason for these low reaction
probabilities is the formation of various hydrocarbon intermediates on the
surface under high-pressure reaction conditions. For example, the dissociation of
CH4 on the Pt(1 11) surface may produce surface intermediates (e.g., —CH, —
CH3, —CCH, and —CCH3) at various temperatures (300–500 K) and high
pressures (1–10 Torr) as observed by sum frequency generation spectroscopy
(SFG) vibrational spectroscopy. These surface intermediates may block the
active sites for the dissociation of CH4 and cause the torturously low probability
(~10−8) of the dissociation process [9].
The rates of surface-catalyzed reactions are usually measured by monitoring
the concentrations of reactants and products as a function of time under steady-
state conditions. Such studies tell us relatively little about the elementary surface
reaction steps. Dynamic methods that alter the flow of reactants or introduce
pulses of isotopically labeled reacting species have been useful to distinguish
between reacting intermediates and adsorbed spectator species on surfaces.
These investigations are carried out by following changes of the concentrations
of adsorbates beginning when changes in flow rate commence, as a function of
time, and by monitoring the time-dependent changes in the concentrations of
isotopically labeled product molecules. Spectroscopic techniques that monitor
reaction intermediates on surfaces can further provide molecular details of
reactive species during catalytic reactions under operating conditions.
(9.4)
where Ri (i = 1, 2) are the turnover rates of two reaction steps, respectively. A
good example of this type of reaction is the stepwise dehydrogenation of
cyclohexane to cyclohexene and then to benzene.
Figure 9.2. Various organic molecules that can all be produced by the catalyzed
reactions of n-hexane [10].
When two or more parallel reaction paths are operative, as is the case during n-
hexane conversion, the reaction scheme is
(9.5)
For a reaction with n pathways, we can define the fractional catalytic selectivity,
Si, as the fraction of reacting molecules that are converted along the ith pathway
(9.6)
An additional possibility is provided by competitive parallel reactions
(2.7)
Here, the ratio of rates, R1/R2, defines the kinetic selectivity. The activity (rate)
and the selectivity are the key parameters of any catalytic reaction.
Figure 9.3. The TEM image (a) and typical particle size distribution (b) for a 3
wt% Pt/ Al2O3 catalyst prepared by incipient wetness from tetrammine Pt(II)–
nitrate (calcination temperature 260°C).
Many of the catalyst poisons act by blocking active surface sites. In addition,
poisons may change the atomic surface structure in a way that reduces the
catalytic activity. Sulfur, for example, is known to change the surface structure
of Ni [34]. By forming chemical bonds of different strengths on the different
crystal planes, it provides a thermodynamic driving force for the restructuring of
the metal particles. Sometimes the rate of deactivation of metal catalysts from
small concentrations of S can be very dramatic. The automobile catalytic
converter necessitated the removal of tetraethyllead from gasoline, one of the
best anti-knocking agents, because it readily poisoned the Pt–Pd catalyst by
depositing lead sulfate on the noble metal surfaces. One of the major causes of
deactivation in crude oil cracking catalysts is the deposition on the catalyst
surface of metallic impurities that are present as compounds in the reactant
mixture. Vanadium- and titanium-containing organometallic compounds
decompose and not only deactivate the catalyst surface, but often plug the pores
of the high-surface-area supports, thereby impeding the reactant–catalyst contact
during petroleum refining.
A freshly prepared catalyst may not exhibit optimum catalytic activity upon its
first introduction into the reactant stream. There may be efficient, but
undesirable, side reactions that need to be eliminated. For this purpose, a small
amount of “poison” is often added to the reaction mixture or introduced in the
form of pretreatment. Thus deactivating impurities may also be used, in small
quantities, to improve the selectivity of the working catalyst.
Figure 9.15. Catalytic activities of transition metals across the periodic table for
the hydrogenolysis of the C—C bond in ethane (C2H6), the C—N bond in
methylamine (CH5N), and the C—Cl bond in methyl chloride (CH3Cl) [10].
The influence of the electronic structure of surface atoms shows up not only in
producing the volcano-shaped trends of transition metal catalytic activity across
the periodic table, but also in producing the structure sensitivity of certain
catalytic reactions on a given transition metal. A catalytic reaction is defined as
structure sensitive if the turnover rate changes markedly as the surface structure
of the catalyst is changed. Reaction studies on single crystals revealed the
importance of steps of atomic height and of kinks in the steps in increasing
reaction rates for H2/D2 exchange, for dehydrogenation and hydrogenolysis.
Theoretical studies indicate large changes in the local density of electronic states
at the surface defect sites that correlate with changes in catalytic activity.
(9.8)
where α is a constant and Θ is called the isokinetic temperature, at which the
rates on all the catalysts are equal. For the methanation reaction [63], α ~ 0 and
Θ = 436 K. Thus ln ~ l.lΔE*kcalmol−1. Figure 9.16 shows the compensation
effect for the methanation reaction for eight different metal catalysts. The ln
versus ΔE* plotsyield a straight-line relationship. Figure 9.17 shows the
compensation effect for the hydrogenolysis reactions whose rates are displayed
in Figure 9.15.
Figure 9.16. Compensation effect for the methanation reaction. The logarithm of
the pre-exponential factor is plotted againt the apparent activation energy, ΔE*,
for this reaction over several transition metal catalysts [63].
Possible explanations for the compensation effect have been explored
extensively and are reviewed by Bond et al. [64]. One of the interpretations of
the compensation effect cited by Bond et al. is a model proposed by Larsson.
This model assumes there is a transfer of energy into the vibrational mode of the
reactant that “most effectively distorts the molecule toward the structure it has in
the “activated complex” of the reaction” [65, 66]. In another interpretation,
Bligaard et al. suggest the compensation effect arises from “a switching of
kinetic regimes”, meaning there is a monotonic relationship between “the
activation energy of the rate-limiting step and the stability of the reaction
intermediates on the surface” [67].
Figure 9.17. Compensation effect for the hydrogenolysis reactions. The
logarithm of the pre-exponential factor is plotted against the apparent activation
energy, ΔE*, for this reaction over several transition metal catalysts. The
squares, triangles, and circles represent values for C5H6, CH5N, and CH3Cl
hydrogenolysis, respectively [10].
During most reactions, the surface of the active metal catalyst is covered with
a strongly chemisorbed overlayer that remained tenaciously bound to the surface
for 102–106 turnovers. During hydrocarbon reactions, this is a carbonaceous
overlayer with a composition of (H/C) ~ 1, during NH3 syntheses it is
chemisorbed N, and during hydrodesulfurization it is a mixture of S and C. It is
believed that this overlayer may play a role in restructuring the surface to create
new active sites and in altering the bonding of reactants, intermediates, and
products.
It is observed that, during hydrocarbon conversion reactions over Pt,>80% of
the metal surface is covered with the carbonaceous deposit [68]. Increasing
evidence shows that the uncovered metal sites as well as the carbonaceous layer
are active parts of the working catalyst. Only when this carbonaceous layer is
totally dehydrogenated will it deactivate the catalyst by forming a cross-linked
graphite coating. This notion has been used to explain the structure insensitivity
of some hydrocarbon conversion reactions (e.g., the hydrogenation of C2H4)
[69]. Basically, if the reactions take place on top of the strongly chemisorbed
overlayer, the structure of the underlining metal surface will not affect the
reaction rates directly. The metal surface only participated indirectly by aiding
the dissociation of molecular hydrogen.
Structure and bonding modifiers are often introduced as important additives
when formulating complex catalyst systems. Structural promoters can change the
surface structure that is often the key to catalyst selectivity. Aluminum oxide
facilitates the restructuring of Fe in the presence of nitrogen to produce surfaces
that are most active during NH3 synthesis. Alloy components may not participate
in the reaction chemistry, but modify structure and site distribution on the
catalyst surface. Site blocking can improve selectivity, which has been proven
for many working catalyst systems. Sulfur and silicon or other strongly adsorbed
atoms that seek out certain active sites can block undesirable side reactions.
Bonding modifiers are employed to weaken or strengthen the chemisorption
bonds of reactants and products. Strong electron donors (e.g., K) or electron
acceptors (e.g., Cl) that are coadsorbed on the catalyst surface are often used for
this purpose. Alloying may create new active sites (mixed-metal sites) that can
greatly modify activity and selectivity. New catalytically active sites can also be
created at the interface between the metal and the high-surface-area oxide
support. In this circumstance, the catalyst exhibits the so-called strong metal-
support interaction (SMSI). Titanium oxide frequently shows this effect when
used as a support for catalysis by transition metals. Often the sites created at the
oxide–metal interface are much more active than the sites on the transition
metal.
Figure 9.19. The SFG spectra of saturation coverages of (a) the di-σ-bonded
C2H4 at 202 K on Pt(111) and (b) C2H3 at 300K on Pt(111) [90].
Both experimental and theoretical evidence exist showing that C2H4 is very
mobile on Pt(111) at 300 K. Scanning tunneling microscopy experiments on
ethylidyne-covered Pt(111) at room temperature have shown that C2H3 is not
visible under these conditions [88] most likely due to the mobility of the
ethylidynes at room temperature. In fact, STM images of C2H3/Pt(1 11) at 300 K
look very similar to images of the clean Pt(111) surface. It was only upon
cooling the Pt crystal that the C2H3 became visible. Extended-Huckel
calculations have been performed for an C2H3 moiety on Pt(111) to determine
the activation barrier for surface diffusion. The results show a low-activation
barrier of 0.11 eV, thus explaining the mobility of C2H3 on the surface at 300 K
[91].
Ethyl is the reaction intermediate proposed in the Horiuti–Polanyi mechanism.
The spectrum of C2H5/Pt(111) at 193 K is shown in Figure 9.20. Several CH
stretch features are observable on the surface. The two lower frequency features
(2860 and 2920 cm−1) have been assigned to a Fermi resonance and the CH3
symmetric stretch of the adsorbed C2H5 group.
Figure 9.20. The SFG spectrum of saturation coverage of ethyl (C2H5) groups
on Pt(111) at 193 K under the UHV condition [90].
Figure 9.21. (a) The SFG spectrum of the Pt(111) surface during C2H4
hydrogenation with 100 Torr of H2 and 35 Torr of C2H4 at 295 K. (b) The SFG
spectrum under the same conditions as (a), but on a Pt(111) surface that was
precovered in UHV with 0.25 ML of C2H3 [90].
Ethylene hydrogenation has been performed on the C2H3 precovered Pt(1 11)
under the same partial pressures and temperature to investigate the role of C2H3
in the reaction process. The SFG spectrum in Figure 9.21b shows that the peak
intensity of the di-σ-bonded C2H4 becomes smaller. The coverage of di-s-bonded
C2H4 is 0.02 ML, which is a factor of four drop from the surface coverage
during the reaction on a clean surface. This result can be understood since C2H3
and di-σ-bonded C2H4 are competing for the fcc threefold hollow sites on Pt(1
11). The preadsorbed C2H4 blocks the adsorption sites for the di-σ-bonded C2H4,
so the relative coverage of these two species depends on the prehistory of the
metal surface.
However, the measured reaction turnover rates on the clean and C2H3
precovered Pt(111) are almost the same (11 and 12 ethane molecules per Pt site−1
s−1, respectively), indicating that once the surface is saturated by the two
strongly adsorbed species, their relative coverage does not affect the rate of
C2H6 formation. The inactive role of C2H3 in the process of C2H4 hydrogenation
has been confirmed by several techniques. Under room temperature and 1 atom
of hydrogen, the measured direct hydrogenation rate of chemisorbed C2H3 is
orders of magnitude slower than the rate of C2H4 hydrogenation (Fig. 9.22).
The di-σ-bonded C2H4 is readily hydrogenated under these reaction
conditions. The fact that the coverage of the di-σ-bonded C2H4 is not directly
correlated with the reaction ratesuggests that it is not the reaction intermediate
leading to the activation barrier of the hydro-genation reaction.
Figure 9.22. Turnover rates for C2H4 hydrogenation, the rehydrogenation of
C2H3, and the deutera-tion of the CH3 group of C2H3 on Pt and Rh crystal
surfaces [20]. Note that C2H4 hydrogenation rates are orders of magnitude faster
than the rate of removal of chemisorbed C2H3.
At present, which reaction intermediate is directly responsible for the
formation of ethane is still under debate. The general properties of the
intermediate include (1) it must be a weakly adsorbed species, since, if it were a
strongly absorbed species like C2H3, it should be easily identified in the IR or
SFG spectra; (2) It must be bonded to a single atom on the metal surface
because, if it were bonded to multiple atoms, the intermediate would have to
compete with C2H3 and the di-σ-bonded C2H4 for adsorption sites, meaning that
the coverage of C2H3 and the di-σ-bonded C2H4 should have a huge effect on the
hydrogenation rate.
The two most-likely active intermediates are the p-bonded C2H4 and C2H5.
The π-bonded C2H4 shows up in the SFG spectrum (Fig. 9.21) as a weak and
broad hump ~3000 cm−1. Its surface coverage is only ~4% ML. The peak
intensities of this species are similar during the reaction over the clean and the
C2H3 precovered Pt(111), which agrees with the unchanged reaction rate on
these two surface. Therefore this species is likely to be the key intermediate in
C2H4 hydrogenation. However, recent efforts by surface techniques other than
SFG could not confirm the existence of this species under similar reaction
conditions because of the very small adsorption energy of this species [92].
The C2H5 group shows up in the SFG spectrum in Figure 9.21 as a shoulder
~2850 cm−1, corresponding to the Fermi resonance peak (see Fig. 9.20). The
existence of a C2H5 group is more evident when the reaction is performed under
a higher partial pressure of hydrogen and at room temperature. Under these
reaction conditions, the surface coverages of C2H3 and the di-σ-bonded C2H4 are
very low and the two peaks at 2850 and 2925 cm−1 become evident in the SFG
spectrum (Fig. 9.23). Recently, time-resolved Fourier transform infrared (FTIR)
spectra of transient-state C2H4 hydrogenation over an alumina-supported Pt
catalyst have been recorded at 25-ms time resolution [93, 94]. Based on the
observation that, at 323 K, the decay time of the surface C2H5 concentration
(~122 ms) coincides with the rise time of C2H6 concentration (~144 ms), it has
been suggested that hydrogenation of surface C2H5 is the rate-limiting step in the
C2H4 hydrogenation process.
Although the amount of deposit does not change much with temperature, the
composition does; it becomes much poorer in hydrogen as the reaction
temperature is increased. The adsorption reversibility decreases markedly with
increasing temperature as the carbonaceous deposit becomes more hydrogen
deficient. As long as the composition is ~CnH1.5n and the temperature is above
450 K, the organic deposit can be removed readily with hydrogen. With
increasing reaction temperatures above 450 K, it converts to an irreversible
adsorbed deposit with a composition of C2nHn that can no longer be readily
removed (hydrogenated) in the presence of excess hydrogen [100].
Nevertheless, the catalytic reaction proceeds readily in the presence of this
active carbonaceous deposit [101, 102]. Above 750 K, this active C layer is
converted to a graphitic layer that deactivates the metal surface, and all chemical
activity for any hydrocarbon conversion reaction ceases. Hydrogen-exchange
studies indicate rapid exchange between the hydrogen atoms in the adsorbing
reactant molecules and hydrogen in the active, but irreversibly adsorbed, deposit.
Only the C atoms in this layer do not exchange. Thus, one important property of
the carbonaceous deposit is its ability to store and exchange hydrogen [101–
103].
The structure of the adsorbed hydrocarbon monolayers was submitted to
detailed studies by LEED and HREELS [104]. In the temperature range of 300–
400 K, the adsorbed alkenes form C2H3 molecules (shown in Chapter 6). The C
—C bond closest to the metal is perpendicular to the surface plane, and its 1.5-Å
length corresponds to a single bond. The C atom that bonds the molecule to the
metal is located in a threefold site 2.0 Å equidistant from itsnearest metallic
neighbors [105]. This bond is appreciably shorter than the covalent metal–
carbon bond (2.2 Å) and is indicative of multiple metal-carbon bonds of the
carbene or carbyne type. Although this layer is ordered, on being heated to ~
100°C it disorders and hydrogen evolution is detectable by a mass spectrometer
attached to the system. As the molecules dehydrogenate, the disordered layer is
composed of CH2, C2H, and CH type fragments that can be identified by
HREELS [104]. Only after being heated to ~400°C do the fragments lose all
their hydrogen and a graphite overlayer forms. These sequential bond-breaking
processes, which occur as a function of temperature, are perhaps the most
important and unique characteristics of the surface chemical bond (Section 6.3).
Although the surface remains active in the presence of organic fragments of C2H
stoichiometry, it loses all activity when the graphite monolayer forms.
Figure 9.25. Carbon- 14-labeled C2H4 (or other alkenes) was chemisorbed as a
function of temperature on a flat Pt( 111) crystal face. The H/C ratio of the
adsorbed species was determined from hydrogen thermal desorption. The
amount of preadsorbed alkene that could not be removed by subsequent
treatment in 1 atm of hydrogen represents the irreversibly adsorbed fraction. The
adsorption reversibility decreases markedly with increasing adsorption
temperature as the surface species become more hydrogen deficient. The
irreversibly adsorbed species have long residence times, on the order of days
[99].
(9.9)
Figure 9.27. The free energy of NH3 formation as a function of temperature
[108].
where w = 1.5 and α = 0.75. The parameters k, Ka, and K3 are constants. The rate
of NH3 formation depends in a rather complex manner on the partial pressures of
N2, H2, and NH3, mostly because of the possibility of a back-reaction. When the
partial pressures are far from equilibrium, this may be neglected and under this
circumstance the rate depends only on the N2 pressure. This indicates that the
rate-limiting step is the dissociative adsorption of N2 on the catalyst surface, a
conclusion shared by most of the practitioners.
Other important rate equations that are applicable in a variety of experimental
conditions have been proposed by Ozaki et al. [113]. The net activation energy
for the reaction is 76kJmol−1, which is in excellent agreement with the 81-kJmol
−1 value determined using single-crystal Fe surfaces [108, 114–117].
Figure 9.32. Rates of NH3 synthesis over clean Fe single-crystals and water-
induced restructured AlxOy/Fe surfaces. Restructuring conditions are given in
the figure [120].
The activity of the Fe(110) and Fe(100) surfaces for NH3 synthesis can also be
enhanced to the level of Fe(111) by water-vapor pretreatments in the absence of
aluminum oxide, but under these circumstances the enhancement in activity is
only transient. Figure 9.33 shows the rate of NH3 synthesis as a function of
reaction time for restructured Fe(110) and AlxOy/Fe(110) surfaces. Both surfaces
have an initial activity similar to that of the clean Fe(111) surface. The
restructured AlxOy/Fe(110) surface maintains this activity for >4h while the
restructured Fe(110) surface loses its activity for NH3 synthesis within 1 h of
reaction.
Figure 9.40. The activation energy for NH3 synthesis on Fe(111) and
K/Fe(111).Within experimental error there is no change, suggesting that K does
not change the reaction mechanism of NH3 synthesis [141].
These results are consistent with the earlier literature in which the effects of K
on doubly promoted (AlxOy and K) catalysts were studied [145, 146]. It was
shown that the turnover number for NH3 synthesis is roughly the same over
singly (AlxOy) and double promoted Fe when 1 -atm reactant pressure of N2 and
H2 is used [ 146]. This implies that at low-pressure conditions, the gas-phase
NH3 concentration is not high enough for K to exert a promoter effect. As higher
reactant pressures are achieved (95–200 atm), the promoter effect of K becomes
significant. It was found that doubly promoted catalysts became increasingly
more active than catalysts without K when the concentration of NH3 increased in
the gas phase [ 145]. This implies that K makes the apparent reaction-order
dependence in NH3 partial pressure less negative over commercial catalysts, in
agreement with the single-crystal work.
(9.10)
where * schematically denotes a vacant site on the catalyst surface.
Figure 9.41. Schematic energy profile of the progress of NH3 synthesis on Fe (in
KJmol−1) [108].
The progress of the reaction may be rationalized in terms of its energy profile
as reproduced in Figure 9.41.
Attempts at theoretical modeling of the kinetics along these lines were recently
performed independently by two groups: Bowker et al. [148, 149] and Stolze and
Norskov [150–155]. The latter group starts with the experimentally well-
established fact that dissociation of adsorbed nitrogen is rate-limiting. The
overall rate can then be calculated from the rate of this step and the equilibrium
constants of all the other steps. This reduces the number of input parameters
significantly. The adsorption-desorption equilibria are treated with the
approximation of competitive Langmuir-type adsorption and by evaluation of the
partition functions for the gaseous and adsorbed species. The data for the
potassium-promoted Fe(111) surface are used for the rate of dissociative
nitrogen adsorption and are also representative of the other crystal planes of the
promoted catalyst, as outlined above. The active area of the commercial catalyst
is assumed to equal that derived from selective carbon monoxide chemisorption
as a well-established standard procedure. A particular strength of this model is
the fact that experimental data from single-crystal studies (such as TPD traces)
are reproduced well with the same set of parameters and the same model as used
for the determination of the rate under “real” conditions. Comparison of the
resulting yields against those determined experimentally with a commercial
catalyst yielded general agreement to within a factor better than 2. In Figure
9.42, a compilation of data over a wide range of conditions is presented that
demonstrate this excellent agreement.
A general conclusion from these models based on single-crystal data is that the
most abundant surface species under practical synthesis conditions will be
adsorbed atomic nitrogen (>90%), despite the fact that its formation is the rate-
limiting step of the overall reaction.
Figure 9.42. Comparison of calculated and measured NH3 production over
commercial iron-based catalysts for a broad range of temperatures, pressures,
N/H ratios, and gas flows [155].
9.6.2.6 Summary.
Surface science research has provided molecular-level detail of ammonia
synthesis which has lead to the improvement of industrial catalysts for this
reaction. The rate-limiting step of this reaction is the dissociative adsorption of
N2. Because N2 is an inert reactant, ammonia synthesis proceeds at higher
temperatures and pressures over metal surfaces capable of forming stronger
chemisorption bonds, as compared to reactions with weakly adsorbed reactants,
like ethylene hydrogenation. In order to assist the dissociative adsorption of N2,
highly active sites with specific atomic arrangements, such as C7 sites on
Fe(111), are required. As a result of this specificity, this reaction is very sensitive
to the catalyst surface structure. For Fe-based catalysts in harsh reaction
conditions, the active sites can be stabilized by the presence of aluminum oxide.
Adding K to Fe catalysts further optimizes the ammonia synthesis process by
both enhancing N2 adsorption and assisting NH3 desorption so as to prevent
product poisoning.
(9.11)
where k1 = γpCOSCO is the adsorption rate constant that is proportional to the CO
partial pressure pCO and the sticking probability SCO;γ is a constant relating the
CO pressure to the impingement rate; k2 = v2exp(–E2/RT) is the desorption rate
with v2 the pre-exponential constant; and E2 the desorption activation energy that
is equal to the adsorption energy. The sticking probability is coverage dependent.
Due to the repulsive interaction between neighboring CO molecules, the sticking
probability decreases as the coverage increases, and becomes zero when the
saturation coverage is reached.
Dioxygen adsorbs dissociatively on the late transition metal surfaces [161]. On
the Pt, Pd, and Rh close-packed surfaces, the adsorption energies are 58, 56, and
47kcalper mole of oxygen atoms, respectively. These energies are higher than
CO adsorption energies, so O2 desorption takes place at a much higher
temperature, and under most reaction conditions, the desorption of O2 can
usually be neglected. The higher adsorption energies also indicate that the
adsorbed O atoms are less mobile than the CO molecules. The adsorption
process,
(9.12)
takes two vacant sites on the surface. The initial sticking probabilities are ~0.5
on these metal surfaces. Due to the strong repulsive interaction between
negatively charged O atoms, the sticking probability decreases dramatically as
the O surface coverage increases. Compared to the adsorbed CO molecules, the
distance between neighboring O atoms at saturation coverage is much larger than
that between CO molecules [162–167]. For example, the saturation coverage of
CO on the Pd(111) surface is 0.66 and the distance between neighboring CO
molecules is ~3 Å, while the saturation coverage of O atoms is 0.25 and the
nearest-neighbor distance is 5.5 Å.
The coadsorption of CO and O2 on the surface is rather complicated due to
strong interactions between COads and Oads [168–170]. The rate constants k1, k2,
and k3 are functions of the surface coverage θCO and θO. On the CO precovereed
surfaces, the adsorption of O2 strongly decreases with the increase of CO
coverage. On the (111) surfaces of Pt, Pd, and Rh, the O2 adsorption is
completely prohibited when the CO coverage, θCO ≥ 0.33. If θCO , 0.33 initially,
the adsorption of O2 leads to the separate islands of Oads and COads on the
surface. On O precovered surfaces, the adsorption of CO is not prohibited even
if the surface is initially saturated with Oads. On (111) surfaces, a small amount
of CO adsorption forces the O to compress into small islands, and leads to the
separate islands of Oads and COads. At high CO exposures, CO molecules start to
occupy the vacant sites inside the O islands, and to form a mixed phase where
the two species are in intimate contact. The coadsorption behavior depends on
the adsorption sequence, which can be understood as follows. On a surface
precovered with a high coverage of CO, it is hard to find an open space for the
dissociative adsorption of O2, which requires two vacant neighboring sites. On a
surface precovered with high coverage of Oads, CO can still find a vacant site in
between O adatoms because the O—O distance is quite large even at saturation
coverage.
Figure 9.43 shows in situ STM images of an O precovered Pt(1 11) surface
during its reaction with CO at 247 K [ 171 ]. The Pt( 111) surface is initially
covered with a submonolayer of oxygen that consists of small O islands with the
(2 × 2) O structure (Fig. 9.43a). At the initial stage of CO exposure, the O
islands are compressed to form larger islands, and domains of CO emerges with
no resolvable structure (Fig. 9.43b). Upon further exposure to CO, large CO
islands with the c(4 × 2) CO structure form (Fig. 9.43c). The length of the
boundary between the islands of the two species continuously changes during
the reaction.
There are two possible reaction mechanisms for CO oxidation. First is the
Langmuir–Hinshelwood process, where the two adsorbed species react on the
surface to produce CO2,
(9.13)
Figure 9.43. (a) The STM image of aPt(111) surface covered with a
submonolayer of oxygen. (b) The STM image after 140-s exposure to 5 ×
10−8mbar CO at 247 K. (c) The STM image after 600-s exposure. Image sizes
are 180 × 170 Å [171].
(9.14)
Under UHV conditions, modulated molecular beam experiment have been used
to identify the Langmuir–Hinshelwood process as the reaction mechanism [172,
173], a result also confirmed by isotope-labeling experiments [174].
According to the Langmuir–Hinshelwood process, the reaction rate of CO2
production
(9.15)
if the coadsorption phase of CO and O is a random mixture with adsorbed CO
and O in intimate contact. However, under most reaction conditions, the random
mixture phase is hardly achieved. The coadsorption may lead to separate islands
of different species (see Fig. 9.43), and the reaction rate is found to be
proportional to the boundary length of islands on the surface. This result
suggests that the reaction occurs on the boundary of islands, where COads and
Oads are in intimate contact. The traditional way to address this problem is to
assume that the rate constant also depends on the coverages; that is, k = k(θO,
θCO).
A simple reaction rate model can be derived under the assumption that the
dominant surface species is CO [158]. The approximate reaction rate is given by
(9.16)
where A is the temperature independent constant, and ECO,des is the desorption
energy of CO. The rate is first order with respect to the partial pressure of O2,
and negative first order to the partial pressure of CO. The rate equation can be
rationalized as follows. On a CO dominant surface, the rate-limiting step is the
O2 dissociative adsorption on the vacant sites that are left after the desorption of
CO molecules. Thus the rate increases with the O2 partial pressure, and the
apparent activity energy of the reaction is close to the desorption energy of CO.
Furthermore, the increase of CO partial pressure increases the CO coverage, and
decreases the number of vacant sites, causing a decrease of reaction rate.
Figure 9.49. Two possible oxide structures on a Pt(110) surface during the CO
oxidation reaction. The left side is the incommensurate PtO2 surface structure.
The right side is the commensurate surface structure with the O—Pt—CO
complex formed [53].
Currently, it is believed that a certain oxidation state is responsible for the
turnover rate increase. However, earlier kinetic studies over single-crystal
surfaces (e.g., the result shown in Fig. 9.44) showed that the metal surface may
be deactivated under a high O2 partial pressure due to the formation of the stable
metal oxide. Also note that most of the studies over single-crystal surfaces that
show the highly active surface phase are under transient state conditions rather
than the steady state. On single-crystal surfaces, it is still in question whether the
highly active surface phase exists under steady-state conditions [159].
One interesting aspect of the nanoparticle size effect, which needs further
investigation, is the coadsorption phase of COads and Oads on the nanoparticle
surface. On single-crystal surfaces, COads usually forms islands with sizes
typically >5 nm (Fig. 9.43b). The formation of islands implies that there is a
long-range interaction between adsorbed CO molecules that stabilize this
aggregation. There could be a critical size of the islands below which the
formation of islands becomes energetically unfavorable and formation of a CO
—O mixture phase becomes favorable. If the size of a nanoparticle is small
enough, we would expect that the formation of a CO—O mixed phase is more
favorable than the formation of these CO islands due to limited surface area of
the nanoparticle. The lower activation energy for the CO2 formation from this
mixed phase may be responsible for the increased turnover rate. Particle size
effect on the formation of the nitrogen island formation during CO + NO
reaction over Rh(111) has been studied by Zaera and co-workers [181]. The
studies using in situ STM, molecular beam, and field ion microscopy (FIM) on
the nanoparticles are expected to address this problem in the future [38, 182,
183].
Figure 9.50. (a) Turnover frequency relative to Rh foil at 50 Torr O2, 20 Torr
CO at 200°C, and activation energy for CO oxidation. (b) The thickness of the
oxide shell scales with particle size; 2 and 7-nm nanoparticles are illustrated here
with the oxide layers shown to scale, as determined by AP-XPS [29].
which is facile once the Pt carbonyl is formed on the defects on the Pt surface.
The rapid and highly exothermic gasification of surface C,
(9.18)
may provides another reaction channel for the ignition process. Since the
ignition temperature on the initially C covered Pt(557) surface is lower than that
on the clean Pt surface under the same reaction condition provides addition
evidence that C deposition is an important step for the onset of ignition.
9.6.3.6 Summary.
Research on CO oxidation suggests that this reaction proceeds via the
Langmuir–Hinshelwood mechanism for most transition metal surfaces at
temperatures below ignition and with partial pressure ratio, pCO/po2, close to the
stoichiometric ratio of 2. Under these reaction conditions, this reaction is surface
structure insensitive over most transition metal catalysts. Recent studies under
oxygen-rich conditions on Pt(110) by in situ STM, however, have shown that the
formation of a specific surface oxide phase may result in an increase in the CO
oxidation turnover rate. In addition, the turnover rate over a series of size-
controlled Rh nanoparticle catalysts has been found to increase with decreasing
particle size. This trend correlates with the increase of an Rh surface oxide, as
observed with in situ XPS. Under ignition conditions, recent studies by in situ
SFG have shown that the ignition temperature of this reaction is surface structure
sensitive over Pt single crystal surfaces, which correlates with the surface
structure sensitivity of CO dissociation temperatures over Pt(100), Pt(557), and
Pt(111) (500, 548, and 673 K, respectively). These new findings exemplify the
complexity of the reaction mechanism under relevant reaction conditions and
demonstrate the existence of regimes of both surface structure sensitivity and
surface structure insensitivity for this reaction.
(9.19)
is determined by the difference of the two barrier heights, ΔG1 – ΔG2. Here, c =
A2/A1. At 300 K, RT ~0.6 kcal mol−1, so a small variation of the difference of the
two barrier heights may change the selectivity significantly (Fig. 9.55).
The relative activation barrier heights among reaction paths determine the final
product distribution; that is, the selectivity for a reaction over a given catalyst.
These fundamental parameters depend on the electronic structures of adsorbed
reactants on the metal surface, which can be tuned by changing the properties of
the catalysts, such as the chemical composition of the metal surface, the surface
structure of the metal catalysts, and the oxide support of metal catalysts; and the
reaction conditions, such as temperature and the partial pressures of reactants.
We may call these tunable molecular level properties of catalysts and reaction
conditions the controllable factors for catalytic selectivity. The purpose of
selectivity research is to find the optimal combination of these controllable
factors for each reaction of economic interests.
Figure 9.55. Illustration of the key role that relative heights among activation
free energy barriers for different reactions play in determining the selectivity of
catalytic processes. The diagram on the on of a reactant to two possible products.
The rates of those reactions are determined by the absolute activation barriers,
indicated here as ΔG1 and ΔG2, but selectivity between the two is controlled by
the difference between those two values. The calculated selectivity of this two-
reaction system, plotted as a function of ΔG1 – ΔG2 on the left side of the figure,
shows how a variation of the barrier height difference by 2 kcal mol−1 leads to a
switch in selectivity from the exclusive formation of one product to the other
[181].
It is fair to say that most current research in this field is focused on
establishing the correlation between these controllable factors and reaction
selectivity, and that detailed understanding of how these factors affect the
electronic structures of the adsorbed reactant and further affect the conversion of
reaction intermediates to produce different products, are still limited. Here, we
first discuss some important correlations between controllable factors and
reaction selectivity established on the model catalytic systems. Current advances
in the nanoparticle synthesis that allows us to study the selectivity of size- and
shape-controlled catalysts will be emphasized.
For some multipath reactions, the active sites on the oxide-metal interface area
may produce a reaction product different from the product produced on the pure
metal sites. The atomic fraction of atoms in the oxide–metal interface area
increases as the size of the metal catalyst decreases. Thus, it is expected that the
selectivity of these multipath reactions can be tuned by changing the size of the
metal catalysts [227]. For example, the selectivity of ring opening of
methylcyclopentane over Pt/MgO catalysts shows a very strong dependence on
the size of Pt particles, which is not observed over Pt on the inert SiO2 support
(Fig. 9.64). The increase of n-hexane as the particle size decreases can be
attributed to the preferential formation of n-hexane at the oxide–metal interface,
which has been confirmed by a series of experiments in which the interfacial
area was deliberately poisoned by NH3, CO, and so on [228, 229].
In the oxide–metal interface area, the metal may easily grab O atoms from the
reducible oxide to form certain partially oxidized states, and to leave oxygen
vacancies on the oxide surface. The electronically modified metal atoms in the
interfacial area or the oxygen vacancies on the oxide surface can serve as active
sites for the formation of specific products [230, 231]. For some reactions, it is
believed that the presence of both of them is necessary [232].
Compared with the pure metal sites, the adsorption energies of reactants are
usually changed significantly on these interfacial sites due to their unique
electronic structures. It was proposed by Schwab and Koller that the SMSI effect
can be traced back to an electron exchange between the support and the catalyst
[233]. This idea was developed further by Akubuiro and Verykios, who
suspected that the Schottky-like barrier at the metal–oxide interface may be
responsible for the formation of the unique electronic structure at the interface
[225]. Recently, the successful fabrication of a catalytic nanodiode, provided for
the first time a way to measure the electron transport through the metal–oxide
interface during exothermic catalytic reactions [234–236]. By establishing the
correlation between the electron transport and the reaction activity and
selectivity for different reactions, the studies using catalytic nanodiodes are
expected to provide a more mechanistic understanding of the SMSI effect [237,
238].
Figure 9.64. Ring opening of methylcyclopentane in excess hydrogen at 513 K:
turnover ratio of branched hexanes [2-methylpentane (2-MP) and 3-
methylpentane (3-MP)] to n-hexane versus mean-particle size of Pt deposited on
silica, alumina, magnesia, and acidic alumina supports. The strong size
dependence of the selectivity was observed on the catalysts with strong metal
support interactions [225, 226].
9.7.6 Oxidation States of Metal Catalyst and
Selectivity
A partially oxidized metal surface may serve as an active phase for catalytic
reactions. One example is the highly active surface oxide phase during a CO
oxidation reaction that we discussed earlier in Section 9.6.3.3. From the
thermodynamics point of view, metal nanoparticles are more easily oxidized as
their size becomes smaller. Figure 9.65 shows XPS spectra of Pt nanoparticles
with a size of 1.0 and 1.5 nm. In the smaller particles, ~60% of the Pt atoms are
oxidized, while only ~10% of Pt atoms are oxidized in the bigger nanoparticles.
The oxidation state of these Pt nanoparticles is believed to play a key role in the
size dependence of the selectivity in pyrrole hydrogenation (Fig. 9.66) [239].
The pyrrole hydro-genation to pyrrolidine is followed by ring opening to n-
butylamine, and the scission of the N—C bond to form butane and ammonia. On
single crystal Pt surfaces, this reaction is known to be poisoned by strongly
adsorbed n-butylamine. The N of n-butylamine is more electron-rich than other
intermediates such as pyrrolidine or pyrroline and thus can form stronger bonds
with the surface and consequently inhibit turnover. Figure 9.66 shows the n-
butylamine formation is more facile over larger Pt nanoparticles. The pyrrolidine
formation occurs more easily on the smaller Pt nanoparticles. The change of the
selectivity becomes significant as the size of nanoparticles is reduced below
about 2 nm, which correlates with the significant increase in the oxidation state
of these smaller nanoparticles. It has been suggested that the higher oxidation
state on the small nanoparticles reduces the adsorption energy of n-butylamine
and consequently increases the surface coverages of weakly adsorbed
intermediates for pyrrolidine formation.
Figure 9.66. (Top) The reaction scheme for pyrrole hydrogenation. (Bottom)
The nanoparticle size dependence of selectivity to the reaction products of
pyrrole hydrogenation [239]. (See color insert.)
Under the UHV conditions, TPD combined with techniques such as HREELS
and reflection–absorption infrared spectroscopy (RAIRS) can be used to identify
possible reaction intermediates and to correlate the intermediates with the
desorbed product at a different temperature [181]. For multipath reactions, the
resulting vibrational spectra of the surface species are usually very complex,
andinterpretation of the spectra becomes difficult [195]. In this regard, the
vibrational frequencies of surface intermediates predicted by DFT calculations
become more and more useful for spectral interpretation [241,242]. Once the
spectral signatures of different intermediates have been determined, they serve as
references for the reaction intermediates under high-pressure reaction conditions.
Figure 9.67. (a) Potential elementary steps available to alkyl moieties (iso-butyl
in this example) when chemisorbed on metal surfaces. All these reactions, with
the exception of b-methyl elimination, have been seen in surface science studies.
(b) Elementary hydride elimination steps lead to different reforming products
[240].
Studies on the intermediates during hydrocarbon reforming suggested that the
β-hydride elimination typically dominates the chemistry of adsorbed alkyl
groups on most transition metal surfaces. The reaction step leads to the
formation of alkene products (Fig. 9.67b). Compared to other metal catalysts, Pt
based catalysts have the unique ability to carry out α- and γ-hydride elimination,
which leads to hydrogenolysis, cyclization, and isomerization products. This is
the reason Pt is the most versatile catalyst for selective reforming.
In situ monitoring of the intermediates during catalytic reactions under high-
pressure conditions can be performed by SFG and polarization-modulated
RAIRS (PM–RAIRS) [243–245]. Figure 9.68a shows the SFG spectra of
cyclohexene hydrogenation and dehydro-genation over Pt(111) under 10Torr of
hydrogen and 1.5Torr of cyclohexene [107]. There are three intermediates: 1,4-
cyclohexadiene, π-allyl c-C6H9, and 1,3-cyclohexadiene (C6H8), which are
detectable at different temperature ranges. At 303 K, 1,4-cyclohexadiene is the
major surface species and small features from 1,3-cyclohexadiene are also
visible. In the temperature range from 303 to 400 K, the turnover rates in Figure
9.68b show cyclo-hexane as the major product.
The corresponding major surface intermediate is π-allyl c-C6H9. At higher
temperatures, the onset of 1,3-cyclohexadiene formation on the surface is
correlated with the production of benzene in the gas phase, which suggests that
1,3-cyclohexadiene is the reactive intermediate.
The selective hydrogenation of the C=O carbonyl group in the presence of a
C=C bond can produce the unsaturated alcohol, an important intermediate in the
pharmaceutical and fragrance industries. Figure 9.69 compares the selectivity of
the hydrogenation of cro-tonaldehyde and prenal over Pt(111) [259]. It is shown
that the formation of the unsaturated alcohol products are more favored in the
hydrogenation of prenal than of crotonaldehyde. The bulky heads of two CH3
groups in prenal may weaken the di-σ(CC) surface bonds of intermediates and
cause the difference in selectivity. Indeed, the SFG spectra of the Pt(111)
surfaces during the two reactions show marked differences (Fig. 9.70). During
the hydrogenation of crotonaldehyde, the major intermediate is the η2-cis-
crontoaldehyde species that is characterized HETEROGENEOUS by peaks at
2860, 2915, 2965, and 2995 cm−1 (Fig. 9.70a). No O–H stretching mode is
observed in this case. During the hydrogenation of prenal (Fig. 9.70b), the O–H
stretch mode is visible as a broad peak centered at 3380 cm−1, which indicates
the existence of chemisorbed unsaturated alcohol on the surface. The existence
of the η3-di-σ(CC)-σ(O)-cis species or η4-di-σ(CC)-di-σ(CO)-cis species, that
form three or four bonds with the surface are indicated by the peak at 2945 cm−1
at higher temperatures. The fact that the O atom in the highly coordinated η3 and
η4 species is close to the surface may cause the preferential hydrogenation of the
C=O bond to selectively produce the unsaturated alcohol product in the
hydrogenation of prenal.
Figure 9.68. (a) The reaction intermediates detected by SFG during cyclohexene
hydrogenation and dehydrogenation over Pt(111) at different temperatures. The
reaction conditions are 1.5 Torr for cyclohexene and 15 Torr for hydrogen. (b)
The turnover rates for benzene and cyclohexane at different temperatures.
Correlation between surface intermediates and formed reaction products may
help identifying the reactive intermediates for the products [107]. (See color
insert.)
Figure 9.72. (a) The 108 ×108-Å STM images showing mirror chiral surfaces
created by (R,R)- and (S,S)-tartaric acid adsorbed on Cu(110) [202]. (b) Three-
dimensional ball models of the Ag(643) surface and its mirror image, the
Ag(643) surface [203].
Currently, metal oxide catalysts are being explored for carrying out
hydrocarbon conversion selectively [261]. Metal oxide catalysts are used in
oxidative dehydrogenation of alkanes to alkenes and selective partial oxidation
of hydrocarbons to oxygenated products. The challenge in these applications is
to avoid the generation of CO2 from the complete oxidation of desired products.
The activity and selectivity of metal oxide catalysts can be correlated with
structure, the oxidation states of the metal, and basicity of the oxide support
[262-265.
9.7.10 Summary
Achieving high catalytic selectivity is the goal of catalysis science in 21st
century. For transition metal catalysts, we have shown that the advances in
nanotechnology enable us to control the concentration of the active sites for
different reaction products by tuning the size and shape of nanoparticles. The
active site for a given product could be a certain atomic arrangement of metal
atoms, the surface atom in a certain oxidation state, bimetallic surface site with
specific composition, or the interface between the metal nanoparticle and the
oxide support. Under reaction conditions, active sites may reconstruct due to the
elevated temperature, or be blocked by strongly adsorbed intermediates. All
these complexities make it essential to develop in situ techniques, and to identify
the active site under reaction conditions.
The development of surface science techniques allows us to study the
molecular factors which control the selectivity of catalytic reactions under
practical reaction conditions. These molecular factors include: 1) Surface
structure (the size and shape of nanoparticles), 2)Surface composition, 3)
Adsorbate mobility, 4) Adsorption-induced surface restructuring, 5) Reaction
intermediates, 6) Surface oxidation state, and 7) Charge transport at metal/
support interfaces.